Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

A comprehensive mechanistic investigation of sustainable carbene N–H insertion catalyzed by engineered His-ligated heme proteins

Rahul L. Khade a, Ronald Daisuke Adukure b, Xinyi Zhao a, Carolyn Wang a, Rudi Fasan *b and Yong Zhang *a
aDepartment of Chemistry and Chemical Biology, Stevens Institute of Technology, 1 Castle Point Terrace, Hoboken, NJ 07030, USA. E-mail: yong.zhang@stevens.edu
bDepartment of Chemistry and Biochemistry, University of Texas at Dallas, Richardson, TX 75080, USA. E-mail: rudi.fasan@utdallas.edu

Received 15th August 2024 , Accepted 10th January 2025

First published on 13th January 2025


Abstract

Engineered heme proteins possess excellent biocatalytic carbene N–H insertion abilities for sustainable synthesis, and most of them have His as the Fe axial ligand. However, information on the basic reaction mechanisms is limited, and ground states of heme carbenes involved in the prior computational mechanistic studies are under debate. A comprehensive quantum chemical reaction pathway study was performed for the heme model with a His analogue as the axial ligand and carbene from the widely used precursor ethyl diazoacetate with aniline as the substrate. The ground state of this heme carbene was calculated by the high-level complete active space self-consistent field (CASSCF) approach, which shows a closed-shell singlet that is consistent with many experimental works. Based on this, DFT calculations of ten main reaction pathways were compared. Results showed that the most favorable pathway involved the initial formation of the metal-bound ylide, followed by a concerted rearrangement/dissociation transition state to form the free enol, which then underwent a water-assisted proton transfer process to yield the final N–H insertion product. This computational prediction was validated via new experimental data using His-ligated myoglobin variants with different types of carbenes. Overall, this is the first comprehensive computational mechanistic study of heme carbene N–H insertions, particularly for neutral His ligated heme proteins and the first high-level CASSCF confirmation of the ground state of the used heme carbene. The experimental results are also the first in this field. Overall, these results build a solid basis for the proposed reaction mechanism to facilitate future biocatalytic carbene N–H insertion studies.


1. Introduction

Carbenoid N–H insertion is a powerful approach to provide a concise synthetic route for α-amino acid derivatives, α-amino ketones, alkaloids, N-heterocyclic compounds, and other bioactive molecules.1–17 It has also been used for bioconjugation purposes.18–21 Initially, iron porphyrins were found and further developed to be used as effective catalysts for this reaction.22–26 More recently, engineered Fe-containing heme proteins, such as cytochrome P450/P411,27–29 myoglobin (Mb),30–35 cytochrome c,36,37 YfeX (peroxidase),38 protoglobin,39 and neuroglobin,40 were explored in a wide range of carbenoid N–H insertions. Clearly, except for P450/P411s, all other experimental studies, which consist of ∼80% work in this field, employed His-ligated heme protein variants (mostly Mb-based), which exhibit promising biocatalytic results with up to 99% yield32,34,40 and excellent chemoselectivity.30,33,36 These His-ligated heme biocatalysts can work under aerobic conditions with reductants31,35 and be extended to aliphatic amine substrates besides aromatic amines33 and effect asymmetric N–H insertions34 with up to 99% ee.37 In addition, all these engineered heme biocatalysts work under ambient conditions such as room temperature and aqueous solutions. The features of biocompatibility and no toxicity of these biocatalytic proteins along with the use of the most abundant and inexpensive transition metal (Fe) support their potential use in sustainable chemical synthesis for new C–N bonds.

Compared to the above-mentioned large body of experimental advances in this area, computational mechanistic work is relatively scarce.28,29,36,37,40–42 While the protein environment effects on chemoselectivity36 and stereoselectivity28,29,37 have been well elucidated computationally, there is only a small set of basic reaction mechanism information of heme carbene N–H insertions, for which density functional theory (DFT) studies have been employed to understand these Fe(II)-based biocatalytic transformations. In 2016, a DFT study was reported that provided some initial mechanistic information. For instance, the hydrogen atom transfer (HAT) pathway (see Scheme 1 with L = Cys; only the first step was calculated) was found to be less favorable than the first step of ylide formation in Scheme 1 for aniline N–H insertion catalyzed by cytochrome P450 with a Cys axial ligand.41 With the Fe(II) metal bound ylide, a concerted dissociation/rearrangement was calculated to be kinetically more favorable than the sequential pathway, which is similar to Pathway-II and Pathway-III branch a up to only IntNC (a free enol dissociated from heme) in Scheme 1, respectively. A direct proton transfer Pathway-I was found to be of higher energy. In contrast, the dissociation Pathway-III branch b was adopted in more recent computational studies for Ser-ligated heme N–H insertions with a cyclic carbene,28,29 which offer interesting insights into the subsequent free ylide pathway to the final product (different from the above-mentioned enol pathway for the acyclic carbene [:CHCO2Et]).41 That work also uncovered the roles of water-assisted proton transfer. Compared with these two computational studies of heme proteins with negatively charged ligands (Cys and Ser), a recent study for the His-ligated heme (cytochrome c) with carbene [:CMeCO2Et]36 revealed a comparison of the first ylide formation step with a direct hydrogen transfer mechanism (similar to the hydride transfer pathway in Scheme 1, as found for Si–H36,43 and C–H insertions44). Clearly, each of the prior computational mechanistic works only studied a few possibilities and limited steps, which show that different mechanisms may be favored for biocatalysts with different axial ligands. So, there has been no comprehensive complete mechanism study from heme carbene to the final product formation, as shown in Scheme 1 (to be described in detail later), particularly for the neutral His ligand. Recent reports of heme protein reactions show that the axial ligand can modulate reactivity43,45–47 and sometimes change preferred pathways.28,48,49 So, it is important to study the full reaction pathways for His-ligated heme catalysts, which constitute ∼80% of the experimentally used engineered heme proteins for N–H insertions.30–40


image file: d4cy00999a-s1.tif
Scheme 1 Different N–H insertion pathways. L = axial ligand. H′: hydrogen, in grey, to be transferred from aniline to carbene carbon.

In addition, all these N–H insertion pathway DFT calculations28,29,36,41 used the open shell singlet (OSS) as the ground state for heme carbenes. However, it is known that certain DFT methods may not be accurate for some electronic structure calculations. Our recent studies show that only the close shell singlet (CSS) of Fe(II)-heme carbene systems yielded accurate predictions of experimental X-ray, Mössbauer and NMR spectroscopic properties.44,50 This CSS feature was further supported by a more recent experimental study using X-ray absorption near edge spectroscopy.51 Moreover, the CSS feature was confirmed in high-level multireference quantum chemical calculations using the complete active space self-consistent field (CASSCF) method for iron porphyrins with carbene [:CMeCO2Et] and several axial ligands N-methylimidazole, methyl thiolate, and hydroxide,52 which are analogues of His, Cys, and Ser ligands. This CSS feature also enabled the successful predictions of a wide range of experimental heme carbene reactivities of cyclopropanations and Si–H and C–H insertions.43,44,47

Considering the above mechanistic problems and the experimental prevalence (∼80%) of His-ligated heme proteins particularly Mb variants30–40 in N–H insertions, we chose to perform a comprehensive mechanistic investigation, as shown in Scheme 1 with L = His for the heme carbene N–H insertion reaction of model [Fe(Por)(5-MeIm)(CHCO2Et)] (R1; FeII(Por): non-substituted porphyrin; 5-MeIm: 5-methylimidazole for His, as used recently43,46,47,53) derived from the widely used carbene precursor ethyl diazoacetate (EDA) in this field31,32,35,38,40 with aniline (R2) substrate, using the DFT method, which enabled accurate predictions of experimental properties of heme carbenes and their reactivities, stereoselectivities, regioselectivities, and chemoselectivities;43,44,46,47,50,53–55 see Experimental details section 2.2. The electronic structure of this heme carbene R1 was first investigated using the high-level CASSCF method to compare CSS, OSS, and other spin states not reported before to validate the ground state for subsequent DFT pathway studies. As shown in Scheme 1, this is the first systematic pathway study for His-ligated heme catalyzed N-H insertion from heme carbene (R1) and substrate (R2) to the final product (PNH) and recycled catalyst (Pheme). In addition, there are several mechanistic features in Scheme 1 that have not been studied before: 1) the comparison of ylide with HAT and hydride pathways; 2) the comparison of four ylide pathway branches: a) concerted direct proton transfer from aniline nitrogen to carbene carbon with C–N bond formation to form final products (Pathway-I), b) concerted dissociation of iron porphyrin (Pheme) with the rearrangement of proton from aniline N to carbonyl O to form an enol intermediate IntNC (Pathway-II), c) dissociation to form the free ylide IntNC-III (Pathway-III), d) rearrangement to form the metal-bound enol Int1-IV, then dissociate to release free enol IntNC (Pathway-IV); 3) the study of three possible pathways from free enol to final products, including the metal-coordinated one via transition state TSO-II and metal-free pathways with and without water assistance through TSNC and TSNC-w respectively; 4) the comparison of indirect and direct pathways from the free ylide to final product for the acyclic carbene: a) rearrange to free enol viaTS2a-III first, then proceed to enol pathways, b) two direct pathways with and without water assistance through TS2b-III and TS2b-III-w, respectively. Overall, there are ten main pathways investigated in this work, with the most favorable mechanism further validated by new experimental results. It should also be noted that there is no prior experimental mechanistic study regarding the basic reaction mechanism in this field. Although the protein environment is important to determine the final biocatalytic reaction, our work, here, as the first study of the His-ligated heme protein catalyzed N–H insertion is focused on the effect of cofactor structure on reaction mechanism as done previously in this field.28,29,36,41

2. Experimental details

2.1 CASSCF calculations

For heme carbene R1, its DFT optimized geometry was used with 6-311G(d) for Fe and its coordinated atoms and 6-31G(d) for all other atoms. For the active space for the CASSCF calculations, as in a recent CASSCF study of similar heme carbenes,52 we also used the same option of 12 electrons in 12 orbitals, which are the largest level among recent CASSCF studies of heme carbene and nitrene systems56,57 as well as other metal complexes.58,59 Similarly, our studied active space also started from the examination of natural orbitals and consists of frontier five Fe 3d orbitals 3dxy, 3dyz, 3xz, 3x2y2, 3dz2, and π/π* and σ carbene orbitals involved in Fe–C bonding, see Fig. 1. Our CASSCF(12,12) calculations used the NRoot = 5 with the state average option using equal weights. More details of the configuration state functions (CSFs) with >1% weight for each studied state are discussed in ESI section S2. We used the Gaussian 16 program for these calculations.60
image file: d4cy00999a-f1.tif
Fig. 1 A) Active space molecular orbitals (MOs) ϕ1–12 of R1. Contour values = ±0.04 au. Atom colors: N – blue, O – red, C – cyan, H – grey, Fe – black. B) Energy levels of the studied singlet, triplet, and quintet spin states.

2.2 DFT calculations

All the models investigated in this work were subject to full geometry optimizations without any symmetry constraints using the PCM method61 with a dielectric constant of 4.0 to simulate the protein environment effect as done previously.28,29,36,41,43–48,53–55,62–64 Compared to this value, additional calculations using a high-end dielectric constant value of 78.3553 for the pure aqueous solvent environment (∼20-fold increase) only increased the reaction barrier by 1.22 kcal mol−1 for the heme carbene formation reaction. As a protein contains many non-polar residues and is much less polar than water, the impact on the absolute value of the reaction barrier will be much less than this value. In fact, the current use of the dielectric constant of 4.0 enabled accurate quantitative predictions of experimental heme reaction barriers (e.g., an average error of 0.11 kcal mol−1 (ref. 48)) and experimental non-heme reaction barriers (e.g., an error of 0.36 kcal mol−1 (ref. 64)) besides reproducing reactivity trends of a number of biocatalytic heme carbene and nitrene transfer reactions.43,45,47,53–55

Geometry optimization was conducted using a range-separated hybrid DFT method with dispersion corrections, ωB97XD,65 with the effective core potential (ECP) basis LanL2DZ66 for iron and the triple-zeta basis 6-311G(d) for all other elements, based on its accurate predictions of reactions involving heme carbenes.43,44,46,47,50,53–55 This basis set is larger than the ones used in recent computational studies of heme N–H insertions.36,41,67,68 The use of a much larger 6-311++G(2d,2p) basis for all non-metal atoms was found to yield similar results for heme carbene reactions44 and thus further support the efficient use of the current basis set here. The use of an ECP basis for metals is common in many reaction studies involving transition metal carbenoids, such as Ir/Ru/Th carbenes.69–71 The advantage of an ECP basis is the inclusion of a relativistic effect basically absent in an all-electron basis set. In addition, it is available for all transition metals, which may allow direct comparisons of the effects of a vast amount of metal centers. The alternative use of an all-electron basis for the metal center47 was recently found to yield qualitatively the same conclusions of geometric, electronic, and energetic features for heme carbene reactions and therefore supports the use of the LanL2DZ basis here, which may help direct comparisons with late transition metals in future studies, for which ECP basis is more readily available and commonly used. The frequency analysis was used to verify the nature of the stationary points on respective potential energy surfaces and to provide zero-point energy corrected electronic energies (EZPE's), enthalpies (H's), and Gibbs free energies (G's) at 1 atm and experimental reaction temperature, i.e., room temperature. All the DFT calculations were performed using the Gaussian 09 program72 as done in recent works on heme carbene reactions.43–47,50,53–55 The atomic charges and spin densities are from the natural population analysis (NPA) and Mulliken schemes implemented in Gaussian 09. Intrinsic reaction coordinate calculations were done to prove the connection between each TS and its corresponding stationary states. We also performed a detailed conformation and spin state study for relevant species, as given in ESI, and then selected the most favorable ones in the reaction pathway discussion here. Except for the CSS states, spin unrestricted DFT calculations were performed, including the OSS states, for which we used the broken-symmetry method.73

2.3 Reagents and analytical methods

All chemicals and reagents were purchased from commercial suppliers (Sigma Aldrich, TCI Chemicals) and used without any further purification.

2.4 Protein expression and purification

Wild-type Mb and engineered Mb variants were cloned and expressed in E. coli C41(DE3) cells as described previously.34 Briefly, cells were grown in TB medium (ampicillin, 100 mg L−1) at 37 °C (180 rpm) until OD600 reached 0.8–1.0. Cells were then induced with 0.25 mM β-D-1-thiogalactopyranoside (IPTG) and 0.3 mM δ-aminolevulinic acid (ALA). After induction, cultures were shaken at 27 °C (180 rpm), harvested after 20 h by centrifugation (4000 rpm, 20 min, 4 °C) and resuspended in an Ni-NTA lysis buffer (50 mM KPi, 250 mM NaCl, 10 mM histidine, pH 8.0). Resuspended cells were frozen and stored at −80 °C. Cell suspensions were thawed at room temperature, lysed by sonication, and clarified by centrifugation (14[thin space (1/6-em)]000 rpm, 50 min, 4 °C). The clarified lysate was transferred to a Ni-NTA column equilibrated with the Ni-NTA lysis buffer. The protein was washed with the Ni-NTA wash buffer (50 mM KPi, 250 mM NaCl, 20 mM histidine, pH 8.0). Proteins were eluted with the Ni-NTA elution buffer (50 mM KPi, 250 mM NaCl, 250 mM histidine, pH 7.0). After buffer exchange (50 mM KPi, pH 7.0), the proteins were stored at +4 °C. Myoglobin concentration was determined by UV/vis spectroscopy using an extinction coefficient of ε410 = 157 mM−1 cm−1.

2.5 N–H insertion reactions

Under standard reaction conditions, reactions were carried out at a 400 μL scale using 10 μM myoglobin, 5 mM amine, 10 mM diazo compound, with or without 50 mM DMPO and 10 mM sodium dithionite. In a typical procedure, in an anaerobic chamber, a solution containing the desired myoglobin variant was mixed with a solution of sodium dithionite in nitrogen-purged potassium phosphate buffer (50 mM, pH 8.0). Reactions were initiated by the addition of 10% DMF or DMPO (50 mM stock in DMF), amine (400 mM stock solution in EtOH), followed by the addition of the diazo compound (400 mM stock solution in EtOH), and the reaction mixtures were stirred in the chamber for 1 h at room temperature. For product analysis, an internal standard (20 μL of benzodioxole at 100 mM in ethanol) was added to the reaction mixture, followed by extraction with dichloromethane (400 μL) and analysis by GC-FID.

2.6 Synthetic procedures and product characterization

Detailed procedures and characterizations for the synthesis of N–H insertion products are provided in the ESI.

3. Results and discussion

3.1 CASSCF study of heme carbene

To obtain a rigorous electronic structure of R1, we first investigated different spin states (S = 0, 1 and 2) using CASSCF calculations. The active space orbitals in our CASSCF(12,12) (12 electrons in 12 active orbitals) calculation are shown in Fig. 1A. Because each spin state can have different contributions or resonance structures of the fragments in R1, each of the five lowest energy levels for S = 0, 1, 2 from CASSCF calculations are shown in Fig. 1B.

As shown in Fig. 1B and Tables S2–S4, the CSS state with a single predominant contribution from the CSS configuration state function that has all frontier occupied orbitals with double occupancy (i.e., paired) is the overall lowest energy spin state for this His-ligated heme with the carbene derived from the widely used ethyl diazoacetate. All other higher-energy singlets exhibit OSS features (see ESI for details). The lowest OSS state has an ΔE of 22.14 kcal mol−1, higher than the CSS ground state. This trend is the same as found in the previous CASSCF result of a similar heme carbene52 and different from previous DFT pathway studies in this field.28,29,36,41 This is not unexpected as DFT is a single determinant method; thus, it is difficult to handle electronic states with multi-reference nature. In fact, most of the calculated electronic states involve a strong mixture of MOs with different characters, see ESI section S2. According to our CASSCF calculations, the next energy level above the CSS ground state is a triplet (it is of ΔE 4.70 kcal mol−1 higher, Table S3), which is lower than the most stable OSS state by 17.44 kcal mol−1. This trend is also consistent with the recently reported CASSCF result of a similar heme carbene.52 We further examined the quintet states and found that the lowest S = 2 state is of even a higher ΔE of 12.46 kcal mol−1 than CSS, see Table S4. These results support that this heme carbene has CSS as the overall ground state as found experimentally and computationally for heme carbenes.43,44,47,50,51 So, this CSS ground state is used for R1 in subsequent pathway studies.

3.2 Ylide pathways

We first investigated ylide pathways in Scheme 1. A detailed conformation and spin state study was done for relevant species (see ESI for details) to provide the most favorable conformations and spin states in the reaction pathway discussion here. As shown in Fig. 2, it is kinetically and thermodynamically very favorable for R1 to react with aniline (R2) to form the C–N bond in the ylide intermediate (Intylide) via transition state (TSylide), a common step for all four ylide pathways.
image file: d4cy00999a-f2.tif
Fig. 2 A) Energy diagrams for the successfully located ylide pathways of heme-carbene-catalyzed N–H insertions. B) Schematic energy profile for the most favourable steps in Pathway-I, -II and -III. The energy levels are based on ΔG. Relative ΔE and ΔG energies (in kcal mol−1) are outside and inside the parenthesis, respectively. ΔEG results for TS1-III shown in italics are from potential energy scan calculations; see ESI for details.

Pathway-I is the concerted pathway with direct proton H′ transfer from N to C along with the Fe–C bond dissociation in the transition state TS1-I to form the product PNH and regenerate the biocatalyst Pheme. Its relative energy to the starting reactants is 4.12 kcal mol−1, and its relative barrier to IntylideG: 33.34 kcal mol−1) is the highest among all studied ylide pathways (to be discussed in detail later), see Fig. 2. This trend for the neutral ligand His ligated heme carbene N–H insertion is the same as found for the heme with a negatively charged Cys ligand.41 The intrinsic reaction coordinate (IRC) calculations for TS1-I (see Fig. S4) confirmed that it is connected with Intylide (not the reactant complex, i.e., the C–N bond in ylide is maintained) and the product complex of PNH and Pheme.

Pathway-II is also concerted but it involves the rearrangement of ylide to enol (aniline first transfers H′ to carbonyl oxygen) along with dissociation of the Fe–C bond. Its transition state TS1-II to form the enol intermediate (IntNC) has a relative barrier to IntylideG: 10.91 kcal mol−1) that is much lower than Pathway-I. This step is similar to that proposed for the Cys-ligated heme with the same carbene.41 For TS1-II, the Fe–C bond length is 2.658 Å, while it is 2.206 Å for TS1-I, as shown in Table S9. A longer Fe–C bond length in TS1-II in Pathway-II leads to more favorable bond dissociation compared to TS1-I in Pathway-I. Along with the Fe–C bond dissociation in both pathways, the N–H′ bond in aniline breaks, and H′ is transferred to either the carbonyl oxygen to form a new O–H′ bond in TS1-II or the carbene's carbon to generate the new C–H′ bond in TS1-I. Here, the O–H′ bond formation is more favorable as compared to the C–H′ bond formation because oxygen is more electronegative than carbon. Therefore, both the bond dissociation and new bond formation parts contribute to a lower barrier of TS1-II as compared to TS1-I.

However, the subsequent steps from IntNC to the final N–H insertion product PNH have not been reported before. For this reason, we investigated three possible subsequent pathways.

Since IntNC is released in the active site with heme nearby, we first studied if the coordination of Fe to carbonyl O could facilitate the rearrangement or transfer of H′ on O to C as in PNH, see Scheme 1. The metal–O coordinated rearrangement mechanism was previously studied in a non-heme carbene system74 but has not been explored for heme carbenes before. It starts with the coordination step to generate intermediate IntO-II and then proceed via the transition state TSO-II to form another intermediate complex, IntComplex-II, which eventually forms products. As shown in Scheme S1, from IntNC with an energy of −26.60 kcal mol−1, the coordination stabilizes IntO-II by 2.29 kcal mol−1. However, TSO-II is significantly uphill by 49.07 kcal mol−1 to be 20.18 kcal mol−1 with respect to starting reactants, although the subsequent IntComplex-II is thermodynamically very favorable. So, this metal-coordinated enol transformation has a relative ΔG of 46.78 kcal mol−1 from IntNC, which is very high.

We then studied the direct transfer mechanism for enol proton in IntNC to carbene's carbon to yield the final product. The relative barrier for this transition state TSNC is 44.26 kcal mol−1, which is still high, although better than the metal-coordinated one.

In light of the beneficial effect of water assistance in proton transfer for ylide in recent computational works of negatively charged ligand ligated heme carbene N–H insertions28 and the availability of water in such biocatalytic environments, we investigated this transition state in the cases of one water and two water molecules and compared with the above water-free result. It was found that the two-water model involving a concerted rearrangement of H′ from carbonyl oxygen to carbene carbon via the two-water cascade and then to product formation, as shown in Scheme 1, is the best; see ESI section S5 for details. Its (TSNC-W) relative barrier from IntNC is decreased to 16.99 kcal mol−1 in contrast with the water-free case of 44.26 kcal mol−1. This trend of water assistance for enol rearrangement is similar to the ylide proton transfer mechanisms calculated recently.28 But, a comparison of the water-assisted proton transfer mechanism for enol with metal-assisted and water-free pathways here is reported for the first time in this field.

As shown in Fig. 2, the most favorable TSNC-W for enol transformation to the final product has a ΔG of −9.61 kcal mol−1 with respect to the starting reactants. Compared to the first transition state in Pathway-II, TS1-II, this step is of higher energy and thus is the rate-determining step (RDS). Therefore, Pathway-II is more favorable than Pathway-I because TSNC-W has a lower ΔG than TS1-I by 13.73 kcal mol−1 (Fig. 2).

As seen from Scheme 1, Pathway-III is stepwise, with the ylide dissociation from heme occurring first viaTS1-III to form the free ylide IntNC-III, followed by two branches to the final product formation. Numerous trials to locate TS1-III were unsuccessful, which is similar to a recent computational work of the dissociation pathway for the Ser-ligated heme carbene N–H insertion.28 We then conducted potential energy scan (PES) calculations along the Fe–C coordinate for this dissociation process (see ESI section S6 for details), which were also used in recent heme carbene N–H insertion calculations of similar systems.41,67,68 Results show that there is an energy cost from both cleaving the Fe–C bond and breaking the hydrogen bond between amine's proton and porphyrin nitrogen, which bears a maximum ΔE cost of 21.74 kcal mol−1 from Intylide, which was added to all relative energetic terms of Intylide to obtain the estimated relative energies for TS1-III. This leads to the relative ΔE and ΔG of −26.70 and −7.48 kcal mol−1 with respect to starting reactants, as shown in Fig. 2.

After the free ylide IntNC-III is formed, we compared the indirect (branch a) and direct (branch b) pathways to the final product PNH, as shown in Scheme 1.

The rearrangement transition state TS2a-III of H′ from N to the carbonyl O to form the free enol IntNC along branch a has a very low relative barrier compared to IntNC-IIIE: 3.02 kcal mol−1; ΔG: −0.35 kcal mol−1, calculated from data in Table S8), which is much lower compared to the cyclic carbene system (ΔE: 9.4 kcal mol−1; ΔG: 6.4 kcal mol−1).28,29 Then, starting from the enol IntNC, subsequent different enol transformation mechanisms have been studied and discussed above for Pathway-II, in which the water-assisted enol rearrangement viaTSNC-w is the most favorable one. As TSNC-w is higher than TS2a-III to form enol, it is the RDS along this indirect (branch a) pathway from free ylide, i.e., branch a has the highest ΔG of −9.61 kcal mol−1.

For the direct pathways in branch b from free ylide to final product, we also studied them with and without water assistance (see Scheme 1) viaTS2b-III and TS2b-III-w, respectively. As found for the cyclic carbene system reported recently,28 the two-water assisted model with first protonation at the carbene carbon via transition state TS2b-III-w and subsequent spontaneous deprotonation at the N position to yield product PNH, is of significantly lower energy than the water-free TS2b-III by ΔG of 33.07 kcal mol−1 (see Scheme S1), and thus is the preferred pathway for the direct transformation of free ylide to product along branch b.

Since the highest energy species along the favorable direct branch b from the free ylide IntNC-III is itself (see Fig. 2), it has a relative Gibbs free energy of −13.62 kcal mol−1 with respect to reactants and thus lower than the relative Gibbs free energy of −9.61 kcal mol−1 for TSNC-w (the highest energy species along the indirect branch a). As a result, the direct branch b is the more favorable one and thus the best Pathway-III mechanism. As shown in Fig. 2, this pathway's RDS step is dissociation with a ΔG of −7.48 kcal mol−1, which is of higher energy than the RDS in Pathway-II (−9.61 kcal mol−1). So, among the first three ylide pathways, Pathway-II with RDS of TS1-II is the most preferred one.

In contrast with Pathway-III, which features a stepwise mechanism with the initial dissociation of free ylide and then rearrangement to enol as studied for the Cys-ligated heme carbene N–H insertion41 or direct formation of the final product for the Ser-ligated case,28Pathway-IV is also stepwise but involves first the rearrangement viaTS1-IV then dissociation TS2-IV for comparison, see Scheme 1. However, all efforts to locate a pure metal-bound rearrangement TS1-IV of H′ from N to the carbonyl O were optimized to the concerted rearrangement and dissociation transition state TS1-II in the most favorable ylide mechanism Pathway-II, as discussed above. This Pathway-IV transition state was also absent for the Cys-ligated heme carbene N–H insertion with the same carbene as studied here.41 In addition, our results show that the lowest energy intermediate Int1-IV after TS1-IV has the Fe–C bond already broken with an RFeC of 3.102 Å, see Table S5. This further supports that the rearrangement causes simultaneous dissociation and thus this stepwise Pathway-IV may not exist.

Overall, these results from heme carbene and substrate to the final products show for the first time that only three out of the four ylide pathways could be successfully located for the His-ligated heme carbene N–H insertion. Fig. 2 collects the branches of each located ylide pathway, which indicates that Pathway-II from the metal-bound ylide to enol IntNC and then via water-assisted rearrangement TSNC-w to generate the final N–H insertion product is the kinetically most favorable ylide mechanism for the studied acyclic carbene, in contrast with the preferred Pathway-III reported for cyclic carbene with Ser-ligated heme.28

As seen from Fig. 2, the overall RDS along the favorable ylide pathway here is ylide formation compared to subsequent steps, as found for both Cys- and Ser-ligated heme biocatalytic N–H insertions.28,41 It has a small barrier and this reaction to the final product formation is also thermodynamically favorable with a net reaction energy of ΔG° −55.04 kcal mol−1, see Fig. 2.

3.3 Alternative pathways

Besides the ylide pathways, we also investigated a stepwise HAT mechanism and concerted hydride transfer mechanism for comparison; see Scheme 1.

As the hydrogen atom bears radical features, the first transition state TS1HAT to abstract a hydrogen atom from aniline to carbene carbon displays an OSS feature with FeIII (S = −1/2, spin density of −1.033 e) and shared spin (S = 1/2) between N and C (total spin density is 0.950 e), see Table S11. This feature can be seen in Fig. 3 for its spin density diagram, in which the transferred H′ atom shows an opposite spin with the remaining aniline and carbene. In the formed intermediate IntHAT, the remaining aniline dissociates with an RCN of 3.963 Å (Table S13). A subsequent radical rebound leads to the formation of the C–N bond in TS2HAT, which has relatively lower energy compared to TS1HAT (ΔΔG: −9.79 kcal mol−1, see Fig. 2). In this step, both C and N again showed radical features, but with opposite spin directions (−0.792 e and 0.648 e, respectively, see Table S11 and Fig. 2), ready for a radical coupling to facilitate the formation of the final product PNH. While the partial C–N bond formation can be seen by its significant distance shortening from 3.963 Å in IntHAT to 2.581 Å in TS2HAT, a concomitant Fe–C bond elongation of ∼0.6 Å in this step also indicates its partial cleavage, which proceeds to the final release of the product.


image file: d4cy00999a-f3.tif
Fig. 3 Energy diagrams for the most favorable steps in the ylide and HAT pathways of heme-carbene catalyzed N–H insertions, where the energy levels are based on ΔG. Relative ΔE and ΔG energies (in kcal mol−1) are outside and inside the parenthesis, respectively. Transition state structures are shown. The spin densities are shown for TS1HAT and TS2HAT. Contour values = ±0.01 au.

As shown in Fig. 3, the RDS barrier (ΔG) for the HAT pathway is 13.64 kcal mol−1, which is significantly higher than the RDS of the ylide pathway by 13.46 kcal mol−1. This shows that the ylide mechanism is more favorable for the His-ligated heme carbene N–H insertion based on a complete reaction pathway to the final product formation, which is the same as found previously for the Cys-ligated system41 based on the first-step comparisons.

We also attempted the concerted nonsynchronous hydride transfer with C–N bond formation and Fe–C dissociation mechanism as found in prior studies for the heme carbene catalyzed C–H/Si–H insertion reactions.36,43,44 However, all attempts to get this transition state TSHydride were optimized as either the concerted proton H′ transfer from N to C with the Fe–C bond dissociation TS1-I in the ylide pathway or TS1HAT in the HAT pathway, suggesting that the hydride pathway is not favorable here. This may be a result of the much higher electronegativity of N vs. C/Si, which makes it difficult to form hydride from an N–H bond.

3.4 Mechanistic features of the overall RDS

Based on the above systematic mechanism study, the ylide Pathway-II is most favorable for His-ligated heme carbene N–H insertion and the overall RDS is ylide formation. The key geometry and charge changes with respect to reactants were analyzed, which were found to be useful to understand the mechanistic features of this RDS step as reported recently for the heme carbene cyclopropanation and C–H/Si–H insertions.43,44,47 As shown in Fig. 4A, the largest geometric change is the partial formation of a C–N bond in TSylide, with visible shortening of the Fe–His bond (∼−0.03 Å) to push the carbene away (elongation of ∼0.03 Å in Fe–C bond length) via the axial ligand's trans effect to facilitate its attack of the substrate. In this transition state, the N–H bond length change (≤0.003 Å) from substrate R2 is negligible compared to its difference (≥0.01 Å) with respect to the formed Intylide; see data in Table S9. This suggests an early transition state feature for this RDS. As demonstrated in Fig. 4B, the largest charge transfer is from the aniline substrate to heme carbene (0.061 e), exhibiting the electrophilicity of carbene, which leads to a negatively charged carbene carbon. However, this charge transfer is significantly smaller than that for many heme carbene cyclopropanation and C–H/Si–H insertions,43,44,47 which are mostly around 0.3–0.4 e for concerted nonsynchronous reaction mechanisms. These results suggest that heme carbene N–H insertions have a distinctive ylide mechanism, and the effects of electronic features of substrates may be less significant than the electrophilic heme carbene cyclopropanation and C–H/Si–H insertions,43,44,47 while the reactant steric effect may affect the reactivity via its influence to the metal-bound ylide formation in RDS. These features are consistent with experimental data that both electron-donating and electron-withdrawing substrates are tolerated, and bulky reactants have relatively lower yields in heme-catalyzed N–H insertions.27,28,32–34,37,38,40
image file: d4cy00999a-f4.tif
Fig. 4 A) Key geometric parameters of TSylide (in black) and geometry changes (in blue) from reactants (R1 and R2) to TSylide. C′ is carbonyl carbon. B) Atomic charge changes (unit: e) from reactants to TSylide in black. Charge transfer is indicated by arrows and numbers in parentheses. Substrate aniline is shown in the box. Oval represents porphyrin.

4. Experimental mechanistic investigation

In a previous study, an engineered active site variant of sperm whale myoglobin, Mb(H64V,V68A), was determined to catalyze an N–H carbene insertion reaction between aniline and ethyl α-diazoacetate (EDA) and ethyl 2-diazopropanoate (EDP) with up to 6150 TON and 130 TON, respectively.34 To investigate their radical vs. non-radical nature, these N–H insertion reactions were carried out in the presence and in the absence of the radical spin trapping agent 5,5-dimethyl-1-pyrroline N-oxide (DMPO). In the previously published work, we successfully applied DMPO to probe the radical mechanism of an engineered myoglobin biocatalyst for another carbene transfer reaction (cyclopropanation): the DMPO-dependent inhibition was observed, and the radical intermediate is bound to the enzyme. The success of those experiments gives us confidence that the DMPO is a suitable probe for trapping the long-lived intermediate in this system if formed.75 As shown in Table 1, no effect in product yield was observed for the reactions with EDA catalyzed by wild-type myoglobin, Mb(H64V,V68A), or hemin in the presence of DMPO when compared to parallel reactions performed in the absence of the spin trapping reagent (entries 1, 3, and 5 vs. 2, 4, and 6, respectively). In contrast, a dramatic reduction in product yield (∼90% reduction) was observed for the Co(TPP)-catalyzed reaction in the presence of the radical trapping reagent (entry 7 vs. 8), which is consistent with the radical reactivity of this catalyst.76 Similar results are observed in the N–H insertion reaction of aniline and EDP (entries 9, 11, and 13 vs. 10, 12, and 15, respectively). These results thus agree with those obtained from our computational analyses in supporting a non-radical mechanism for hemoprotein-catalyzed N–H insertion.
Table 1 Catalytic activity of myoglobin variants and transition metal catalysts in the N–H insertion reaction with diazo compounds in the presence and absence of a radical trap reagenta

image file: d4cy00999a-u1.tif

Entry Catalyst (R) DMPOb Yieldc,d (%)
a Reaction conditions: 5 mM aniline, 10 mM diazo compound, 10 μM Mb variant (or hemin), 10 mM sodium dithionite in 50 mM potassium phosphate buffer (pH 7) containing 10% DMF or 50 mM DMPO, room temp., 1 hour, anaerobic conditions. Reaction conditions B: 0.24 mmol aniline, 1.5 equiv. diazo compound, 5 mol% Co(TPP) in toluene, 60 °C, 16 hours, anaerobic conditions. b With or without 10 equiv. DMPO relative to styrene. c GC yields calculated using calibration curves with authentic standards. d Standard deviations calculated from triplicate experiments.
1 Mb H No 66 (±3)
2 Mb H Yes 60 (±3)
3 Mb(H64V, V68A) H No 79 (±4)
4 Mb(H64V, V68A) H Yes 81 (±6)
5 Hemin H No 47 (±2)
6 Hemin H Yes 49 (±4)
7 Co(TPP)B H No 60 (±3)
8 Co(TPP)B H Yes 2 (±1)
9 Mb Me No 65 (±4)
10 Mb Me Yes 63 (±2)
11 Mb(H64V, V68A) Me No 51 (±3)
12 Mb(H64V, V68A) Me Yes 42 (±4)
13 Hemin Me No 32 (±1)
14 Hemin Me Yes 33 (±5)
15 Co(TPP)B Me No 54 (±1)
16 Co(TPP)B Me Yes 3 (±1)


5. Conclusions

This work provides a most comprehensive mechanistic investigation of heme carbene N–H insertion with many pathway comparisons revealed for the first time, as briefly summarized at the end of the Introduction. Regarding the electronic structure of the involved heme carbene from the widely used precursor EDA, it was found to have a CSS ground state based on high-level CASSCF calculations, consistent with prior CASSCF works on similar heme carbenes52 and computational predictions of their experimental spectroscopic properties44,50 as well as related experimental works.51 Results show that the HAT and hydride transfer mechanisms are less favorable than the ylide pathway, which has four main sub-pathways with a number of subsequent branches. Overall, the most favorable pathway for the His-ligated heme catalyzed N–H insertion with the acyclic carbene studied here starts with the formation of a metal-bound ylide, then undergoes a concerted rearrangement of ylide to enol and dissociation of the Fe–C bond, and finally generate the N–H insertion product via the water-assisted rearrangement of enol. The preference for the ylide pathway over the HAT mechanism found in the above systematic computational study was further supported by the additional experimental work. A detailed analysis of the geometric and electronic features of the overall RDS (ylide formation) shows that this reaction is of mild electrophilicity, and the steric effect may also influence reactivity. Future studies to include the whole protein will be performed to offer a more detailed understanding of the effect of the protein environment on such biocatalytic reactions. Given that 79% of the experimental biocatalytic heme N–H insertions employ the cofactor with a His axial ligand and 86% of the experimental work in this area utilized acyclic carbenes,27–40 the novel and significant mechanistic results here will facilitate future heme-based biocatalytic N–H insertion research for sustainable chemical synthesis.

Data availability

The authors declare that the data supporting the findings of this study are available within the paper and its ESI files.

Author contributions

Y. Z. conceived the idea and designed the research. R. L. K., X. Z., and C. W. conducted the computational studies. R. D. A. performed the experimental work under the guidance and supervision of R. F. All authors participated in the data analyses and preparation of data tables and figures. R. L. K. and Y. Z. wrote the manuscript with input and additions from all other co-authors.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by a U.S. National Science Foundation grant CHE-2054897 to Y. Z and a U. S. National Institute of Health GM098628 to R. F. R. F. acknowledges chair endowment support from the Robert A. Welch Foundation (Chair, AT-0051).

References

  1. C. Bolm, A. Kasyan, K. Drauz, K. Günther and G. Raabe, Angew. Chem., Int. Ed., 2000, 39, 2288–2290 CrossRef CAS PubMed .
  2. F. A. Davis, T. Fang and R. Goswami, Org. Lett., 2002, 4, 1599–1602 CrossRef CAS PubMed .
  3. K. Yamazaki and Y. Kondo, Chem. Commun., 2002, 210–211 RSC .
  4. K. E. Bashford, A. L. Cooper, P. D. Kane, C. J. Moody, S. Muthusamy and E. Swann, J. Chem. Soc., Perkin Trans. 1, 2002, 1672–1687 RSC .
  5. S.-H. Lee, B. Clapham, G. Koch, J. Zimmermann and K. D. Janda, J. Comb. Chem., 2003, 5, 188–196 CrossRef CAS PubMed .
  6. A. C. B. Burtoloso and C. R. D. Correia, Tetrahedron Lett., 2004, 45, 3355–3358 CrossRef CAS .
  7. J. R. Davies, P. D. Kane and C. J. Moody, Tetrahedron, 2004, 60, 3967–3977 CrossRef CAS .
  8. H. Matsushita, S.-H. Lee, K. Yoshida, B. Clapham, G. Koch, J. Zimmermann and K. D. Janda, Org. Lett., 2004, 6, 4627–4629 CrossRef CAS PubMed .
  9. I. Aviv and Z. Gross, Chem. Commun., 2006, 4477–4479 RSC .
  10. B. Liu, S.-F. Zhu, W. Zhang, C. Chen and Q.-L. Zhou, J. Am. Chem. Soc., 2007, 129, 5834–5835 CrossRef CAS PubMed .
  11. M. E. Morilla, M. M. Díaz-Requejo, T. R. Belderrain, M. C. Nicasio, S. Trofimenko and P. J. Pérez, Chem. Commun., 2002, 2998–2999 RSC .
  12. J. R. Davies, P. D. Kane and C. J. Moody, J. Org. Chem., 2005, 70, 7305–7316 CrossRef CAS PubMed .
  13. J. Bariwal and E. Van der Eycken, Chem. Soc. Rev., 2013, 42, 9283–9303 RSC .
  14. H. Kohls, F. Steffen-Munsberg and M. Höhne, Curr. Opin. Chem. Biol., 2014, 19, 180–192 CrossRef CAS PubMed .
  15. J. E. Kim, S. Choi, M. Balamurugan, J. H. Jang and K. T. Nam, Trends Chem., 2020, 2, 1004–1019 CrossRef CAS .
  16. Y. Chen, R. Zhang, Z. Chen, J. Liao, X. Song, X. Liang, Y. Wang, J. Dong, C. V. Singh, D. Wang, Y. Li, F. D. Toste and J. Zhao, J. Am. Chem. Soc., 2024, 146, 10847–10856 CrossRef CAS PubMed .
  17. S. Harada, S. Hirose, M. Takamura, M. Furutani, Y. Hayashi and T. Nemoto, J. Am. Chem. Soc., 2024, 146, 733–741 CrossRef CAS PubMed .
  18. D. Gillingham and N. Fei, Chem. Soc. Rev., 2013, 42, 4918–4931 RSC .
  19. J. M. Antos and M. B. Francis, J. Am. Chem. Soc., 2004, 126, 10256–10257 CrossRef CAS PubMed .
  20. C.-M. Ho, J.-L. Zhang, C.-Y. Zhou, O.-Y. Chan, J. J. Yan, F.-Y. Zhang, J.-S. Huang and C.-M. Che, J. Am. Chem. Soc., 2010, 132, 1886–1894 CrossRef CAS PubMed .
  21. K. Tishinov, K. Schmidt, D. Häussinger and D. G. Gillingham, Angew. Chem., Int. Ed., 2012, 51, 12000–12004 CrossRef CAS PubMed .
  22. L. K. Baumann, H. M. Mbuvi, G. Du and L. K. Woo, Organometallics, 2007, 26, 3995–4002 CrossRef CAS .
  23. I. Aviv and Z. Gross, Chem. – Eur. J., 2008, 14, 3995–4005 CrossRef CAS PubMed .
  24. P. Le Maux, I. Nicolas, S. Chevance and G. Simonneaux, Tetrahedron, 2010, 66, 4462–4468 CrossRef CAS .
  25. H. X. Wang, Q. Wan, K. H. Low, C. Y. Zhou, J. S. Huang, J. L. Zhang and C. M. Che, Chem. Sci., 2020, 11, 2243–2259 RSC .
  26. C. Ma, S. Wang, Y. Sheng, X. L. Zhao, D. Xing and W. Hu, J. Am. Chem. Soc., 2023, 145, 4934–4939 CrossRef CAS PubMed .
  27. Z. J. Wang, N. E. Peck, H. Renata and F. H. Arnold, Chem. Sci., 2014, 5, 598–601 RSC .
  28. Z. Liu, C. Calvo-Tusell, A. Z. Zhou, K. Chen, M. Garcia-Borras and F. H. Arnold, Nat. Chem., 2021, 13, 1166–1172 CrossRef CAS PubMed .
  29. C. Calvo-Tusell, Z. Liu, K. Chen, F. H. Arnold and M. Garcia-Borras, Angew. Chem., Int. Ed., 2023, 62, e202303879 CrossRef CAS PubMed .
  30. E. J. Moore, V. Steck, P. Bajaj and R. Fasan, J. Org. Chem., 2018, 83, 7480–7490 CrossRef CAS PubMed .
  31. G. Sreenilayam, E. J. Moore, V. Steck and R. Fasan, Adv. Synth. Catal., 2017, 359, 2076–2089 CrossRef CAS PubMed .
  32. G. Sreenilayam and R. Fasan, Chem. Commun., 2015, 51, 1532–1534 RSC .
  33. V. Steck, G. Sreenenilayam and R. Fasan, Synlett, 2020, 31, 224–229 CrossRef CAS PubMed .
  34. V. Steck, D. M. Carminati, N. R. Johnson and R. Fasan, ACS Catal., 2020, 10, 10967–10977 CrossRef CAS PubMed .
  35. M. Pott, M. Tinzl, T. Hayashi, Y. Ota, D. Dunkelmann, P. R. E. Mittl and D. Hilvert, Angew. Chem., Int. Ed., 2021, 60, 15063–15068 CrossRef CAS PubMed .
  36. M. Garcia-Borras, S. B. J. Kan, R. D. Lewis, A. Tang, G. Jimenez-Oses, F. H. Arnold and K. N. Houk, J. Am. Chem. Soc., 2021, 143, 7114–7123 CrossRef CAS PubMed .
  37. D. Nam, A. Tinoco, Z. Shen, R. D. Adukure, G. Sreenilayam, S. D. Khare and R. Fasan, J. Am. Chem. Soc., 2022, 144, 2590–2602 CrossRef CAS PubMed .
  38. V. S. Alfaro, S. O. Waheed, H. Palomino, A. Knorrscheidt, M. Weissenborn, C. Z. Christov and N. Lehnert, Chem. – Eur. J., 2022, 28, e202201474 CrossRef PubMed .
  39. N. J. Porter, E. Danelius, T. Gonen and F. H. Arnold, J. Am. Chem. Soc., 2022, 144, 8892–8896 CrossRef CAS PubMed .
  40. L. J. Sun, H. Wang, J. K. Xu, S. Q. Gao, G. B. Wen and Y. W. Lin, Inorg. Chem., 2023, 62, 16294–16298 CrossRef CAS PubMed .
  41. D. A. Sharon, D. Mallick, B. Wang and S. Shaik, J. Am. Chem. Soc., 2016, 138, 9597–9610 CrossRef CAS PubMed .
  42. Y. Zhang, Chem. – Eur. J., 2019, 25, 13231–13247 CrossRef CAS PubMed .
  43. R. L. Khade, A. L. Chandgude, R. Fasan and Y. Zhang, ChemCatChem, 2019, 11, 3101–3108 CrossRef CAS PubMed .
  44. R. L. Khade and Y. Zhang, Chem. – Eur. J., 2017, 23, 17654–17658 CrossRef CAS PubMed .
  45. Y. Wei, M. Conklin and Y. Zhang, Chem. – Eur. J., 2022, 28, e202202006 CrossRef CAS PubMed .
  46. R. L. Khade and Y. Zhang, J. Am. Chem. Soc., 2015, 137, 7560–7563 CrossRef CAS PubMed .
  47. Y. Wei, A. Tinoco, V. Steck, R. Fasan and Y. Zhang, J. Am. Chem. Soc., 2018, 140, 1649–1662 CrossRef CAS PubMed .
  48. Y. Shi and Y. Zhang, Angew. Chem., Int. Ed., 2018, 57, 16654–16658 CrossRef CAS PubMed .
  49. K. Chen, S. Q. Zhang, O. F. Brandenberg, X. Hong and F. H. Arnold, J. Am. Chem. Soc., 2018, 140, 16402–16407 CrossRef CAS PubMed .
  50. R. L. Khade, W. Fan, Y. Ling, L. Yang, E. Oldfield and Y. Zhang, Angew. Chem., Int. Ed., 2014, 53, 7574–7578 CrossRef CAS PubMed .
  51. Y. Liu, W. Xu, J. Zhang, W. Fuller, C. E. Schultz and J. Li, J. Am. Chem. Soc., 2017, 139, 5023–5026 CrossRef CAS PubMed .
  52. G. D. Stroscio, M. Srnec and R. G. Hadt, Inorg. Chem., 2020, 59, 8707–8715 CrossRef CAS PubMed .
  53. D. A. Vargas, R. L. Khade, Y. Zhang and R. Fasan, Angew. Chem., Int. Ed., 2019, 58, 10148–10152 CrossRef CAS PubMed .
  54. D. Nam, J. P. Bacik, R. L. Khade, M. C. Aguilera, Y. Wei, J. D. Villada, M. L. Neidig, Y. Zhang, N. Ando and R. Fasan, Nat. Commun., 2023, 14, 7985 CrossRef CAS PubMed .
  55. A. Tinoco, Y. Wei, J. P. Bacik, E. J. Moore, N. Ando, Y. Zhang and R. Fasan, ACS Catal., 2019, 9, 1514–1524 CrossRef CAS PubMed .
  56. G. D. Stroscio, M. Srnec and R. G. Hadt, Inorg. Chem., 2020, 59, 8707–8715 CrossRef CAS PubMed .
  57. X. Li, L. Dong and Y. Liu, Inorg. Chem., 2020, 59, 1622–1632 CrossRef CAS PubMed .
  58. S. E. Stavretis, M. Atanasov, A. A. Podlesnyak, S. C. Hunter, F. Neese and Z.-L. Xue, Inorg. Chem., 2015, 54, 9790–9801 CrossRef CAS PubMed .
  59. W. J. Yang, X. B. Chen, H. Z. Su, W. H. Fang and Y. Zhang, Chem. Commun., 2015, 51, 9616–9619 RSC .
  60. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr , J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford CT, 2019 Search PubMed .
  61. B. Mennucci and J. Tomasi, J. Chem. Phys., 1997, 106, 5151–5158 CrossRef CAS .
  62. R. A. Torres, T. Lovell, L. Noodleman and D. A. Case, J. Am. Chem. Soc., 2003, 125, 1923–1936 CrossRef CAS PubMed .
  63. M. R. A. Blomberg, T. Borowski, F. Himo, R.-Z. Liao and P. E. M. Siegbahn, Chem. Rev., 2014, 114, 3601–3658 CrossRef CAS PubMed .
  64. Y. Shi, M. A. Michael and Y. Zhang, Chem. – Eur. J., 2021, 27, 5019–5027 CrossRef CAS PubMed .
  65. J.-D. Chai and M. Head-Gordon, Phys. Chem. Chem. Phys., 2008, 10, 6615–6620 RSC .
  66. P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270–283 CrossRef CAS .
  67. Z. Liu, C. Calvó-Tusell, A. Z. Zhou, K. Chen, M. Garcia-Borràs and F. H. Arnold, Nat. Chem., 2021, 13, 1166–1172 CrossRef CAS PubMed .
  68. C. Calvó-Tusell, Z. Liu, K. Chen, F. H. Arnold and M. Garcia-Borràs, Angew. Chem., Int. Ed., 2023, 62, e202303879 CrossRef PubMed .
  69. J.-C. Wang, Z.-J. Xu, Z. Guo, Q.-H. Deng, C.-Y. Zhou, X.-L. Wan and C.-M. Che, Chem. Commun., 2012, 48, 4299–4301 RSC .
  70. K.-H. Chan, X. Guan, V. K.-Y. Lo and C.-M. Che, Angew. Chem., Int. Ed., 2014, 53, 2982–2987 CrossRef CAS PubMed .
  71. E. Nakamura, N. Yoshikai and M. Yamanaka, J. Am. Chem. Soc., 2002, 124, 7181–7192 CrossRef CAS PubMed .
  72. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr , E. J. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision B.01, Gaussian, Inc., Wallingford CT, 2010 Search PubMed .
  73. P. G. Blachly, G. M. Sandala, D. A. Giammona, D. Bashford, J. A. McCammon and L. Noodleman, Inorg. Chem., 2015, 54, 6439–6461 CrossRef CAS PubMed .
  74. X.-C. Wang, X.-S. Song, L.-P. Guo, D. Qu, Z.-Z. Xie, F. Verpoort and J. Cao, Organometallics, 2014, 33, 4042–4050 CrossRef CAS .
  75. D. M. Carminati and R. Fasan, ACS Catal., 2019, 9, 9683–9697 CrossRef CAS PubMed .
  76. H. Lu, W. I. Dzik, X. Xu, L. Wojtas, B. de Bruin and X. P. Zhang, J. Am. Chem. Soc., 2011, 133, 8518–8521 CrossRef CAS PubMed .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cy00999a

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.