Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Nov 2;127(45):9465–9472. doi: 10.1021/acs.jpca.3c05021

An Anthracene-Thiolate-Ligated Ruthenium Complex: Computational Insights into Z-Stereoselective Cross Metathesis

Juan Pablo Martínez 1,*, Bartosz Trzaskowski 1,*
PMCID: PMC10658622  PMID: 37916964

Abstract

graphic file with name jp3c05021_0005.jpg

Stereoselective control of the cross metathesis of olefins is a crucial aspect of synthetic procedures. In this study, we utilized density functional theory methods to calculate thermodynamic and kinetic descriptors to explore the stereoselectivity of cross metathesis between allylbenzene and 2-butene-1,4-diyl diacetate. A ruthenium-based complex, characterized primarily by an anthracene-9-thiolate ligand, was designed in silico to completely restrict the E conformation of olefins through a bottom-bound mechanism. Our investigation of the kinetics of all feasible propagation routes demonstrated that Z-stereoisomers of metathesis products can be synthesized with an energy cost of only 13 kcal/mol. As a result, we encourage further research into the synthetic strategies outlined in this work.

1. Introduction

Historically, one of the main challenges in cross metathesis (CM) was the control of the stereoselectivity of the reaction.1 For nonstereoselective catalysts, metathesis products typically consist of a mixture of E and Z isomers, creating operational difficulties due to the complicated and costly nature of component separation.2 In 2009, Schrock et al. synthesized the first Z-selective olefin metathesis catalysts based on Mo3,4 and W5 complexes (see Figure 1). These catalysts were able to achieve yields above 50% with a Z isomer ratio greater than 98%. At that juncture, the field of alkylidene chemistry involving Mo and W had advanced substantially, resulting in the availability of numerous exceptionally active and tunable catalysts.6 In addition, over the past decade, significant progress has been made in the development of Z-selective Mo- and W-based catalysts for CM, which have provided new opportunities for the synthesis of Z-olefins with high selectivity and efficiency.7 These include, for example, a series of halogenated Mo-alkylidene complexes that were reported as exceptionally effective catalysts for olefin CM reactions, resulting in the production of acyclic 1,2-disubstituted Z-alkenyl halides with ratios Z:E of 93:7 or higher.8 Monoaryloxide chloride Mo-based catalysts also exhibited Z-selective CM reactions involving a series of olefins and Z-1,1,1,4,4,4-hexafluoro-2-butene in benzene at 22 °C.9 Furthermore, in the same context of the Mo-catalyzed CM framework, notable advances in the synthesis of Z-trisubstituted alkenes have been recently achieved by Hoveyda et al.1012

Figure 1.

Figure 1

Selected Z-stereoselective catalysts: (a) early Mo- and W-based complexes, (b) set of catalysts designed by Grubbs, and (c) Ru-based complexes ligated to thiolates. (d) Comparison between the side-bound and bottom-bound conformations as a strategy in the design of Z-selective catalysts. (e) Cross metathesis studied in this work.

In the case of Z-selective Ru-based catalysts, a series of functional-group-tolerant systems were reported by Grubbs et al. since 2011.13,14 These catalysts are composed of an N-heterocyclic carbene (NHC) ligand and were derived from the intramolecular carboxylate-driven C–H bond insertion of an N-bound substituent (e.g., adamantyl in Figure 1b).1518 Nitrato-substituted catalysts were found to be more stable and efficient for CM.19 In 2013, Hoveyda et al. used a similar approach to synthesize a Z-selective catalyst, where the replacement of chlorides with a catechodithiolate ligand favored the formation of the Z isomer in ring-opening metathesis polymerization and ring-opening/cross-metathesis.20 In the same year, Jensen et al. achieved Z-selective CM by substituting one chloride of a Ru-based catalyst complex with the bulky 2,4,6-triphenylbenzenethiolate (see Figure 1c).21 This substitution resulted in Z selectivity of up to 96% in the metathesis homocoupling of terminal olefins. Continued investigations in Jensen’s group resulted in additional phosphine-based Z-selective complexes.22 In 2015, within Ru-based catalysts, Hoveyda et al. presented a Ru-disulfide complex to catalyze the synthesis of acyclic Z allylic alcohols via CM transformations, with yields of up to 80% and Z:E ratios as high as 98:2.23 In 2018, Mauduit et al. synthesized a variation of the adamantyl-NHC catalysts, which incorporated N-Dipp (N-2,6-diisopropylphenyl) and N-adamantyl substituents in the nonsaturated NHC ligand.24 This catalyst exhibited exceptional catalytic efficacy in both self- and CM reactions, even at a low catalyst loading, yielding internal olefins with remarkable conversion rates and high Z-selectivity exceeding 99%. Later in 2023, within the Ru-based framework, Mauduit’s group synthesized a series of 16 Z-selective catalysts, and in the case of a catechodithiolate complex, it resulted in a remarkable performance in the Z-selective (>98%) asymmetric ring-opening CM of exonorbornenes.25 Furthermore, the first Z-stereoselective catechodithiolate-based ruthenium complexes containing cyclic(alkyl)(amino)carbene (CAAC) ligands were recently reported by Bertrand and co-workers (see Figure 1c).26 Moderate to good yields and high Z-selectivity (>98%) were obtained in various metathesis transformations, including CM.

Grubbs et al. in collaboration with Houk’s group provided a rationale for the design of Z-stereoselective catalysts by distinguishing between two possible metathesis mechanisms, bottom-bound27 and side-bound28 (see Figure 1d).29 They presented evidence to support the notion that the side-bound mechanism leads to improved d-orbital back-donation and van der Waals interactions between the catalyst and the Z isomer of olefin in transition states, and the steric compression of the olefin substituents is reduced. A possible explanation for the Z-selectivity observed in catechodithiolate-based catalysts may be rationalized in terms of the side-bound mechanism as reported by Houk et al.29 Sterically demanding ligands are known to stabilize the 14-electron Ru-activated catalyst30 and prevent the formation of the E isomer.31,32 However, Z-selective catalysts that incorporate monodentate chelates (e.g., I, (2,4,6-Ph3)Ph-S, or Dipp-O) that allow bottom-bound pathways have also been developed. As a result, the question of whether the Z-stereoselectivity arises entirely from the side-bound mechanism is a primary focus in the current study.

Recent developments in Z-stereoselective olefin metathesis have led to specific applications such as continuous-flow synthesis of natural products, including pheromones and macrocyclic odorant molecules.33 Inspired by the current state of Z-selective olefin metathesis, we aimed to investigate by means of quantum-chemical methods the development of an even more efficient catalyst for stereoselective CM reactions. To achieve this goal, we incorporated a polycyclic aromatic-carbon (anthracene) thiolate ligand to exclusively induce the formation of Z olefins. It may be argued that this framework replicates the design principles of Jensen’s catalyst,21 in which 2,4,6-triphenylbenzenethiolate was incorporated as a ligand that induces the selective formation of Z olefins. Although similarities exist, ligands such as anthracenes or extended polycyclic rings cannot be spatially reoriented so that the formation of E olefins is prohibitive. Therefore, our current research is motivated by the observation that E olefins could still be detected in catalysts with improved Z-selectivity, which were synthesized through the incorporation of Z-inducing ligands such as thiolates,22 sulfonates, and phosphates.34

Accordingly, we chose a model reaction from a set of six metathesis transformations proposed by Grubbs et al. to characterize newly synthesized catalysts,31,35 specifically the CM of allylbenzene (A) and 2-butene-1,4-diyl diacetate (B) to produce 4-phenyl-2-buten-1-yl acetate (P) (see Figure 1e). The objective of our work was to conduct a thorough analysis of the thermodynamic and kinetic factors that determine the formation of the Z isomer of the products as well as to elucidate key strategies to encourage further experimental assessment. In this context, several researchers have previously documented predictive and descriptive catalyses through studies based on chemical calculations and reactivity models, some of them validated by experimental investigations.3640

2. Computational Details

Density functional theory (DFT) calculations were performed using the quantum-chemistry code Jaguar version 11.2.41 Becke’s three-parameter functional combined with the Lee–Yang–Parr correlation functional (B3LYP) was employed to optimize the geometries.42,43 During the geometry optimization process, the dispersion energy corrections developed by Grimme et al. (D3) were incorporated.44 The electronic configuration of the molecular systems was described using the Gaussian 6-31G** basis set, which includes polarization functions for the H, C, N, O, and S atoms. In the case of ruthenium, the small-core, quasi-relativistic effective core potential developed at Los Alamos National Laboratory (LA),45 along with an associated double ζ plus polarization basis set, was used (standard LACVP** keyword in the Jaguar code). Analytical frequency calculations were performed for all localized stationary points at the B3LYP-D3/6-31G**∼LACVP** level of theory. Transition states and connecting local minima were located via linear-transit calculations by using the same DFT method. Electronic energies were obtained from the single-point M06-D3 calculations46 that included a triple ζ basis set plus polarization and diffuse functions, 6-311G++(d,p) (for Ru: standard LACV3P++** keyword in Jaguar). The polarizable continuum model (PCM)47,48 was employed to account for solvent effects, with dichloromethane as the solvent. Gibbs free energies were derived from the corrected electronic energies at the (PCM: CH2Cl2) M06-D3/6-311G++**∼LACV3P++** // B3LYP-D3/6-31G**∼LACVP** level of theory. Corrections for zero-point energy, thermal contributions to internal energy, and entropy term were calculated from vibrational frequencies at 298.15 K, assuming an ideal gas under standard conditions. Standard convergence criteria and a fine grid for DFT calculations were utilized in all cases. It can be argued that the M06 functional already incorporates medium-range dispersion, leading to overestimation of dispersion contributions due to double counting of these effects in M06-D3.49 However, the inclusion of D3 corrections to M06 was demonstrated to enhance the results for catalyzed olefin metathesis, particularly in treating weak interactions.50 Furthermore, the computational approach we chose has previously been benchmarked against experimental data.51,52

3. Results and Discussion

The primary aim of our current work is to demonstrate, via quantum-chemical calculations, the feasibility of designing Z-stereoselective catalysts using the concept of the bottom-bound mechanism. To achieve this goal, we have proposed a Ru-based complex that contains an anionic ligand, anthracene-9-thiolate, which can induce the formation of the Z isomer of olefins through a bottom-bound mechanism. The NHC ligands are constituted of 2,4,6-trimethylphenyl (mesityl) and 2,4-dimethyl-6-(propylthiolate)phenyl groups, where the thiolate group is coordinated with Ru. In this regard, the alkylthiolate moiety was extended until the NHC plane reached a perpendicular alignment with the S–Ru–S axis, resulting in propyl as the alkyl chain. In the context of polycyclic aromatic hydrocarbons, Grela et al. previously synthesized Ru-based catalysts characterized by either a thiophene-based53 or a 10-phenyl-9-phenanthryl54 ligand bonded to an unsymmetrical NHC moiety. In fact, in the case of Hoveyda–Grubbs complexes bearing N-(9-alkylfluorenyl)imidazol-2-ylidene ligands (alkyl: CH3, C2H5, or C6H5), they conducted thorough investigations of olefin metathesis reactions, with a specific focus on achieving Z-selectivity with Z/E ratios of up to 94/6.55 Furthermore, Ru-based catalysts incorporating sulfurated chelates have also been reported by Lemcoff et al.5658 Therefore, we have formulated a possible synthesis route to the precatalyst in question (details in Figure S1 in the Supporting Information), which involves the preparation of thioether-imidazolium chlorides,59 anilines,60,61 and modifications to the Grubbs first-generation catalyst.34,62

3.1. The Initiation Route

The catalytic cycle commences with the generation of an active catalyst, ACPh, as a result of the dissociation of pyridine from pre. Two distinct conformations pre and pre′ are presented in Figure 2, which are differentiated by the orientation of benzylidene relative to the propyl-thiolate of NHC. The Ru-pyridine bond dissociation reported in Figure S2 in the Supporting Information reveals a flat potential (electronic) energy surface that reaches maxima of 16 and 21 kcal/mol for pre and pre′, respectively. Nonetheless, for the localized stationary points, the thermal contributions resulted in a Gibbs activation energy of <15 kcal/mol, as shown in Figure 2. As a result, activation of the catalyst is expected to take place under mild conditions at room temperature. The subsequent step involves olefin coordination (OC), and given our interest in synthesizing 4-phenyl-2-buten-1-yl acetate P through a productive metathesis reaction, there exist two possible sequences of the cycle: allylbenzene A association followed by propagation with 2-butene-1,4-diyl diacetate B, as illustrated in Figure 1e, or vice versa. Nevertheless, ACPh is expected to exhibit a preference for reacting with allylbenzene initially, owing to the greater reactivity of less substituted alkenes such as olefin A,63 as confirmed in our earlier work.64 As a result, we focus only on the initiation route through A.

Figure 2.

Figure 2

(Top) Gibbs free energy changes for catalyst activation through the dissociation of pyridine. (Bottom) Conformers of the activated catalyst coordinated to allylbenzene (OCA) and Gibbs free energy comparisons relative to OCA1. Inset: The activation-strain model for the 2,2-cycloaddtion transition state (TS1A). All energies are given in kcal/mol.

Our analysis and energy comparisons were conducted starting from the olefin coordination intermediate OCA. We identified eight distinct configurations for the coordination of A with the active catalyst ACPh (or AC′Ph in which benzylidene is trans to propyl-thiolate), as depicted in Figure 2. The phenyl group in A can acquire either a contracted or an expanded conformation. Similarly to our previous research findings, it has been demonstrated that the destabilization of OC intermediates is correlated with the destabilization of transition states.65 Therefore, we investigated the initiation phase exclusively for the most stable conformers. Consequently, we suggest the criterion of ΔGrel < 3 kcal/mol to select the intermediates that undergo a kinetically favored initiation phase, which in this study are all OCA intermediates characterized by the expanded configuration of phenyl in A. This thermodynamic stabilization can be attributed to a more active alkene orientation, as described by Straub.66

The activated complex pyr/ACPh or pyr/AC′Ph is isoenergetic with respect to OCA since we did not observe significant energy variations for the exchange of pyridine with olefin AG = 0.4 and −1.2 kcal/mol respectively; see Figure 3). This fact indicates that catalyst activation is susceptible to reversibility when the concentration of A is insufficient. The 2,2-cycloaddition from OCA via transition state TS1A produces metallacyclobutane MCBA. The rupture of MCBA through TS2A leads to the 2,2-cycloreversion intermediate CRA, followed by the release of the propagating activated catalyst ACA (or AC′A in which alkylidene is trans to propyl-thiolate) and the corresponding olefin. Species OCA1 and OCA5 are associated with productive metathesis so that styrene (S0) is generated through the initiation route. On the other hand, OCA3 and OCA7 are related to nonproductive metathesis, so that substrate S1 (1-propene-1,3-diphenyl) is instead generated. The Gibbs free energy profiles, ΔG, for the initiation phase via OCA1 and OCA5 (productive metathesis) are depicted in Figure 3, along with the reaction energies and barriers for the 2,2-cycloaddition steps (ΔG1 and ΔG1) and 2,2-cycloreversion steps (ΔG2 and ΔG2).

Figure 3.

Figure 3

Gibbs free energy profiles for the initiation phase for the most stable OCA species associated with productive metathesis. Illustration and energies (kcal/mol) are relative to OCA1. The equivalent schematization regarding the benzylidene of AC′Ph localized in the opposite direction corresponds to path A5. ΔG3 = G(TS2) – G(OC).

To establish the kinetic bottleneck of the reaction, we evaluated all energy barriers in relation to the lowest-energy intermediate and highest-energy transition state, a concept that is referred as the energetic span model.6769 That is, the third energy barrier reported in Figure 3 is calculated as ΔG3 = G(TS2A) – G(OCA) for each pathway. The ΔG3 value for path A1 is greater than the ΔG1 and ΔG2 values, indicating that the critical states of this reaction are OCA1 and TS2A1. In the case of path A5, ΔG2 is slightly greater than ΔG3; therefore, the crucial states of this pathway are MCBA5 and TS2A5 describing the 2,2-cycloreversion step. As indicated in Figure 3, our findings suggest that the catalyst can be promptly activated toward productive metathesis through pathways A1 and A5, which have overall costs of 8.3 and 6.1 kcal/mol, respectively. The generation of the active catalysts, AC′A and ACA, through the release of styrene via paths A1 and A5, respectively, led to favorable driving forces of 6.5 and 4.5 kcal/mol, which were computed using the most stable isolated fragments: G(AC′A) or G(ACA) + G(styrene) – G(OCA1). In the cases of AC′A and ACA, we optimized eight conformers and chose the most thermodynamically stable structures, as illustrated in Figure S3 in the Supporting Information. In contrast, nonproductive routes A3 and A7 are comparatively less kinetically competitive due to limiting initiation barriers higher than 10 kcal/mol. Furthermore, the formation of substrate S1 is endergonic, unlike the exergonic initiation through the release of styrene (see Figure S4 in the Supporting Information for a detailed description of the corresponding energy profiles).

As a final point, we employed the activation-strain model strategy7073 to assess the relative energy of the system in the 2,2-cycloaddition transition states, considering the structural strain and intermolecular interactions involved. The strain energy quantifies the amount of energy required to alter the geometry of both olefin A and active catalyst ACPh (or AC′Ph) to form the structure of TS1A, while the interaction energy assesses the extent of interactions between the distorted fragments when they form TS1A. Similarly to our previous work,64 the overall cost ΔG3 for the initiation route is linearly correlated (R2 = 0.77) with the level of structural strain and, to some extent, the molecular interactions (R2 = 0.53, see inset in Figure 2). Consequently, faster reaction rates are associated with a lower degree of structural strain in the reacting entities as well as to moderate molecular interactions that prevent the formation of highly stabilized metallacyclobutanes, thereby promoting the course of the reaction.

3.2. The Propagation Route

In relation to the process of productive metathesis, the active catalysts AC′A and ACA are subsequently coordinated with compound B to form the propagating OCP species. However, the flexibility of the structure of olefin B results in the formation of several conformers of OCP (the selection of OCP species is reported in Figure S5 in the Supporting Information). Similarly to the preceding subsection, our analysis of the propagation phase focused exclusively on the most stable conformers, specifically, OCP2 and OCP3. The corresponding reaction mechanisms are illustrated in Figure 4.

Figure 4.

Figure 4

Gibbs free energy profiles (in kcal/mol) for the propagation phase through 2-butene-1,4-diyl diacetate as the substrate (B) and the catalyst activated via the metathesis of allylbenzene (AC′A).

The Gibbs free energy profiles for the propagation phase through OCP2,3 are reported in relation to the sum of the energies of AC′A and B. To obtain the energy of B, we selected the most stable structure from a group of three cis conformers (see Figure S3 in the Supporting Information for details). For example, the formation of OCP2 resulted in no energy change (ΔG = 0.1 kcal/mol) and the 2,2-cycloaddition energy barrier is ΔG1 = 10.9 kcal/mol. MCBP2 is destabilized since ΔG1 = 9.0 kcal/mol, but the 2,2-cycloreversion proceeds with an energy barrier of only ΔG2 = 4.1 kcal/mol under an exergonic process of ΔG2 = −3.4 kcal/mol. Olefin decoordination from CRP2 resulted in a favorable driving force of 4.2 kcal/mol, giving rise to the acetate propagating alkylidene (ACP) and the metathesis product P. The OCP2 and TS2P2 stationary points are the rate-limiting states in this part of the pathway, resulting in an overall cost of ΔG3 = 13.1 kcal/mol, so that the reaction can be carried out under laboratory conditions. In our prior communication,64 we reported equivalent energy barriers for a second-generation Grubbs catalyst, so that a similar catalytic activity is expected for the complex in question. Path P3 is kinetically less competitive but accessible since we calculated ΔG3 = 22.1 kcal/mol as the limiting barrier. Furthermore, if the phenyl-alkylidene is located in the opposite direction relative to AC′A (that is, ACA in which the phenyl-alkylidene is cis to propyl-thiolate), the propagating structure OCP6, which is analogous to OCP2, was destabilized by 6 kcal/mol compared to OCP2. Nonetheless, the overall energy required for the formation of P via path P6 was calculated to be only 9 kcal/mol, as illustrated in Figure S6 and discussed in the Supporting Information. Consequently, the catalyst in question is versatile, which means that metathesis products can be generated regardless of the orientation of the phenyl-alkylidene within the activated catalyst. Our calculations indicate that Z-stereoselective catalysts can be effectively designed by considering the bottom-bound mechanism, making our study a reference point for future synthetic procedures as an alternative to strategies based on the side-bound mechanism.

4. Conclusions

We performed predictive DFT calculations to assess the catalytic activity of a Ru-based complex containing an anthracene-9-thiolate ligand. The catalyst was designed in silico to induce the formation of the Z isomer of metathesis products through the bottom-bound mechanism. Additionally, the anthracene-9-thiolate ligand impedes the formation of the E isomer. Through an in-depth analysis of the thermodynamics and kinetics of the formulated mechanisms, we explored the potential applications of the Ru complex in the CM of allylbenzene and 2-butene-1,4-diyl diacetate. Our results indicate that this new catalyst should facilitate efficient catalyst activation, as the energy barrier for the initiation route was <15 kcal/mol essentially attributed to the Ru-pyridine bond dissociation. Furthermore, we demonstrated that the overall energy cost for olefin metathesis through the propagation phase, which leads to the synthesis of the Z isomer of the metathesis product, 4-phenyl-2-buten-1-yl acetate, was only 13 kcal/mol. Our findings hold significant promise for the future exploration of this new catalyst candidate through experimental studies, and we hope that they will serve as a useful guide for further computational investigations aimed at developing new Z-stereoselective catalysts. In this regard, we additionally proposed a laboratory procedure for the synthesis of this novel catalyst.

Acknowledgments

J.P.M. thanks a postdoctoral fellowship granted by The Polish National Agency for Academic Exchange (Narodowa Agencja Wymiany Akademickiej) under the ULAM program 2021, in accordance with Decision No. PPN/ULM/2020/1/00117/DEC/01 and Agreement No. PPN/ULM/2020/1/00117/U/00001 of 2021-01-11. B.T. and J.P.M. thank National Science Centre (Poland) grants Opus 22, UMO-2021/43/B/ST4/00122 and Opus 15, UMO-2018/29/B/ST4/00805 for generous funding.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpca.3c05021.

  • Discussion of alternative initiation and propagation phases; detailed analysis of molecular structures and energies of the investigated chemical species and their structural isomers, along with their respective DFT absolute energies (PDF)

  • Computed molecular Cartesian coordinates in a format convenient for visualization (XYZ)

Author Contributions

J.P.M. and B.T. contributed equally to the work.

The authors declare no competing financial interest.

Supplementary Material

jp3c05021_si_001.pdf (1.1MB, pdf)
jp3c05021_si_002.xyz (350.1KB, xyz)

References

  1. Hamilton J. G.; Ivin K. J.; McCann G. M.; Rooney J. J. The Origins of Stereoselectivities in olefin Metathesis. J. Chem. Soc. Chem. Commun. 1984, 1379–1381. 10.1039/c39840001379. [DOI] [Google Scholar]
  2. Fürstner A. Teaching Metathesis “Simple” Stereochemistry. Science 2013, 341, 1229713. 10.1126/science.1229713. [DOI] [PubMed] [Google Scholar]
  3. Ibrahem I.; Yu M.; Schrock R. R.; Hoveyda A. H. Highly Z- and Enantioselective Ring-Opening/Cross-Metathesis Reactions Catalyzed by Stereogenic-at-Mo Adamantylimido Complexes. J. Am. Chem. Soc. 2009, 131, 3844–3845. 10.1021/ja900097n. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Flook M. M.; Jiang A. J.; Schrock R. R.; Müller P.; Hoveyda A. H. Z-Selective olefin Metathesis Processes Catalyzed by a Molybdenum Hexaisopropylterphenoxide Monopyrrolide Complex. J. Am. Chem. Soc. 2009, 131, 7962–7963. 10.1021/ja902738u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Jiang A. J.; Zhao Y.; Schrock R. R.; Hoveyda A. H. Highly Z-Selective Metathesis homocoupling of Terminal Olefins. J. Am. Chem. Soc. 2009, 131, 16630–16631. 10.1021/ja908098t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Schrock R. R. Recent Advances in High Oxidation State Mo and W Imido alkylidene Chemistry. Chem. Rev. 2009, 109, 3211–3226. 10.1021/cr800502p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Dawood K. M.; Nomura K. Recent Developments in Z-Selective olefin Metathesis Reactions by Molybdenum, Tungsten, Ruthenium, and Vanadium Catalysts. Adv. Synth. Catal. 2021, 363, 1970–1997. 10.1002/adsc.202001117. [DOI] [Google Scholar]
  8. Koh M. J.; Nguyen T. T.; Zhang H.; Schrock R. R.; Hoveyda A. H. Direct Synthesis of Z-Alkenyl Halides through Catalytic Cross-Metathesis. Nature 2016, 531, 459–465. 10.1038/nature17396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Koh M. J.; Nguyen T. T.; Lam J. K.; Torker S.; Hyvl J.; Schrock R. R.; Hoveyda A. H. Molybdenum Chloride Catalysts for Z-Selective olefin Metathesis Reactions. Nature 2017, 542, 80–85. 10.1038/nature21043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Nguyen T. T.; Koh M. J.; Mann T. J.; Schrock R. R.; Hoveyda A. H. Synthesis of E- and Z-Trisubstituted alkenes by Catalytic Cross-Metathesis. Nature 2017, 552, 347–354. 10.1038/nature25002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Qin C.; Koengeter T.; Zhao F.; Mu Y.; Liu F.; Houk K. N.; Hoveyda A. H. Z-Trisubstituted α,β-Unsaturated Esters and Acid Fluorides through Stereocontrolled Catalytic Cross-Metathesis. J. Am. Chem. Soc. 2023, 145, 3748–3762. 10.1021/jacs.2c13245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Koengeter T.; Qin C.; Mai B. K.; Liu Q.; Mu Y.; Liu P.; Hoveyda A. H. Catalytic Cross-Metathesis Reactions That Afford E- and Z-Trisubstituted Alkenyl Bromides: Scope, Applications, and Mechanistic Insights. J. Am. Chem. Soc. 2023, 145, 3774–3785. 10.1021/jacs.2c13289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Herbert M. B.; Grubbs R. H. Z-Selective Cross Metathesis with Ruthenium Catalysts: Synthetic Applications and Mechanistic Implications. Angew. Chem., Int. Ed. 2015, 54, 5018–5024. 10.1002/anie.201411588. [DOI] [PubMed] [Google Scholar]
  14. Montgomery T. P.; Johns A. M.; Grubbs R. H. Recent Advancements in Stereoselective olefin Metathesis Using Ruthenium Catalysts. Catalysts 2017, 7, 87. 10.3390/catal7030087. [DOI] [Google Scholar]
  15. Endo K.; Grubbs R. H. Chelated Ruthenium Catalysts for Z-Selective olefin Metathesis. J. Am. Chem. Soc. 2011, 133, 8525–8527. 10.1021/ja202818v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Keitz B. K.; Endo K.; Patel P. R.; Herbert M. B.; Grubbs R. H. Improved Ruthenium Catalysts for Z-Selective olefin Metathesis. J. Am. Chem. Soc. 2012, 134, 693–699. 10.1021/ja210225e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Herbert M. B.; Suslick B. A.; Liu P.; Zou L.; Dornan P. K.; Houk K. N.; Grubbs R. H. Cyclometalated Z-Selective Ruthenium Metathesis Catalysts with Modified N-Chelating Groups. Organometallics 2015, 34, 2858–2869. 10.1021/acs.organomet.5b00185. [DOI] [Google Scholar]
  18. Cannon J. S.; Zou L.; Liu P.; Lan Y.; O’leary D. J.; Houk K. N.; Grubbs R. H. Carboxylate-Assisted C(Sp3)-H Activation in olefin Metathesis-Relevant Ruthenium Complexes. J. Am. Chem. Soc. 2014, 136, 6733–6743. 10.1021/ja5021958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Herbert M. B.; Marx V. M.; Pederson R. L.; Grubbs R. H. Concise Syntheses of Insect Pheromones Using Z-Selective Cross Metathesis. Angew. Chem., Int. Ed. 2013, 52, 310–314. 10.1002/anie.201206079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Khan R. K. M.; Torker S.; Hoveyda A. H. Readily Accessible and Easily Modifiable Ru-Based Catalysts for Efficient and Z-Selective Ring-Opening Metathesis Polymerization and Ring-Opening/Cross-Metathesis. J. Am. Chem. Soc. 2013, 135, 10258–10261. 10.1021/ja404208a. [DOI] [PubMed] [Google Scholar]
  21. Occhipinti G.; Hansen F. R.; Törnroos K. W.; Jensen V. R. Simple and Highly Z-Selective Ruthenium-Based olefin Metathesis Catalyst. J. Am. Chem. Soc. 2013, 135, 3331–3334. 10.1021/ja311505v. [DOI] [PubMed] [Google Scholar]
  22. Smit W.; Koudriavtsev V.; Occhipinti G.; Törnroos K. W.; Jensen V. R. Phosphine-Based Z-Selective Ruthenium olefin Metathesis Catalysts. Organometallics 2016, 35, 1825–1837. 10.1021/acs.organomet.6b00214. [DOI] [Google Scholar]
  23. Koh M. J.; Khan R. K. M.; Torker S.; Yu M.; Mikus M. S.; Hoveyda A. H. High-Value Alcohols and Higher-Oxidation-State Compounds by Catalytic Z-Selective Cross-Metathesis. Nature 2015, 517, 181–186. 10.1038/nature14061. [DOI] [PubMed] [Google Scholar]
  24. Dumas A.; Tarrieu R.; Vives T.; Roisnel T.; Dorcet V.; Baslé O.; Mauduit M. A Versatile and Highly Z -Selective olefin Metathesis Ruthenium Catalyst Based on a Readily Accessible N-Heterocyclic carbene. ACS Catal. 2018, 8, 3257–3262. 10.1021/acscatal.8b00151. [DOI] [Google Scholar]
  25. Morvan J.; Bouëtard D.; Kong L.; Chou Y.; Vives T.; Albalat M.; Roisnel T.; Crévisy C.; Nava P.; Humbel S.; Vanthuyne N.; Clavier H.; Mauduit M. Chiral Atropisomeric-NHC Catechodithiolate Ruthenium Complexes for Z-Selective Asymmetric Ring-Opening Cross Metathesis of Exo-Norbornenes. Chem.—Eur. J. 2023, 29, e202300341 10.1002/chem.202300341. [DOI] [PubMed] [Google Scholar]
  26. Morvan J.; Vermersch F.; Lorkowski J.; Talcik J.; Vives T.; Roisnel T.; Crévisy C.; Vanthuyne N.; Bertrand G.; Jazzar R.; Mauduit M. Cyclic(Alkyl)(Amino)carbene Ruthenium Complexes for Z-Stereoselective (Asymmetric) olefin Metathesis. Catal. Sci. Technol. 2023, 13, 381–388. 10.1039/D2CY01795D. [DOI] [Google Scholar]
  27. Tallarico J. A.; Bonitatebus P. J.; Snapper M. L. Ring-Opening Metathesis. A Ruthenium Catalyst Caught in the Act. J. Am. Chem. Soc. 1997, 119, 7157–7158. 10.1021/ja971285r. [DOI] [Google Scholar]
  28. Trnka T. M.; Day M. W.; Grubbs R. H. Novel η3-Vinylcarbene Complexes Derived from Ruthenium-Based olefin Metathesis Catalysts. Organometallics 2001, 20, 3845–3847. 10.1021/om010314a. [DOI] [Google Scholar]
  29. Liu P.; Xu X.; Dong X.; Keitz B. K.; Herbert M. B.; Grubbs R. H.; Houk K. N. Z-Selectivity in olefin Metathesis with Chelated Ru Catalysts: Computational Studies of Mechanism and Selectivity. J. Am. Chem. Soc. 2012, 134, 1464–1467. 10.1021/ja2108728. [DOI] [PubMed] [Google Scholar]
  30. Jawiczuk M.; Marczyk A.; Trzaskowski B. Decomposition of Ruthenium olefin Metathesis Catalyst. Catalysts 2020, 10, 887. 10.3390/catal10080887. [DOI] [Google Scholar]
  31. Ogba O. M.; Warner N. C.; O’Leary D. J.; Grubbs R. H. Recent Advances in Ruthenium-Based olefin Metathesis. Chem. Soc. Rev. 2018, 47, 4510–4544. 10.1039/C8CS00027A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Díez-González S.; Nolan S. P. Stereoelectronic Parameters Associated with N-Heterocyclic carbene (NHC) Ligands: A Quest for Understanding. Coord. Chem. Rev. 2007, 251, 874–883. 10.1016/j.ccr.2006.10.004. [DOI] [Google Scholar]
  33. Morvan J.; McBride T.; Curbet I.; Colombel-Rouen S.; Roisnel T.; Crévisy C.; Browne D. L.; Mauduit M. Continuous Flow Z-Stereoselective olefin Metathesis: Development and Applications in the Synthesis of Pheromones and macrocyclic Odorant Molecules. Angew. Chem., Int. Ed. 2021, 60, 19685–19690. 10.1002/anie.202106410. [DOI] [PubMed] [Google Scholar]
  34. Teo P.; Grubbs R. H. Facile Synthesis of Effcient and Selective Ruthenium olefin Metathesis Catalysts with Sulfonate and Phosphate Ligands. Organometallics 2010, 29, 6045–6050. 10.1021/om1007924. [DOI] [Google Scholar]
  35. Ritter T.; Hejl A.; Wenzel A. G.; Funk T. W.; Grubbs R. H. A Standard System of Characterization for olefin Metathesis Catalysts. Organometallics 2006, 25, 5740–5745. 10.1021/om060520o. [DOI] [Google Scholar]
  36. Pump E.; Poater A.; Bahri-Laleh N.; Credendino R.; Serra L.; Scarano V.; Cavallo L. Regio, Stereo and Chemoselectivity of 2nd Generation Grubbs Ruthenium-Catalyzed olefin Metathesis. Catal. Today 2020, 388–389, 394–402. 10.1016/j.cattod.2020.04.071. [DOI] [Google Scholar]
  37. Bahri-Laleh N.; Credendino R.; Cavallo L. The Intriguing Modeling of Cis-Trans Selectivity in Ruthenium-Catalyzed olefin Metathesis. Beilstein J. Org. Chem. 2011, 7, 40–45. 10.3762/bjoc.7.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Luján C.; Nolan S. P. E/Z Selectivity in Ruthenium-Mediated Cross Metathesis. Catal. Sci. Technol. 2012, 2, 1027–1032. 10.1039/c2cy00457g. [DOI] [Google Scholar]
  39. Rouen M.; Queval P.; Borré E.; Falivene L.; Poater A.; Berthod M.; Hugues F.; Cavallo L.; Baslé O.; Olivier-Bourbigou H.; Mauduit M. Selective Metathesis of α-Olefins from Bio-Sourced Fischer–Tropsch Feeds. ACS Catal. 2016, 6, 7970–7976. 10.1021/acscatal.6b01428. [DOI] [Google Scholar]
  40. Müller D. S.; Curbet I.; Raoul Y.; Le Nôtre J.; Baslé O.; Mauduit M. Stereoretentive olefin Metathesis Made Easy: In Situ Generation of Highly Selective Ruthenium Catalysts from Commercial Starting Materials. Org. Lett. 2018, 20, 6822–6826. 10.1021/acs.orglett.8b02943. [DOI] [PubMed] [Google Scholar]
  41. Bochevarov A. D.; Harder E.; Hughes T. F.; Greenwood J. R.; Braden D. A.; Philipp D. M.; Rinaldo D.; Halls M. D.; Zhang J.; Friesner R. A. Jaguar: A High-Performance Quantum Chemistry Software Program with Strengths in Life and Materials Sciences. Int. J. Quantum Chem. 2013, 113, 2110–2142. 10.1002/qua.24481. [DOI] [Google Scholar]
  42. Becke A. D. III The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648. 10.1063/1.464913. [DOI] [Google Scholar]
  43. Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
  44. Goerigk L.; Grimme S. A Thorough Benchmark of Density Functional Methods for General Main Group Thermochemistry, Kinetics, and Noncovalent Interactions. Phys. Chem. Chem. Phys. 2011, 13, 6670–6688. 10.1039/c0cp02984j. [DOI] [PubMed] [Google Scholar]
  45. Hay P. J.; Wadt W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299. 10.1063/1.448975. [DOI] [Google Scholar]
  46. Zhao Y.; Truhlar D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215–241. 10.1007/s00214-007-0310-x. [DOI] [Google Scholar]
  47. Tomasi J.; Mennucci B.; Cammi R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
  48. Tomasi J.; Persico M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 94, 2027–2094. 10.1021/cr00031a013. [DOI] [Google Scholar]
  49. Goerigk L. Treating London-Dispersion Effects with the Latest Minnesota Density Functionals: Problems and Possible Solutions. J. Phys. Chem. Lett. 2015, 6, 3891–3896. 10.1021/acs.jpclett.5b01591. [DOI] [PubMed] [Google Scholar]
  50. Minenkov Y.; Singstad Å.; Occhipinti G.; Jensen V. R. The Accuracy of DFT-Optimized Geometries of Functional Transition Metal Compounds: A Validation Study of Catalysts for olefin Metathesis and Other Reactions in the Homogeneous Phase. Dalton Trans. 2012, 41, 5526–5541. 10.1039/c2dt12232d. [DOI] [PubMed] [Google Scholar]
  51. Martínez J. P.; Trzaskowski B. Electrophilicity of Hoveyda-Grubbs olefin Metathesis Catalysts as the Driving Force That Controls Initiation Rates. ChemPhysChem 2022, e202200580 10.1002/cphc.202200580. [DOI] [PubMed] [Google Scholar]
  52. Małecki P.; Gajda K.; Gajda R.; Woźniak K.; Trzaskowski B.; Kajetanowicz A.; Grela K. Specialized Ruthenium olefin Metathesis Catalysts Bearing Bulky Unsymmetrical NHC Ligands: Computations, Synthesis, and Application. ACS Catal. 2019, 9, 587–598. 10.1021/acscatal.8b04783. [DOI] [Google Scholar]
  53. Smoleń M.; Kośnik W.; Gajda R.; Woźniak K.; Skoczeń A.; Kajetanowicz A.; Grela K. Ruthenium Complexes Bearing Thiophene-Based Unsymmetrical N-Heterocyclic carbene Ligands as Selective Catalysts for olefin Metathesis in Toluene and Environmentally Friendly 2-Methyltetrahydrofuran. Chem.—Eur. J. 2018, 24, 15372–15379. 10.1002/chem.201803460. [DOI] [PubMed] [Google Scholar]
  54. Dąbrowski M.; Wyrębek P.; Trzybiński D.; Woźniak K.; Grela K. In a Quest for Selectivity Paired with Activity: A Ruthenium olefin Metathesis Catalyst Bearing an Unsymmetrical Phenanthrene-Based N-Heterocyclic carbene. Chem.—Eur. J. 2020, 26, 3782–3794. 10.1002/chem.201904549. [DOI] [PubMed] [Google Scholar]
  55. Gajda K.; Sytniczuk A.; Vendier L.; Trzaskowski B.; Lugan N.; Kajetanowicz A.; Bastin S.; Grela K.; César V. Towards (C,C)-Cyclometalated N-(9-Alkylfluorenyl)NHC Ruthenium Complexes for Z-Selective olefin Metathesis. Eur. J. Inorg. Chem. 2023, 26, e202300169 10.1002/ejic.202300169. [DOI] [Google Scholar]
  56. Nechmad N. B.; Phatake R.; Ivry E.; Poater A.; Lemcoff N. G. Unprecedented Selectivity of Ruthenium Iodide Benzylidenes in olefin Metathesis Reactions. Angew. Chem., Int. Ed. 2020, 59, 3539–3543. 10.1002/anie.201914667. [DOI] [PubMed] [Google Scholar]
  57. Nechmad N. B.; Kobernik V.; Tarannam N.; Phatake R.; Eivgi O.; Kozuch S.; Lemcoff N. G. Reactivity and Selectivity in Ruthenium Sulfur-Chelated Diiodo Catalysts. Angew. Chem., Int. Ed. 2021, 60, 6372–6376. 10.1002/anie.202014929. [DOI] [PubMed] [Google Scholar]
  58. Phatake R. S.; Nechmad N. B.; Reany O.; Lemcoff N. G. Highly Substrate-Selective macrocyclic Ring Closing Metathesis. Adv. Synth. Catal. 2022, 364, 1465–1472. 10.1002/adsc.202101515. [DOI] [Google Scholar]
  59. Kuriyama M.; Shimazawa R.; Shirai R. Design and Synthesis of Thioether-Imidazolium Chlorides as Efficient Ligands for Palladium-Catalyzed Suzuki–Miyaura Coupling of Aryl Bromides with Arylboronic Acids. Tetrahedron 2007, 63, 9393–9400. 10.1016/j.tet.2007.06.097. [DOI] [Google Scholar]
  60. Schmidt V. A. Reactions for Making Widely Used Aniline Compounds Break Norms of Synthesis. Nature 2020, 584, 46–47. 10.1038/d41586-020-02283-8. [DOI] [PubMed] [Google Scholar]
  61. Min Z.; Ying-Guo F. An Efficient Monochlorination of Electron-Rich Aromatic Compounds Catalysed by Ammonium Iodide. J. Chem. Res. 2014, 38, 197–199. 10.3184/174751914X13929076043809. [DOI] [Google Scholar]
  62. Schwab P.; France M. B.; Ziller J. W.; Grubbs R. H. A Series of Well-Defined Metathesis Catalysts–Synthesis of [RuCl2(=CHR′)(PR3)2] and Its Reactions. Angew. Chem., Int. Ed. 1995, 34, 2039–2041. 10.1002/anie.199520391. [DOI] [Google Scholar]
  63. Chatterjee A. K.; Choi T. L.; Sanders D. P.; Grubbs R. H. A General Model for Selectivity in olefin Cross Metathesis. J. Am. Chem. Soc. 2003, 125, 11360–11370. 10.1021/ja0214882. [DOI] [PubMed] [Google Scholar]
  64. Martínez J. P.; Trzaskowski B. On the stereoselectivity of the Cross Metathesis of Olefins Catalyzed by a Second-Generation Catalyst. Catal. Commun. 2022, 172, 106552 10.1016/j.catcom.2022.106552. [DOI] [Google Scholar]
  65. Martínez J. P.; Trzaskowski B. olefin Metathesis Catalyzed by a Hoveyda-Grubbs-like Complex Chelated to Bis(2-Mercaptoimidazolyl) Methane: A Predictive DFT Study. J. Phys. Chem. A 2022, 126, 720–732. 10.1021/acs.jpca.1c09336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Straub B. F. Origin of the High Activity of Second-Generation Grubbs Catalysts. Angew. Chem., Int. Ed. 2005, 44, 5974–5978. 10.1002/anie.200501114. [DOI] [PubMed] [Google Scholar]
  67. Uhe A.; Kozuch S.; Shaik S. Automatic Analysis of Computed Catalytic Cycles. J. Comput. Chem. 2011, 32, 978–985. 10.1002/jcc.21669. [DOI] [PubMed] [Google Scholar]
  68. Kozuch S.; Martin J. M. L. The Rate-Determining Step Is Dead. Long Live the Rate-Determining State!. ChemPhysChem 2011, 12, 1413–1418. 10.1002/cphc.201100137. [DOI] [PubMed] [Google Scholar]
  69. Kozuch S.; Shaik S. How to Conceptualize Catalytic Cycles? The Energetic Span Model. Acc. Chem. Res. 2011, 44, 101–110. 10.1021/ar1000956. [DOI] [PubMed] [Google Scholar]
  70. Fernández I.; Bickelhaupt F. M. The Activation Strain Model and Molecular Orbital Theory: Understanding and Designing Chemical Reactions. Chem. Soc. Rev. 2014, 43, 4953–4967. 10.1039/C4CS00055B. [DOI] [PubMed] [Google Scholar]
  71. Ess D. H.; Houk K. N. Distortion/Interaction Energy Control of 1,3-Dipolar Cycloaddition Reactivity. J. Am. Chem. Soc. 2007, 129, 10646–10647. 10.1021/ja0734086. [DOI] [PubMed] [Google Scholar]
  72. van Zeist W. J.; Bickelhaupt F. M. The Activation Strain Model of Chemical Reactivity. Org. Biomol. Chem. 2010, 8, 3118–3127. 10.1039/b926828f. [DOI] [PubMed] [Google Scholar]
  73. Bickelhaupt F. M.; Houk K. N. Analyzing Reaction Rates with the Distortion/Interaction-Activation Strain Model. Angew. Chem., Int. Ed. 2017, 56, 10070–10086. 10.1002/anie.201701486. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

jp3c05021_si_001.pdf (1.1MB, pdf)
jp3c05021_si_002.xyz (350.1KB, xyz)

Articles from The Journal of Physical Chemistry. a are provided here courtesy of American Chemical Society

RESOURCES