Skip to main content
JACS Au logoLink to JACS Au
. 2022 Feb 25;2(3):745–761. doi: 10.1021/jacsau.2c00014

DFT Mechanistic Insights into Aldehyde Deformylations with Biomimetic Metal–Dioxygen Complexes: Distinct Mechanisms and Reaction Rules

Ruihua Zhao †,, Bei-Bei Zhang , Zheyuan Liu §, Gui-Juan Cheng ‡,*, Zhi-Xiang Wang †,*
PMCID: PMC8970012  PMID: 35373207

Abstract

graphic file with name au2c00014_0014.jpg

Aldehyde deformylations occurring in organisms are catalyzed by metalloenzymes through metal–dioxygen active cores, attracting great interest to study small-molecule metal–dioxygen complexes for understanding relevant biological processes and developing biomimetic catalysts for aerobic transformations. As the known deformylation mechanisms, including nucleophilic attack, aldehyde α-H-atom abstraction, and aldehyde hydrogen atom abstraction, undergo outer-sphere pathways, we herein report a distinct inner-sphere mechanism based on density functional theory (DFT) mechanistic studies of aldehyde deformylations with a copper (II)–superoxo complex. The inner-sphere mechanism proceeds via a sequence mainly including aldehyde end-on coordination, homolytic aldehyde C–C bond cleavage, and dioxygen O–O bond cleavage, among which the C–C bond cleavage is the rate-determining step with a barrier substantially lower than those of outer-sphere pathways. The aldehyde C–C bond cleavage, enabled through the activation of the dioxygen ligand radical in a second-order nucleophilic substitution (SN2)-like fashion, leads to an alkyl radical and facilitates the subsequent dioxygen O–O bond cleavage. Furthermore, we deduced the rules for the reactions of metal–dioxygen complexes with aldehydes and nitriles via the inner-sphere mechanism. Expectedly, our proposed inner-sphere mechanisms and the reaction rules offer another perspective to understand relevant biological processes involving metal–dioxygen cores and to discover metal–dioxygen catalysts for aerobic transformations.

Keywords: aldehyde deformylations, metal−dioxygen complexes, aldehyde deformylation mechanism, inner-sphere mechanism, homolytic aldehyde C−C bond cleavage, dioxygen O−O bond cleavage, DFT mechanistic study

Introduction

Aldehyde deformylations are chemical reactions taking place in organisms during metabolism and biosynthesis.1,2 These biochemical reactions are catalyzed by metalloenzymes such as cytochrome P450 and aldehyde decarbonylase (ADO).36 By binding to molecular oxygen, the metal cofactors of the metalloenzymes form metal–dioxygen active cores that perform these reactions.710 Inspired by these biological processes, much effort has been devoted to the study of small-molecule transition-metal–dioxygen complexes (denoted TM–O2 complexes hereafter).1116 On the one hand, TM–O2 complexes serve as biomimetic models to deeply understand the relevant biological processes like deformylations; on the other hand, studying the chemistry of TM–O2 complexes could lead to the discovery of aerobic catalytic transformations.1740 The use of abundant molecular O2 as an oxidant offers a green and sustainable approach to produce valuable chemical products.

TM–O2 complexes often feature amphoteric reactivity, including the electrophilic reactivity as exhibited in the hydrogen atom abstraction reactions4143 and the nucleophilic reactivity as found in the aldehyde deformylations.18,31,4448 In the past two decades, using aldehydes such as 2-phenylpropionaldehyde (2-PPA), trimethylacetaldehyde (TMA), cyclohexanecarboxaldehyde (CCA), and other electrophiles as probes, the nucleophilic reactivity of a wide range of TM–O2 complexes has been investigated, where TM = V,49,50 Mn,26,31,40,44,45 Fe,23,47 Co,46,48 Ni,28,32 and Cu.18 Notably, using O2 as an oxidant, catalytic deformylation of 2-PPA has also been reported.19,51,52 As de Visser and co-workers in 2021 have presented an excellent and comprehensive review on this topic,14 we herein highlight the works relevant to the present study.

The copper–dioxygen core present in oxygenase (e.g., lytic polysaccharide monooxygenase53) and oxidase (e.g., heme-copper oxidases54) enzymes55,56 encouraged the synthesis and the reactivity study of various biomimetic copper–dioxygen model complexes.12,15 In this context, Tolman and co-workers in 2010 synthesized a copper (II)–superoxo complex [CuO2]K(18-crown-6) supported by a dianionic pyridine-2,6-dicarboxamide ligand.20 In 2014, McDonald and co-workers studied the nucleophilic reactivity of the complex and found that the complex reacted with acyl chlorides and a range of aldehydes (eqs 1–4, Scheme 1).18 To improve the stability and solubility of [CuO2]K(18-crown-6), Tolman and co-workers in 2019 used phase-transfer agent Krpytofix222 (Krypt) to replace 18-crown-6, furnishing [CuO2][K(krypt)].22 It was reported that the new complex ([CuO2][K(krypt)]) reacted with 2-PPA to give copper(II) enolate complex [LCu(OC=C(Me)Ph)] in the presence of water in a different solvent (19:1 tetrahydrofuran (THF)/CH3CN). Because the two experimental studies were conducted with [CuO2] complexes having different countercations under different experimental conditions, the question of whether the copper–dioxygen core possesses nucleophilic reactivity remains unclear. It is highly desirable to have an “ideal” approach that could avoid the effects of these undefinable experimental factors. While it is not easy for experimental studies, theoretical calculations offer a convenient approach to reach the goal. Theoretical calculations produce detailed geometric, energetic, and electronic information, which is important for deeply understanding reaction mechanisms and reactivity.

Scheme 1. Deformylation Reactions Studied in This Work.

Scheme 1

Mechanistically, a widely accepted mechanism to account for aldehyde deformylations with TM–O2 complexes is the outer-sphere nucleophilic attack mechanism,18,31,4448 termed as ONA mechanism hereafter. Under the mechanism, the O2 ligand undergoes nucleophilic attack at the aldehyde carbonyl carbon to start a deformylation reaction.18,31,4448 Using a quantum mechanics/molecular mechanics (QM/MM) modeling method, Hackett and co-workers demonstrated the ONA mechanism to be feasible for the deformylation catalyzed by sterol-14α-demethylase (CYP51).57 In the experimental studies, the ONA mechanism was often postulated.18,31,4448 As exemplified in Scheme 2A,18 under the ONA mechanism, the reaction begins with nucleophilic attack to form an intermediate, followed by Baeyer–Villiger oxidation or hydrogen atom abstraction pathways, leading to the deformylation products. Because the mechanism involves abstraction of α-hydrogen of aldehydes (2-PPA and CCA) and TMA does not have an α-hydrogen, the deformylation of TMA (eq 2) terminates after the decomposition of the nucleophilic attack intermediate. As the mechanism could account for reactions in eqs 1–3 reasonably, it is puzzling that TMA could undergo deformylation (eq 2), but the more electrophilic benzaldehyde could not (eq 4), although it is possible that the nucleophilic attack is not the rate-determining step.

Scheme 2. Previously Proposed Outer-Sphere Deformylation Mechanisms (A–C) and Our Proposed Inner-Sphere Mechanisms (D).

Scheme 2

For clarity, the −1 charge states of Cu-containing structures and +1 charge state of Mn-containing structures were omitted.

As the ONA mechanism is commonly accepted, Sastri and de Visser et al.33,41 in 2016 and 2017 first discovered that the ONA mechanism could not be applied for their 2-PPA deformylation with the cationic manganese (III)–peroxo complex supported by dimethyl-2,4-di(2-pyridyl)-3-benzyl-7-(pyridin-2-ylmethyl)-3,7-diazabicyclo[3.3.1]nonan-9-one-1,5-dicarboxylate ligand, denoted 5[MnO2]+ (quintet) hereafter. They proposed an electrophilic deformylation mechanism, where the peroxo ligand of 5[MnO2]+ first abstracts the α-H atom of 2-PPA, denoted α-HAA mechanism (Scheme 2B).41 Their density functional theory (DFT) calculations indicated that the nucleophilic attack transition state (TS) is 8.5 kcal/mol higher than the α-HAA TS and the TS for the nucleophilic attack at the in situ generated enol species is slightly higher than the α-HAA TS by 1.2 kcal/mol. More interestingly, the same group in 2019 found that the replacement of aldehyde α-hydrogen with deuterium could switch the deformylation product of α-[D1]-CCA from conventional cyclohexanone to cyclohexanecarboxylic acid.58 This finding led them to propose a mechanism in which the peroxo ligand first extracts the aldehyde hydrogen (denoted ald-HAA mechanism, Scheme 2C).58

In 2020, Baik–Cho joint group59 and our group60 independently carried out DFT calculations to gain insight into the dioxygenase-like reactivity of [CoIII(TBDAP)(O2)]+ (denoted 3[CoO2]+, TBDAP = N,N-di-tert-butyl-2,11-diaza[3.3](2,6)-pyridinophane) toward nitriles (R–C≡N; R = Me, Et, and Ph).61 These studies showed that the reaction starts with nitrile coordination to form a five-ring metallacycle as a key intermediate, followed by O–O bond cleavage, finally leading to the activation product (vide infra). Our further analyses indicated that the O–O bond cleavage could be essentially considered as a cobalt-aided [1,3]-sigmatropic rearrangement.60 Because aldehyde may also coordinate to TM–O2 complexes, we envisioned that the deformylations with TM–O2 complexes could start with aldehyde coordination, thus initiating an inner-sphere pathway that is different from the known outer-sphere pathways (i.e., ONA,18,57 α-HAA,41 and ald-HAA58 pathways). Using the experimental reactions (eqs 1–4) as representatives,18 our DFT study demonstrated that these deformylations indeed prefer the inner-sphere mechanism (Scheme 2D) over the outer-sphere mechanisms (Scheme 2A–C).18,41,57,58 Compared to the outer-sphere mechanisms, a marked difference lies in that the inner-sphere mechanism involves homolytic aldehyde C–C bond cleavage prior to O–O bond homolytic cleavage to generate an alkyl radical. The C–C bond cleavage is enabled via O2 ligand radical activation in a way similar to a second-order nucleophilic substitution (SN2) process; thus, we term the mechanism as the radical SN2-like mechanism. Expectedly, the inner-sphere mechanism will complement the known outer-sphere mechanisms for deeply understanding the reactivity of TM–O2 complexes and offer a new way of thinking for the rational development of catalysts to perform aerobic catalytic transformations.

Computational Methods

Using experimental compounds, all structures were optimized at the B3LYP/6-31G** level in the experimental solvent (a solvent mixture with THF/dimethylformamide (DMF) = 3:1 for the 3[CuO2] system18 and 2[FeO2]2– system,57 benzene solvent for the 3[CoO2]+ system61). The solvent effects of the solvent mixture were described by the generic polarizable continuum model (PCM) solvent model with a dielectric constant of 14.88 (=ε(THF) × 0.75 + ε(DMF) × 0.25 = 14.88, where ε(THF) = 7.43 and ε(DMF) = 37.22) and those of benzene solvent were described by the PCM solvent model.62,63 Depending on the nature of a species, the B3LYP calculations could be restricted B3LYP (RB3LYP) for closed-shell singlet species, unrestricted B3LYP (UB3LYP) for open-shell species, or broken-symmetry B3LYP (BSB3LYP, implemented with the “guess = mix” keyword in the Gaussian program) for open-shell singlet species. Particular attention was paid to the closed-shell singlet species. When the wave function of a closed-shell singlet species was found to be unstable, the open-shell singlet was recalculated with BSB3LYP and reported. In the nomenclature hereafter, we specify the spin multiplicity of a structure with a left superscript. A left superscript OS represents an open-shell singlet, and no superscript means a closed-shell singlet. Harmonic frequency analysis calculations at the same level were performed to verify the optimized geometries to be minima with no imaginary frequency or transition states having unique one imaginary frequency. With the optimized structures, the energies were further improved by M06L64(PCM)/6-311++G(d,p)//B3LYP(PCM)/6-31G(d,p) single-point calculations with solvent effects simulated by the PCM solvent model using the experimental solvent (supra infra). The M06L single-point calculations maintained the wave function nature in the geometric optimizations using the keyword “guess = read” in the Gaussian program. Harmonic vibration frequencies at the B3LYP(PCM)/6-31G(d,p) level were used to correct the single-point energies to free energies at the experimental temperature (193.15 K for the 3[CuO2] system, 298.15 K for 2[FeO2]2– and 3[CoO2]+ systems). Considering that the ideal gas model used in the Gaussian program inevitably overestimates the entropy contribution to the free energy for reactions in a solvent, a −2.6 kcal/mol correction was applied for each component decrement (i.e., correcting the overestimation of the entropy penalty) and vice versa.6569

The crystal structure of 3[CuO2] has not been obtained, but that of the precursor [Cu]-(4-tBu-pyridine) was reported.20 In addition, 3[CuO2] was characterized to have a triplet ground state, an O–O bond vibrational frequency of 1104 cm–1, and a characteristic UV–vis absorption at 627 nm.20 Using these relevant experimental results, we evaluated the reliability of our protocol by comparing to other protocols (M06L(PCM)//B3LYP-D3(PCM), M067073(PCM)//B3LYP-D374(PCM), B3LYP-D3(PCM)//B3LYP-D3(PCM), MN1575(PCM)//MN15(PCM)) with different basis set combinations (6-311++G**//6-31G** or Def2TZVP7678//Def2SVP7679). Detailed results in the Supporting Information (Section SI1) indicate that the protocol (M06L(PCM)/6-311++G**//B3LYP(PCM)/6-31G**) has the best overall performance in reproducing these available experimental results. In addition, we also confirmed that the reaction system could not involve [Cu]–O–O–[Cu] species (see Section SI2 for details).

The natural bond orbital (NBO) charges,8083 the Mulliken spin populations, and biorthogonal molecular orbitals (BMOs) were obtained at the B3LYP(PCM)/6-31G(d,p) level. All standard DFT calculations were performed using the Gaussian 09 program.84

Results and Discussion

Deformylation Mechanism of 2-PPA with 3[CuO2]

The ONA mechanism was often applied to account for the deformylations with TM–O2 complexes as deformylating reagents.18,31,4448 We first investigated whether 3[CuO2] could act similarly. At a glance, because of the anionic nature of 3[CuO2], the ONA mechanism appears reasonable. Nevertheless, the nucleophilic attack to give 3IM2aCu is highly endergonic by 24.5 kcal/mol (Figure 1). The geometric optimization to locate a singlet counterpart of 3IM2aCu led to separate OS[CuO2] and 2-PPA. These results suggest that the ONA mechanism may not operate in this reaction. We next considered the α-HAA and ald-HAA mechanisms. Similar to the previous finding,41 the α-HAA TSs, OS,3TS1aCuG = 18.0/15.9 kcal/mol), are substantially lower than 3IM2aCuG = 24.5 kcal/mol). The ald-HAA TSs, OS,3TS1bCuG = 23.4/22.3 kcal/mol), are higher than the α-HAA TSs but still lower than 3IM2aCu.

Figure 1.

Figure 1

Reaction pathway for the deformylation of 2-PPA (eq 1), along with the relative free energies in kcal/mol. The values in black and orange in 3IM2Cu and 3TS2Cu are the bond lengths in angstroms and NBO charges in e, respectively. The −1 charge states of the Cu-containing structures were omitted for clarity.

Essentially, the three processes (ONA, α-HAA, and ald-HAA) take place in an outer-sphere fashion without the participation of the copper center. Although the α-HAA and ald-HAA TSs are not inaccessible, our previous study60 inspired us to explore whether the reaction could undergo an inner-sphere mechanism, starting with carbonyl coordination to the copper center. Compared to the outer-sphere mechanism, we foresaw that the inner-sphere mechanism could be benefited from the coordination interaction and Cuδ+···Oδ− Coulomb attraction. Figure 1 illustrates our computed pathways. First, the carbonyl group of 2-PPA undergoes end-on coordination with 3[CuO2] by crossing a low barrier of 2.8 kcal/mol (3TS1Cu), forming a slightly stable intermediate 3IM2Cu. Proceeding forward, two pathways were examined. The pathway to break the O–O bond, giving 3IM3aCu, is highly endergonic by 37.8 kcal/mol, unambiguously excluding the possibility. The alternative breaks the aldehyde C4–C5 bond of 2-PPA. Remarkably, the TS (3TS2Cu) to break the bond is 24.0 kcal/mol lower than 3IM3aCu. In addition, we considered the two-state reactivity mechanism85 to break the C4–C5 bond, but the open-shell singlets OSIM2Cu and OSTS2Cu are higher than their corresponding triplet counterparts. The energetic difference of 1.7 kcal/mol between OSTS2Cu and 3TS2Cu indicates that the two-state reactivity can only play a minor role if any.

The homolytic C4–C5 bond cleavage affords 2IM3Cu and a benzyl radical R1. The bond cleavage substantially elongates the O1–O2 bond from 1.31 Å in 3IM2Cu to 1.45 Å in 2IM3Cu; thus, the O–O bond can break with a barrier of 11.2 kcal/mol (2TS3Cu relative to 2IM3Cu). In comparison, intermediate 3IM3aCu led by directly breaking the O1–O2 bond of 3IM2Cu has relative energy of 37.8 kcal/mol. Therefore, the C4–C5 bond cleavage greatly facilitates the O–O bond cleavage. Furthermore, we considered whether the O1–O2 bond cleavage could be facilitated by associating with the R1 radical, as described by OS,3TS3aCu; however, both TSs are higher than 2TS3Cu, suggesting that the association of R1 cannot help the O1–O2 bond cleavage. It is worth noting that, referring to OS,3TS3aCu, the bulky ligand of the 3[CuO2] complex prevents R1 associating with O1.

The O–O bond cleavage leads to 2IM4Cu, in which O1 bears a single electron. Thus, the R1 radical couples with O1, giving IM5Cu. Finally, IM5Cu undergoes intramolecular hydrogen atom transfer via OSTS4Cu, affording the experimental products acetophenone and formic acid. Examining the inner-sphere pathway, the ketone oxygen atom originates from 3[CuO2], which agrees with the isotope labeling experiment with 3[Cu18O2] as the deformylating reagent. To our delight, a complex similar to 3IM2Cu has been characterized spectroscopically for the deformylation of 2-PPA with an analogue of 3[CoO2]+ during our preparation of the paper, which could be an experimental evidence to support our proposed mechanism (see the box in Figure 1).86

Overall, the reaction is exergonic by 71.3 kcal/mol with a rate-determining barrier of 14.7 kcal/mol for breaking the aldehyde C–C bond, indicating that the reaction could take place facilely. The rate-determining 3TS2CuG = 13.8 kcal/mol) is lower than the intermediate 3IM2aCuG = 24.5 kcal/mol) led by nucleophilic attack; OS,3TS1aCuG = 18.0/15.9 kcal/mol), 3TS2aCuG = 20.4 kcal/mol), and 3TS3bCuG = 20.5 kcal/mol) involved in the α-HAA pathway; and OS,3TS1bCuG = 23.4/22.3 kcal/mol) for the ald-HAA process. The favorability of the inner-sphere mechanism could be attributed to the favorable interactions (coordination and Coulomb attraction) between the copper center and the carbonyl oxygen of 2-PPA (see the NBO charges and Cu···O3 distances given in 3IM2Cu and 3TS2Cu) and the high stability of the benzyl radical (vide infra). Compared to the ONA and α-HAA mechanisms in which the O–O bond breaks prior to the C–C bond cleavage, the inner-sphere mechanism first breaks the C–C bond, which facilitates subsequent O–O bond cleavage.

To corroborate the favorability of the inner-sphere mechanism, we further compared the energetics of the four mechanisms at the other five levels of DFT calculations (see Section SI3 for detailed results). In spite of some numerical differences, the energetic results from the six levels of calculations consistently indicate that the inner-sphere mechanism is most favorable.

Comparing the Deformylation Mechanism of 3[CuO2] and Acetonitrile Activation Mechanism of 3[CoO2]+

Acetonitrile activation with 3[CoO2]+ involves O–O bond cleavage but not C–C bond breaking.59,60 It is insightful to compare the key steps of the two systems. In the following, we gain insight into the differences and similarities of the two systems (Figure 2).

Figure 2.

Figure 2

Comparing the key steps involved in 2-PPA deformylation with 3[CuO2] and acetonitrile activation with 3[CoO2]+. For clarity, the −1 charge states for Cu-containing structures and +1 charge states for Co-containing structures were omitted.

First, the O2 ligand in 3[CuO2] adopts end-on coordination with unequal Cu–O bond lengths (1.91 and 2.55 Å, respectively), while the O2 ligand in 3[CoO2]+ features side-on coordination with equal Co–O bond lengths (ca. 1.88Å). The coordination difference indicates that a Cu(II) oxidation state is preferred in 3[CuO2], while a Co(III) oxidation state is preferred in 3[CoO2]+, although Cu(III)–peroxo complexes were synthesized with strongly electron-donating bidentate-diketiminate or anilido-imine ligands.87,88 Due to the coordination difference, the oxygen ligand in 3[CuO2] bears a single electron (Mulliken spin population: O2 1.17), while the ligand in 3[CoO2]+ possesses no radical character with the two single electrons residing at the cobalt center (Mulliken spin populations: Co 1.88, O2 0.02). However, the O2 ligand in 3[CoO2]+ can gain a single electron upon acetonitrile coordination, as indicated by the Mulliken spin population of O2 (0.94) in 3IM1Co. The Cu(II) oxidation state preference explains the general feature that the Cu center in the intermediate of the 3[CuO2] system only interacts with one of the two neighbor oxygen atoms tightly, while the Co center in the intermediate of the 3[CoO2]+ system can interact with neighboring two atoms tightly.

Second, 3IM1Cu converts to a loose five-ring structure 3IM2Cu, while 3IM1Co isomerizes to a tight five-ring metallacycle 3IM2Co. Indicated by the Mulliken spin populations and the O–O bond length in 3IM2Cu, the O2 ligand in 3IM2Cu maintains its radical and partial double bond character in 3IM1Cu. In contrast, the O2 ligand in 3IM2Co bears no single electron with the two single electrons located at the Co center and the O–O bond is a regular single bond. In addition, the O2–C4 bond in 3IM2Cu is only partially formed with a bond length of 1.65 Å, while the bond in 3IM2Co is a well-formed C–O single bond with a bond length of 1.36 Å. To gain deeper insight into the origin of these differences, we first go over the isomerization mechanism of 3IM1Co to 3IM2Co, proposed in our previous study (Figure 3).60 To begin with, the single α-electron located on the O2 ligand in 3IM1Co activates an acetonitrile N≡C π-bond, forming the C–O σ-bond in 3IM2Co and leaving a single α-electron on the N=C double bond. Then, the single α-electron on the N=C double bond interacts with a pair of nonbonding electrons on Co to form the Co–N bond in 3IM2Co. With this mechanism established, we understand why 3IM1Cu is converted to 3IM2Cu, rather than an intermediate similar to 3IM2Co. Figure 4A compares the key relevant biorthogonal molecular orbitals (BMOs) of 3IM1Cu and 3IM1Co. The two intermediates have similar singly occupied BMOs (SOBMOs) dominated by O2 ligands (see SOBMO189Cu/SOBMO128Co), which are approximately perpendicular to the Cu–O–O/Co–O–O plane. In 3IM1Cu, the C=O π-bond orbital of 2-PPA (DOBMO137Cu, doubly occupied BMO) lying in the Cu–O–O plane is nearly perpendicular to SOBMO189Cu. Therefore, the O2 ligand radical cannot activate the C=O bond effectively, preventing the single-electron activation mode (see box C in Figure 4B). Nevertheless, the carbonyl group can be activated through coordination activation, as illustrated by the electron flow chart in box A (Figure 4B). In principle, the full activation of the C=O π-bond would result in 3IM2Cu-RA; however, the activation may not be so effective. Overall, 3IM2Cu can be represented by the two resonance structures 3IM2Cu-RA and 3IM2Cu-RB, without involving 3IM2Cu-RC. The resonance structure description agrees with the relatively long lengths of Cu–O1 (R = 2.34Å), C4–O2(1.65 Å), and O3=C4 (1.30 Å) and shortened Cu–O3 (1.92 Å) in 3IM2Cu. In contrast to the 2-PPA unit in 3IM1Cu, the acetonitrile unit in 3IM1Co has two π-bonds (see DOBMO100Co and DOBMO112Co in Figure 4A). The perpendicular DOBMO100Co is nearly parallel to the single-electron orbital (SOBMO128Co), which is favorable for single-electron activation. Thus, single-electron activation can take place effectively to convert 3IM1Co to 3IM2Co with the mechanism described by 3TS2Co in Figure 3.

Figure 3.

Figure 3

Mechanism for the isomerization of 3IM1Co to 3IM2Co. For clarity, the +1 charge states for Co-containing structures were omitted.

Figure 4.

Figure 4

(A) Comparing the singly occupied biorthogonal molecular orbitals of 3IM1Cu and 3IM1Co. (B) Mechanism for converting 3IM1Cu to 3IM2Cu. For clarity, the −1 charge states for Cu-containing structures and +1 charge states for Co-containing structures were omitted.

Third, 3IM2Cu undergoes aldehyde C4–C5 bond cleavage, giving an alkyl radical, while 3IM2Co breaks the O1–O2 bond directly (Figure 2). A straightforward explanation for the difference would be that the O–O bond in the former is a partial double bond (R = 1.31Å), while the bond in the latter is a single bond (R = 1.46Å). Previously, we have shown that the O–O bond cleavage is essentially a cobalt-aided one-electron [1,3]-sigmatropic rearrangement (for electron flow chart description, see Figure 2). In the following, we analyze the mechanism for breaking the C–C bond in 3IM2Cu. As shown in Figure 1, 3IM2Cu cannot break the O–O bond because the process to afford 3IM3aCu is highly endergonic by 37.8 kcal/mol. The high energy of 3IM3aCu can be attributed to its unfavorable characteristics such as the disfavored Cu(III) oxidation state and hypervalent carbon. Therefore, an approach to circumvent the hypervalent carbon problem is to break the aldehyde C4–C5 bond before O–O bond cleavage, which can be achieved through the activation of the O2 ligand radical. Figure 5A illustrates the principle for 3IM2Cu to break the C–C bond via 3TS2Cu. The C4–C5 σ-bond orbital in 3TS2Cu points outward the five-ring plane, which is symmetrically suitable for the interaction of the C4–C5 σ-bond orbital with the single-electron orbital of the O2 ligand, as depicted by SOBMO189Cu of the bond cleavage TS 3TS2Cu. As illustrated by the electron flow chart of 3TS2Cu, the bond cleavage takes place in a way similar to an SN2 process but is enabled by single-electron activation of the O2 ligand radical, rather than an electron pair of a nucleophile in the conventional SN2 reaction. Because of the bond cleavage feature and the step is the rate-determining step of the deformylation mechanism, we propose to term the inner-sphere deformylation mechanism as the radical SN2-like mechanism.

Figure 5.

Figure 5

(A) Mechanism to break the C–C bond in 3TS2Cu. (B) Mechanism to break the C–C bond in 3TS1BaCo (see Figure 6 for 3TS1BaCo). The −1 charge state of 3TS2Cu and +1 charge states of 3TS1BaCo were omitted for clarity.

As compared in Figure 2, intermediate 2IM3Cu led by C4–C5 bond cleavage is similar to 3IM2Co, except for the Cu–O3 coordination bond in the former, while the Co–N covalent bond in the latter (due to the different preference of oxidation states, Cu(II) versus Co(III)). In particular, the ligand O1–O2 bond in 2IM2Cu becomes a single bond (R = 1.45 Å) in 2IM3Cu; thus, the O–O bond can break via a mechanism depicted by the curly arrows in 2TS3Cu. Comparing the Mulliken spin populations of 2IM4Cu and 3IM3Co, the O–O bond cleavage of 2IM3Cu leads to an α-electron located on O1, while the process of 3IM2Co leads to an α-electron and a β-electron located at O1 and N, respectively. Thus, 3IM3Co undergoes intramolecular radical–radical coupling to form the N–O bond, completing a [1,3]-sigmatropic rearrangement to give 3IM4Co. In comparison, 2IM4Cu undergoes intermolecular radical–radical coupling with R1, giving IM5Cu for subsequent processes to complete the deformylation. The occurrence of the formal Cu(III) oxidation state in 2IM4Cu can be ascribed to the strong π-donor nature of the oxo ligand. On the other hand, the Cu–O3 distance (2.18 Å) is substantially longer than that (1.93 Å) in the X-ray structure of a Cu(II) complex [CuOC(O)H] supported by a similar ligand in [CuO2],89 indicating that the Cu–O3 bond is only a weak covalent bond and 2IM4Cu is not a well-formed Cu(III) species.

After understanding the differences and similarities between 2-PPA deformylation with 3[CuO2] and acetonitrile activation with 3[CoO2]+, we further questioned whether 3[CoO2]+ could also promote deformylations via a similar mechanism, whether 3[CuO2] could activate acetonitrile, and whether the benzyl substituent of 2-PPA plays a role in C–C bond cleavage. To answer these questions, we computed the key steps for the reactions of 3[CoO2]+ with 2-PPA and 2-phenylpropanenitrile (N≡CCH(Me)Ph, 2-PPN), respectively (Figure 6). For the reaction of 3[CoO2]+ with 2-PPA, the quintet pathway shown in Figure 6 is slightly more favorable than the triplet pathway (see Section SI4 for detailed results). The quintet results are discussed below and the discussion holds true for the triplet case. The C–C bond cleavage is similar to that of 3[CuO2] (see Figure 2). First, the coordination complex 5IM1ACo is similar to 3IM2Cu. Second, 5IM1ACo also favors C–C bond cleavage over O–O bond cleavage, albeit to less extent, compared to the case of 3[CuO2]. As 3IM3aCu is 24.0 kcal/mol higher than C–C bond cleavage TS 3TS2Cu, the energy difference between 5IM2ACo and 5TS1ACo is much smaller (5.3 kcal/mol). The large difference can be attributed to the different preferred oxidation states of Cu(II) and Co(III) because the preferred Co(III) oxidation state in 5IM2ACo avoids the unfavorable effect in 3IM3aCu due to the disfavored Cu(III) oxidation state.

Figure 6.

Figure 6

Key steps for the reactions of 3[CoO2]+ with 2-PPA and 2-PPN, along with the relative energies in kcal/mol. The +1 charge states of the Co-containing structures were omitted for clarity.

For the reaction of 3[CoO2]+ with 2-PPN, the five-ring metallacycle 3IM1BCo is similar to 3IM2Co, as shown in Figure 2. Interestingly, 3IM1BCo could break its O–O bond via two reaction modes. The mode depicted by 3TS1BbCo is similar to that of 3TS3Co in acetonitrile activation, as shown in Figure 2, but the mode described by 3TS1BaCo has no counterpart, which breaks O–O and C–C bonds concertedly. Energetically, the latter mode is slightly more favorable than the former. Considering that the aldehyde C–C bond cleavage in 3IM2Cu is activated by the single electron on the O2 ligand (Figure 5A), it is interesting to note that the O2 ligand in 3IM1BCo (Figure 6) possesses no single electron and the N=C bond lies in the ring plane, but 3IM1BCo can also undergo C–C bond cleavage, which we understand as follows. Referring to Figure 5B, the homolytic O1–O2 bond cleavage results in O2 with β radical character (denoted (O2) hereafter). Radical (O2) could activate either of its neighboring C4–C5 bond or C4=N3 bond, depending on which activation is more feasible. If the substituent is a good radical-leaving group such as a benzyl group (vide infra), the (O2) radical would prefer to activate the C–C bond (see SOBMO151Co of 3TS1BaCo), forming the C4=O2 double bond and releasing an alkyl radical. Referring to Figure 2, if the substituent is a poor radical-leaving group such as a methyl group in 3IM2Co, the (O2) radical would prefer to act on the C=N π-bond, forming a C4=O2 double bond and leaving a single electron on the nitrogen center in 3IM3Co. This activation mechanism explains why the C–C and O–O bond cleavages of 3IM1BCo takes place concertedly because there is a need to generate the reactive (O2) radical for the subsequent C–C or C=N bond activation. As demonstrated by the much more stable 4IM2BCo than 3IM2BCo, the C–C bond cleavage to release a benzyl radical has the advantage to avoid the destabilizing effect due to the location of a single electron on nitrogen to maintain the preferred Co(III) oxidation state. These studies indicate that the benzyl substituent of 2-PPA and 2-PPN plays a role in allowing the C–C bond cleavage, which will be analyzed in detail later.

The activation of acetonitrile and 2-PPN with 3[CuO2] was also examined. However, no intermediate similar to either 3IM2Cu or 3IM2Co could be located, probably due to the following reason. Because the anionic 3[CuO2] is less electrophilic than the cationic 3[CoO2]+, the former could not activate nitrile as effectively as the latter through nitrile coordination, which makes the N≡C π bond activation difficult for the O2 ligand radical of 3[CuO2].

On the basis of the discussion above, we envision that 3[CoO2]+ could react with 2-PPA but 3[CuO2] could not activate acetonitrile. Indeed, it was reported that an analogue of 3[CoO2]+ could deformylate 2-PPA during our preparation of the paper.86 A five-membered intermediate characterized spectroscopically (see the box in Figure 1) is similar to our 3,5IM1ACo. In addition, we predict that 3[CoO2]+ may react with 2-PPN in a way different from acetonitrile activation and call experimental verification.

Deformylation Mechanisms of TMA/CCA with 3[CuO2]

2-PPA has an α-H atom with ketone as the deformylation product. We further computed the deformylation pathways of aldehyde TMA without an α-H atom. In addition, CCA (with an α-H atom) deformylation also afforded a ketone product, providing another case to corroborate our mechanism. Figure 7 shows our computed pathways for the deformylations of TMA and CCA. Proceeding along the blue pathway (PathATMA/CCA), TMA/CCA first undergoes a process similar to that of 2-PPA, breaking the aldehyde C–C bond via 3TS2TMA/CCA to generate 2IM2TMA/CCA and an alkyl radical R2/R3. However, the subsequent processes are different. First, the O–O bond cleavage in the case of 2-PPA takes place directly without the participation of the benzyl radical generated from C–C bond cleavage. Differently, in the case of TMA/CCA, the association of the R2/R3 radical with the O2 atom lowers the O–O bond cleavage barriers by 3.6/3.2 kcal/mol (the energetic difference between 2TS3aTMA/CCA and 3TS3TMA/CCA). Note that, due to the steric hindrance of the ligand, the alkyl radical cannot associate with the O1 atom. We will rationalize the difference in the next section. Second, because of the above difference, the subsequent processes are different. For CCA, due to the association of the alkyl radical R3 in O–O bond cleavage, the process results in cyclohexyl formate (IM3CCA) and copper-oxo species 3[CuO]. The two species then react via a sequence including H-atom abstraction via 3TS4CCA and nucleophilic attack via 3TS5CCA, giving 3IM5CCA. Finally, the cleavage of the C–O bond via 3TS6CCA leads to the experimental products cyclohexanone and formic acid (eq 3). Note that, an alternative to the nucleophilic attack, 2IM4CCA and radical (IM3CCA) could undergo rebound via 3TS5aCCA, but 3TS5aCCA is significantly higher than 3TS5CCA and 3TS6CCA, excluding the possibility. Because TMA does not have an α-H atom, the reaction terminates after O–O bond cleavage, giving the experimental product, tert-butyl formate (eq 2).

Figure 7.

Figure 7

Deformylation pathways of TMA and CCA, along with the relative free energies in kcal/mol. The −1 charge states of the Cu-containing structures were omitted for clarity.

We also examined the outer-sphere ONA, α-HAA, and ald-HAA pathways. The nucleophilic attack TS (3TS1bTMA/CCA, ΔG = 24.1/26.3 kcal/mol) is much higher than the rate-determining C–C bond cleavage TS (3TS2TMA/CCA, ΔG = 16.5/21.2 kcal/mol). PathBCCA describes the α-HAA pathway. Although the α-HAA TS (OS,3TS1aCCA, ΔG = 20.7, 21.4 kcal/mol) is slightly lower than the rate-determining barrier 3TS2CCAG = 21.2 kcal/mol) for C–C bond cleavage, the mechanism encounters a high barrier 3TS3aCCAG = 30.2 kcal/mol) for the nucleophilic attack of 3[CuO2] at the enol intermediate. PathCCCA describes the ald-HAA pathway of CCA. Although the ald-HAA TS (3TS1cCCA) is only slightly higher than the rate-determining barrier TS 3TS2CCA, the recombination TS (OS,3TS2cCCA, ΔG = 31.8, 32.9 kcal/mol) is much higher than 3TS2CCAG = 21.2 kcal/mol). In the case of TMA, the only alternative is ald-HAA, but 3TS1aTMA for the process is 5.6 kcal/mol higher than 3TS2TMA. These results exclude the outer-sphere ONA, α-HAA, and ald-HAA mechanisms for the two deformylation reactions, further affirming the preference of the inner-sphere mechanism.

Key Factors to Enable the Radical SN2-like Deformylation

In addition to 2-PPA (A1), TMA (A2), and CCA (A3), McDonald et al.18 also carried out deformylations of other aldehydes including propionaldehyde (A4), acetaldehyde (A5), and benzaldehyde (A6), but no reactions were observed for A4A6 (eq 4). These experimental results encouraged us to find a controlling parameter to enable deformylation at the aldehyde end. On the basis of our proposed inner-sphere mechanism involving aldehyde C–C bond cleavage as the rate-determining step to generate an alkyl radical, we reasoned that the stability of the radical could play an important role in promoting the aldehyde C–C bond cleavage. As shown in Figure 8, we calculated the bond cleavage barriers (ΔG) and the stabilization free energies (ΔG) of the formed radicals relative to benzyl radical R1 (see Section SI5 for details on how to estimate ΔG). Remarkably, the barriers ΔG are linearly correlated to the stabilization energies ΔG of the radicals with a regression coefficient of 0.95. The excellent linear correlation of ΔG to ΔG demonstrates that the stability of the radical originating from the aldehyde substituent is a key factor for the deformylation reactions. Because radicals R4R6 are much less stable than R1–R3, the rate-determining barriers of A4–A6 are significantly higher than those of A1A3, explaining the experimental observations of no reaction for A4–A6 (eq 4). In addition, the stability of the radical explains why benzaldehyde (A6) could not undergo deformylation, while TMA (A2) could. Because the corresponding phenyl radical (R6) of benzaldehyde is 17.6 kcal/mol less stable than the tert-butyl radical (R2) of TMA, the former has much higher C–C bond cleavage barrier than the latter (33.5 versus 16.5 kcal/mol). Experimentally, various derivatives of A6 (p-R-PhC(O)H, R = NMe2, OMe, Me, Cl, NO2) were also attempted for their deformylations, but no reaction was observed.18 Our calculations showed that the radicals corresponding to these derivatives have relative stabilization energies higher than that of A6 (>25.4 kcal/mol, see Section SI5). Therefore, these derivatives would have rate-determining barriers higher than that (33.5 kcal/mol) of A6, explaining the observations of no reaction for these A6 derivatives. The rationalizations of these experimental observations support our proposed radical SN2-like mechanism. Moreover, the stability of the radical also elucidates the aforementioned different O–O bond cleavage mechanisms of TMA and CCA from that of 2-PPA. The corresponding radical (R1) of 2-PPA is much more stable than R2 and R3 of TMA and CCA by 7.8 and 12.5 kcal/mol, respectively. Thus, in the case of 2-PPA, the entropic penalty due to associating a benzyl radical to 2IM3Cu could override the energetic benefit of radical–radical association (see the process from 2IM3Cu to OS,3TS3aCu in Figure 1).

Figure 8.

Figure 8

Relationship between the rate-determining C–C bond cleavage barrier and the stability of the radicals relative to benzyl radical.

Next, we considered the geometric factor of TM–O2 complexes in enabling deformylations via the inner-sphere mechanism. Examining the structures of 3[CuO2] and 3[CoO2]+, the inner-sphere mechanism requires TM–O2 complexes to have available room for aldehyde coordination. In this context, we computed 2-PPA deformylation with the porphyrin-ligated 2[FeO2]2– model complex of CYP51 (Figure 9A). Previously, Hackett et al. demonstrated the feasibility of the ONA mechanism.57 In agreement with their study, our calculations predicted that [FeO2]2– favors a low-spin doublet state and the quartet is 6.4 kcal/mol higher than the doublet. Not surprisingly, the geometric optimization to locate a coordination complex similar to 3IM2Cu or 3IM1Co or 3IM2Co between 2[FeO2]2– and 2-PPA drove the two parts apart due to the steric hindrance of the rigid planar porphyrin ligand. Thus, the outer-sphere mechanisms (ONA, α-HAA, and ald-HAA) are possible alternatives. For the ONA mechanism, the nucleophilic attack intermediate 2IM2aFe has relative free energy of 7.2 kcal/mol, which is significantly lower than that (24.5 kcal/mol) of the copper counterpart 3IM2aCu. The barrier 2TS2aFeG = 18.8 kcal/mol) to break the aldehyde C–C bond of 2IM2aFe is also lower than that of the copper counterpart 3TS2bCuG = 22.6 kcal/mol). Furthermore, we examined whether the reaction could undergo α-HAA and ald-HAA mechanisms (Figure 9B). For the α-HAA mechanism, the 2TS2bFeG = 9.6 kcal/mol) is not high, but the pathway needs to cross a high barrier of 23.8 kcal/mol (2TS3bFe) for nucleophilic attack at the enol intermediate (referring to Scheme 2B). The 2TS2cFe for ald-HAA reaches 28.4 kcal/mol. These barriers are higher than 2TS2aFeG = 18.8 kcal/mol), indicating that the ONA mechanism is most favorable.

Figure 9.

Figure 9

(A) Comparing the outer-sphere nucleophilic attack of 3[CuO2] and 2[FeO2]2– at 2-PPA. (B) Key TSs involved in 2-PPA deformylation with 2[FeO2]2– via the α-HAA mechanism (2TS2bFe and 2TS3bFe) and ald-HAA mechanism (2TS2cFe). (C) Outer-sphere nucleophilic attack of 3[CuO2] with a strong nucleophile benzoyl chloride. The values in the parentheses are free energies (in kcal/mol) relative to reactants.

Figure 9A also compares the outer-sphere nucleophilic attacks of 2[FeO2]2– and 3[CuO2] at 2-PPA. The NBO charge (q) and Mulliken spin population results of 2IM2aFe and 3IM2aCu indicate that the processes are essentially different. As the carbonyl oxygen in 2IM2aFe is an O anion with a Mulliken spin population of 0.00 and charge of −0.92 e, that in 3IM2aCu is an O radical with a Mulliken spin population of 0.67 and charge of −0.46 e. In addition, the nucleophilic attack of 2[FeO2]2– significantly increases the Mulliken spin population at the Fe center from 0.31 in 2[FeO2]2– to 0.92 in 2IM2aFe, while the process of 3[CuO2] only slightly increases the Mulliken spin population at the Cu center from 0.57 to 0.62. According to the Mulliken spin population and NBO charge results of 2IM2aFe, we reasoned that the nucleophilic attack of 2[FeO2]2– takes place via a two-electron mechanism, while the process of 3[CuO2] proceeds via a one-electron mechanism, as compared in the black boxes. Due to the difference, as shown by the Mulliken spin population results of 3TS2bCu and 2TS2aFe, the C–C bond cleavage of 3IM2aCu is homolytic, giving benzyl radicals, while the process of 2IM2aFe is heterolytic, giving benzyl anions. In addition to the geometric factor, the stronger nucleophilicity of the dianionic 2[FeO2]2– than the monoanionic 3[CuO2] should also contribute to these differences.

In the radical SN2-like process, the leaving group is an alkyl radical. Intrigued by the question of what if the substrate is more electrophilic than 2-PPA and bears a good anionic leaving group, in this context, we further studied the nucleophilic attack of 3[CuO2] at benzoyl chloride. Experimentally, it has been found that 3[CuO2] could react with acyl chlorides.18Figure 9C shows the intermediates formed via inner-sphere (CS,3IM2Cu–Cl, CS denoting closed-shell singlet) and outer-sphere (CS,3IM2aCu–Cl) mechanisms, respectively. Interestingly, regardless of the inner- or outer-sphere mechanisms, the attacks are conventional SN2 processes, breaking the C–Cl bond heterolytically; no intermediate similar to 3IM2Cu (Figure 2) or 3IM2aCu (Figure 9A) could be formed. In addition, the intermediates prefer singlet as the ground state (CSIM2Cu–Cl/CSIM2aCu–Cl is 15.3/13.4 kcal/mol lower than their triplet counterparts), which is different from the attack of 3[CuO2] at 2-PPA. Although the nucleophilic attack with benzoyl chloride is different from that with 2-PPA, the inner-sphere pathway is still favored, CSIM2Cu–Cl/3IM2Cu–Cl being lower than CSIM2aCu–Cl/3IM2aCu–Cl. The preference could be attributed to the Coulomb attraction and the weak coordination between the positive copper center and the carbonyl oxygen in CS,3IM2Cu–Cl. The results demonstrate that the inner-sphere pathway, which could take the advantage of the favorable interaction between the metal center and the functional group of substrates, should not be overlooked even if a deformylation reaction proceeds via the conventional nucleophilic attack.

Summarizing the discussion in this section, two factors could be crucial for TM–O2 complexes to undergo the inner-sphere mechanism with aldehydes. First, TM–O2 complexes feature the coordination environment that allows aldehyde end-on coordination to the metal center. Second, aldehydes feature a substituent corresponding to a relatively stable radical as a good radical-leaving group.

Rules for the Reactions of TM–O2 Complexes with Aldehydes or Nitriles via the Inner-Sphere Mechanism

On the basis of the present and previous studies,59,60 we reasoned that the reactions of TM–O2 complexes with aldehydes or nitriles may follow the rules sketched in Scheme 3 if the reactions undergo the inner-sphere mechanism. We clarify that the inner-sphere mechanism does not rule out the possibility that some TM–O2 complexes may favor outer-sphere mechanisms, as exemplified by the [FeO2]2– complex in Figure 9.

Scheme 3. Rules for the Reactions of TM–O2 Complexes with Aldehydes and Nitriles via the Inner-Sphere Radical SN2-like Mechanism.

Scheme 3

The inner-sphere mechanism begins with the end-on coordination of the substrate aldehydes/nitriles to the metal center of TM–O2, forming coordination complexes IMI/IMIV. If TM–O2 is a metal–peroxo complex, the coordination renders the O2 ligand to gain a single electron.

For the reactions of TM–O2 with aldehydes, subsequent to the formation of IMI, if the radical R corresponding to aldehyde substituent is a poor radical-leaving group (RIII-type), the aldehyde C–C bond cannot be broken and the inner-sphere mechanism cannot proceed. If R is a good radical-leaving group (RI- and RII-type), the O2 ligand radical could activate the aldehyde C–C bond, generating an alkyl radical R and an intermediate IMII. Subsequently, the O–O bond cleavage takes place via two possible modes. If the R radical like a benzyl radical (RI-type) is very stable, the O–O bond breaks directly via TSI-type TS (e.g., 2TS3Cu in Figure 1), giving IMIII. Then, radical R and IMIII couples, followed by intramolecular hydrogen abstraction via TSa, finally affording the products, ketone and formic acid. If R is medially stable (RII-type), the association of radical R with O could aid O–O bond cleavage, taking place via TSII-type TS (e.g., 3TS3TMA/CCA in Figure 7). The O–O bond cleavage leads to formate ester. If the aldehyde (e.g., TMA) does not have an α-H, the reaction stops at the step and gives formate ester as the final product. If the aldehyde possesses an α-H, the M=O species abstracts the α-H in the formate ester via TSb (e.g., 3TS4CCA in Figure 7), giving MOH species, then the reaction affords the product.

For the reaction of TM–O2 with nitriles, the coordination complex IMIV first isomerizes to a five-ring metallacycle IMV via O2 ligand radical activation of the nitrile group, forming the metal–N and C–O bonds, followed by O–O bond cleavage. If the R group of a nitrile is a poor radical-leaving group like RIII-type, the O–O bond breaks via TSIII-type TS (e.g., 3TS3Co in Figure 2), followed by intramolecular radical–radical coupling, finally leading to the nitrile activation product. If the R group is a good radical-leaving group like RI-type, IMV breaks O–O and C–C bonds concertedly via TSIV-type TS (e.g., 3TS1BaCo in Figure 6). For the nitriles with RII-type substituent, TSIII- and TSIV-type modes would compete. To our knowledge, the activation of nitrile with RI-type substituents has not been studied experimentally. It would be interesting to characterize the final products, and we invite experimental study to verify our rules.

Conclusions

In summary, we have performed DFT study of the deformylation mechanism of aldehydes with copper(II)–superoxo as a deformylating reagent. In contrast to the known outer-sphere mechanisms, including nucleophilic attack, α-hydrogen atom abstraction, and aldehyde hydrogen atom abstraction, the study found that the present deformylations prefer an inner-sphere mechanism that could be termed as the radical SN2-like mechanism. The mechanism proceeds via a sequence mainly consisting of end-on aldehyde coordination, aldehyde C–C bond cleavage, O–O bond cleavage, and α-hydrogen abstraction, finally giving ketone or formate ester as the product. The aldehyde C–C bond cleavage is the rate-determining step, which is achieved by the activation of the O2 ligand radical via an SN2-like fashion. The C–C bond cleavage greatly facilitates the subsequent O–O bond cleavage. Because the rate-determining step results in an alkyl radical corresponding to an aldehyde substituent, the stability of the alky radical is a key factor to enable the inner-sphere mechanism. In addition, the metal–dioxygen complex should possess a coordination environment suitable for aldehyde end-on coordination. Taking the deformylation mechanism and the mechanism for acetonitrile activation with cobalt(III)–peroxo into considerations together, we have derived reaction rules for deformylations and nitrile activations with metal–dioxygen complexes and identified four modes for breaking the O–O bond in metal–dioxygen complexes, depending on the properties of the substrates and the metal–dioxygen complexes. This study demonstrates that the inner-sphere mechanism could be favored over the known outer-sphere mechanism and should not be overlooked when studying the reactivity of TM–O2 complexes. Expectedly, the inner-sphere mechanism, as well as the reaction rules, will complement the known outer-sphere mechanisms, thus offering a new way of thinking for understanding the reactivity of metal–dioxygen core-related biological processes such as deformylations and the discovery of metal–dioxygen catalysts for catalytic aerobic transformations.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.2c00014.

  • Additional results, total energies, and Cartesian coordinates of optimized structures (Sections SI1–SI6) (PDF)

This research was supported by the National Science Foundation of China (Grant Nos. 22173103 and 21773240).

The authors declare no competing financial interest.

Supplementary Material

au2c00014_si_001.pdf (1.6MB, pdf)

References

  1. Jasniewski A. J.; Que L. Dioxygen Activation by Nonheme Diiron Enzymes: Diverse Dioxygen Adducts, High-Valent Intermediates, and Related Model Complexes. Chem. Rev. 2018, 118, 2554–2592. 10.1021/acs.chemrev.7b00457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Guengerich F. P. Common and Uncommon Cytochrome P450 Reactions Related to Metabolism and Chemical Toxicity. Chem. Res. Toxicol. 2001, 14, 611–650. 10.1021/tx0002583. [DOI] [PubMed] [Google Scholar]
  3. Hargrove T. Y.; Wawrzak Z.; Guengerich F. P.; Lepesheva G. I. A requirement for an active proton delivery network supports a compound I-mediated C–C bond cleavage in CYP51 catalysis. J. Biol. Chem. 2020, 295, 9998–10007. 10.1074/jbc.RA120.014064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Basri R. S.; Rahman R. N. Z. R. A.; Kamarudin N. H. A.; Ali M. S. M. Cyanobacterial aldehyde deformylating oxygenase: Structure, function, and potential in biofuels production. Int. J. Biol. Macromol. 2020, 164, 3155–3162. 10.1016/j.ijbiomac.2020.08.162. [DOI] [PubMed] [Google Scholar]
  5. Spector S.; Flynn J. M.; Tidor B.; Baker T. A.; Sauer R. T. Expression of N-formylated proteins in Escherichia coli. Protein Expr. Purif. 2003, 32, 317–322. 10.1016/j.pep.2003.08.004. [DOI] [PubMed] [Google Scholar]
  6. Vatsis K. P.; Coon M. J. Oxidative aldehyde deformylation catalyzed by NADPH-cytochrome P450 reductase and the flavoprotein domain of neuronal nitric oxide synthase. Biochem. Biophys. Res. Commun. 2005, 337, 1107–1111. 10.1016/j.bbrc.2005.09.167. [DOI] [PubMed] [Google Scholar]
  7. Shaik S.; Cohen S.; Wang Y.; Chen H.; Kumar D.; Thiel W. P450 Enzymes: Their Structure, Reactivity, and Selectivity—Modeled by QM/MM Calculations. Chem. Rev. 2010, 110, 949–1017. 10.1021/cr900121s. [DOI] [PubMed] [Google Scholar]
  8. Tang M.-C.; Zou Y.; Watanabe K.; Walsh C. T.; Tang Y. Oxidative Cyclization in Natural Product Biosynthesis. Chem. Rev. 2017, 117, 5226–5333. 10.1021/acs.chemrev.6b00478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Guengerich F. P.; Yoshimoto F. K. Formation and Cleavage of C–C Bonds by Enzymatic Oxidation–Reduction Reactions. Chem. Rev. 2018, 118, 6573–6655. 10.1021/acs.chemrev.8b00031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Huang X.; Groves J. T. Oxygen Activation and Radical Transformations in Heme Proteins and Metalloporphyrins. Chem. Rev. 2018, 118, 2491–2553. 10.1021/acs.chemrev.7b00373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Adam S. M.; Wijeratne G. B.; Rogler P. J.; Diaz D. E.; Quist D. A.; Liu J. J.; Karlin K. D. Synthetic Fe/Cu Complexes: Toward Understanding Heme-Copper Oxidase Structure and Function. Chem. Rev. 2018, 118, 10840–11022. 10.1021/acs.chemrev.8b00074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cramer C. J.; Tolman W. B. Mononuclear Cu-O2 Complexes: Geometries, Spectroscopic Properties, Electronic Structures, and Reactivity. Acc. Chem. Res. 2007, 40, 601–608. 10.1021/ar700008c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cho J.; Sarangi R.; Nam W. Mononuclear Metal-O2 Complexes Bearing Macrocyclic N-Tetramethylated Cyclam Ligands. Acc. Chem. Res. 2012, 45, 1321–1330. 10.1021/ar3000019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bagha U. K.; Satpathy J. K.; Mukherjee G.; Sastri C. V.; de Visser S. P. A comprehensive insight into aldehyde deformylation: mechanistic implications from biology and chemistry. Org. Biomol. Chem. 2021, 19, 1879–1899. 10.1039/D0OB02204G. [DOI] [PubMed] [Google Scholar]
  15. Elwell C. E.; Gagnon N. L.; Neisen B. D.; Dhar D.; Spaeth A. D.; Yee G. M.; Tolman W. B. Copper-Oxygen Complexes Revisited: Structures, Spectroscopy, and Reactivity. Chem. Rev. 2017, 117, 2059–2107. 10.1021/acs.chemrev.6b00636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Fukuzumi S.; Cho K. B.; Lee Y. M.; Hong S.; Nam W. Mechanistic dichotomies in redox reactions of mononuclear metal-oxygen intermediates. Chem. Soc. Rev. 2020, 49, 8988–9027. 10.1039/D0CS01251C. [DOI] [PubMed] [Google Scholar]
  17. Pan H.-R.; Chen H.-J.; Wu Z.-H.; Ge P.; Ye S.; Lee G.-H.; Hsu H.-F. Structural and Spectroscopic Evidence for a Side-on Fe(III)–Superoxo Complex Featuring Discrete O–O Bond Distances. JACS Au 2021, 1, 1389–1398. 10.1021/jacsau.1c00184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Pirovano P.; Magherusan A. M.; McGlynn C.; Ure A.; Lynes A.; McDonald A. R. Nucleophilic Reactivity of a Copper(II)–Superoxide Complex. Angew. Chem., Int. Ed. 2014, 53, 5946–5950. 10.1002/anie.201311152. [DOI] [PubMed] [Google Scholar]
  19. Winslow C.; Lee H. B.; Field M. J.; Teat S. J.; Rittle J. Structure and Reactivity of a High-Spin, Nonheme Iron(III)- Superoxo Complex Supported by Phosphinimide Ligands. J. Am. Chem. Soc. 2021, 143, 13686–13693. 10.1021/jacs.1c05276. [DOI] [PubMed] [Google Scholar]
  20. Donoghue P. J.; Gupta A. K.; Boyce D. W.; Cramer C. J.; Tolman W. B. An Anionic, Tetragonal Copper(II) Superoxide Complex. J. Am. Chem. Soc. 2010, 132, 15869–15871. 10.1021/ja106244k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Shokri A.; Que L. Jr. Conversion of Aldehyde to Alkane by a Peroxoiron(III) Complex: A Functional Model for the Cyanobacterial Aldehyde-Deformylating Oxygenase. J. Am. Chem. Soc. 2015, 137, 7686–7691. 10.1021/jacs.5b01053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Bailey W. D.; Gagnon N. L.; Elwell C. E.; Cramblitt A. C.; Bouchey C. J.; Tolman W. B. Revisiting the Synthesis and Nucleophilic Reactivity of an Anionic Copper Superoxide Complex. Inorg. Chem. 2019, 58, 4706–4711. 10.1021/acs.inorgchem.9b00090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Cho J.; Jeon S.; Wilson S. A.; Liu L. V.; Kang E. A.; Braymer J. J.; Lim M. H.; Hedman B.; Hodgson K. O.; Valentine J. S.; Solomon E. I.; Nam W. Structure and reactivity of a mononuclear non-haem iron(III)-peroxo complex. Nature 2011, 478, 502–505. 10.1038/nature10535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Colmer H. E.; Howcroft A. W.; Jackson T. A. Formation, Characterization, and O–O Bond Activation of a Peroxomanganese(III) Complex Supported by a Cross-Clamped Cyclam Ligand. Inorg. Chem. 2016, 55, 2055–2069. 10.1021/acs.inorgchem.5b02398. [DOI] [PubMed] [Google Scholar]
  25. Geiger R. A.; Wijeratne G. B.; Day V. W.; Jackson T. A. Steric and Electronic Influences on the Structures of Peroxomanganese(III) Complexes Supported by Tetradentate Ligands. Eur. J. Inorg. Chem. 2012, 2012, 1598–1608. 10.1002/ejic.201101391. [DOI] [Google Scholar]
  26. Geiger R. A.; Chattopadhyay S.; Day V. W.; Jackson T. A. Nucleophilic reactivity of a series of peroxomanganese(III) complexes supported by tetradentate aminopyridyl ligands. Dalton Trans. 2011, 40, 1707–1715. 10.1039/c0dt01570a. [DOI] [PubMed] [Google Scholar]
  27. Colmer H. E.; Geiger R. A.; Leto D. F.; Wijeratne G. B.; Day V. W.; Jackson T. A. Geometric and electronic structure of a peroxomanganese(iii) complex supported by a scorpionate ligand. Dalton Trans. 2014, 43, 17949–17963. 10.1039/C4DT02483D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Cho J.; Sarangi R.; Annaraj J.; Kim S. Y.; Kubo M.; Ogura T.; Solomon E. I.; Nam W. Geometric and electronic structure and reactivity of a mononuclear ″side-on″ nickel(III)-peroxo complex. Nat. Chem. 2009, 1, 568–572. 10.1038/nchem.366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Pegis M. L.; Martin D. J.; Wise C. F.; Brezny A. C.; Johnson S. I.; Johnson L. E.; Kumar N.; Raugei S.; Mayer J. M. Mechanism of Catalytic O2 Reduction by Iron Tetraphenylporphyrin. J. Am. Chem. Soc. 2019, 141, 8315–8326. 10.1021/jacs.9b02640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chen H.; Cho K. B.; Lai W.; Nam W.; Shaik S. Dioxygen Activation by a Non-Heme Iron(II) Complex: Theoretical Study toward Understanding Ferric-Superoxo Complexes. J. Chem. Theory. Comput. 2012, 8, 915–926. 10.1021/ct300015y. [DOI] [PubMed] [Google Scholar]
  31. Annaraj J.; Cho J.; Lee Y. M.; Kim S. Y.; Latifi R.; de Visser S. P.; Nam W. Structural characterization and remarkable axial ligand effect on the nucleophilic reactivity of a nonheme manganese(III)-peroxo complex. Angew. Chem., Int. Ed. 2009, 48, 4150–4153. 10.1002/anie.200900118. [DOI] [PubMed] [Google Scholar]
  32. Panda C.; Chandra A.; Corona T.; Andris E.; Pandey B.; Garai S.; Lindenmaier N.; Künstner S.; Farquhar E. R.; Roithová J.; Rajaraman G.; Driess M.; Ray K. Nucleophilic versus Electrophilic Reactivity of Bioinspired Superoxido Nickel(II) Complexes. Angew. Chem., Int. Ed. 2018, 57, 14883–14887. 10.1002/anie.201808085. [DOI] [PubMed] [Google Scholar]
  33. Barman P.; Upadhyay P.; Faponle A. S.; Kumar J.; Nag S. S.; Kumar D.; Sastri C. V.; de Visser S. P. Deformylation Reaction by a Nonheme Manganese(III)-Peroxo Complex via Initial Hydrogen-Atom Abstraction. Angew. Chem., Int. Ed. 2016, 55, 11091–11095. 10.1002/anie.201604412. [DOI] [PubMed] [Google Scholar]
  34. Schaefer A. W.; Roveda A. C.; Jose A.; Solomon E. I. Geometric and Electronic Structure Contributions to O–O Cleavage and the Resultant Intermediate Generated in Heme-Copper Oxidases. J. Am. Chem. Soc. 2019, 141, 10068–10081. 10.1021/jacs.9b04271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Schaefer A. W.; Kieber-Emmons M. T.; Adam S. M.; Karlin K. D.; Solomon E. I. Phenol-Induced O–O Bond Cleavage in a Low-Spin Heme–Peroxo–Copper Complex: Implications for O2 Reduction in Heme–Copper Oxidases. J. Am. Chem. Soc. 2017, 139, 7958–7973. 10.1021/jacs.7b03292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Kim H.; Rogler P. J.; Sharma S. K.; Schaefer A. W.; Solomon E. I.; Karlin K. D. Heme-FeIII Superoxide, Peroxide and Hydroperoxide Thermodynamic Relationships: FeIII-O2•– Complex H-Atom Abstraction Reactivity. J. Am. Chem. Soc. 2020, 142, 3104–3116. 10.1021/jacs.9b12571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Geiger R. A.; Chattopadhyay S.; Day V. W.; Jackson T. A. A Series of Peroxomanganese(III) Complexes Supported by Tetradentate Aminopyridyl Ligands: Detailed Spectroscopic and Computational Studies. J. Am. Chem. Soc. 2010, 132, 2821–2831. 10.1021/ja910235g. [DOI] [PubMed] [Google Scholar]
  38. Krishnan V. M.; Shopov D. Y.; Bouchey C. J.; Bailey W. D.; Parveen R.; Vlaisavljevich B.; Tolman W. B. Structural Characterization of the [CuOR](2+) Core. J. Am. Chem. Soc. 2021, 143, 3295–3299. 10.1021/jacs.0c13470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Magherusan A. M.; Kal S.; Nelis D. N.; Doyle L. M.; Farquhar E. R.; Que L. Jr.; McDonald A. R. A Mn(II) Mn(III) -Peroxide Complex Capable of Aldehyde Deformylation. Angew. Chem., Int. Ed. 2019, 58, 5718–5722. 10.1002/anie.201900717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Lin Y. H.; Kutin Y.; van Gastel M.; Bill E.; Schnegg A.; Ye S.; Lee W. Z. A Manganese(IV)-Hydroperoxo Intermediate Generated by Protonation of the Corresponding Manganese(III)-Superoxo Complex. J. Am. Chem. Soc. 2020, 142, 10255–10260. 10.1021/jacs.0c02756. [DOI] [PubMed] [Google Scholar]
  41. Cantú Reinhard F. G.; Barman P.; Mukherjee G.; Kumar J.; Kumar D.; Kumar D.; Sastri C. V.; de Visser S. P. Keto-Enol Tautomerization Triggers an Electrophilic Aldehyde Deformylation Reaction by a Nonheme Manganese(III)-Peroxo Complex. J. Am. Chem. Soc. 2017, 139, 18328–18338. 10.1021/jacs.7b10033. [DOI] [PubMed] [Google Scholar]
  42. Zhu W.; Jang S.; Xiong J.; Ezhov R.; Li X. X.; Kim T.; Seo M. S.; Lee Y. M.; Pushkar Y.; Sarangi R.; Guo Y.; Nam W. A Mononuclear Non-heme Iron(III)-Peroxo Complex with an Unprecedented High O–O Stretch and Electrophilic Reactivity. J. Am. Chem. Soc. 2021, 143, 15556–15561. 10.1021/jacs.1c03358. [DOI] [PubMed] [Google Scholar]
  43. Bhadra M.; Transue W. J.; Lim H.; Cowley R. E.; Lee J. Y. C.; Siegler M. A.; Josephs P.; Henkel G.; Lerch M.; Schindler S.; Neuba A.; Hodgson K. O.; Hedman B.; Solomon E. I.; Karlin K. D. A Thioether-Ligated Cupric Superoxide Model with Hydrogen Atom Abstraction Reactivity. J. Am. Chem. Soc. 2021, 143, 3707–3713. 10.1021/jacs.1c00260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Narulkar D. D.; Ansari A.; Vardhaman A. K.; Harmalkar S. S.; Lingamallu G.; Dhavale V. M.; Sankaralingam M.; Das S.; Kumar P.; Dhuri S. N. A side-on Mn(iii)–peroxo supported by a non-heme pentadentate N3Py2 ligand: synthesis, characterization and reactivity studies. Dalton Trans. 2021, 50, 2824–2831. 10.1039/D0DT03706K. [DOI] [PubMed] [Google Scholar]
  45. Annaraj J.; Cho J.; Lee Y. M.; Kim S. Y.; Latifi R.; de Visser S. P.; Nam W. Structural characterization and remarkable axial ligand effect on the nucleophilic reactivity of a nonheme manganese(III)-peroxo complex. Angew. Chem., Int. Ed. 2009, 48, 4150–4153. 10.1002/anie.200900118. [DOI] [PubMed] [Google Scholar]
  46. Jo Y.; Annaraj J.; Seo M. S.; Lee Y. M.; Kim S. Y.; Cho J.; Nam W. Reactivity of a cobalt(III)-peroxo complex in oxidative nucleophilic reactions. J. Inorg. Biochem. 2008, 102, 2155–2159. 10.1016/j.jinorgbio.2008.08.008. [DOI] [PubMed] [Google Scholar]
  47. Bae S. H.; Li X. X.; Seo M. S.; Lee Y. M.; Fukuzumi S.; Nam W. Tunneling Controls the Reaction Pathway in the Deformylation of Aldehydes by a Nonheme Iron(III)-Hydroperoxo Complex: Hydrogen Atom Abstraction versus Nucleophilic Addition. J. Am. Chem. Soc. 2019, 141, 7675–7679. 10.1021/jacs.9b02272. [DOI] [PubMed] [Google Scholar]
  48. Zhang Q.; Bell-Taylor A.; Bronston F. M.; Gorden J. D.; Goldsmith C. R. Aldehyde Deformylation and Catalytic C–H Activation Resulting from a Shared Cobalt(II) Precursor. Inorg. Chem. 2017, 56, 773–782. 10.1021/acs.inorgchem.6b02127. [DOI] [PubMed] [Google Scholar]
  49. Rana S.; Haque R.; Santosh G.; Maiti D. Decarbonylative Halogenation by a Vanadium Complex. Inorg. Chem. 2013, 52, 2927–2932. 10.1021/ic302611a. [DOI] [PubMed] [Google Scholar]
  50. Rana S.; Pandey B.; Dey A.; Haque R.; Rajaraman G.; Maiti D. A Doubly Biomimetic Synthetic Transformation: CatalyticDecarbonylation and Halogenation at Room Temperature by Vanadium Pentoxide. ChemCatChem 2016, 8, 3367–3374. 10.1002/cctc.201600843. [DOI] [Google Scholar]
  51. Annaraj J.; Suh Y.; Seo M. S.; Kim S. O.; Nam W. Mononuclear Nonheme Ferric-Peroxo Complex in Aldehyde Deformylation. Chem. Commun. 2005, 4529–4531. 10.1039/b505562h. [DOI] [PubMed] [Google Scholar]
  52. Corcos A. R.; Villanueva O.; Walroth R. C.; Sharma S. K.; Bacsa J.; Lancaster K. M.; MacBeth C. E.; Berry J. F. Oxygen Activation by Co(II) and a Redox Non-Innocent Ligand: Spectroscopic Characterization of a Radical-Co(II)-Superoxide Complex with Divergent Catalytic Reactivity. J. Am. Chem. Soc. 2016, 138, 1796–1799. 10.1021/jacs.5b12643. [DOI] [PubMed] [Google Scholar]
  53. Courtade G.; Ciano L.; Paradisi A.; Lindley P. J.; Forsberg Z.; Sørlie M.; Wimmer R.; Davies G. J.; Eijsink V. G. H.; Walton P. H.; Aachmann F. L. Mechanistic basis of substrate-O2 coupling within a chitin-active lytic polysaccharide monooxygenase: An integrated NMR/EPR study. Proc. Natl. Acad. Sci. U.S.A. 2020, 117, 19178–19189. 10.1073/pnas.2004277117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Jose A.; Schaefer A. W.; Roveda A. C.; Transue W. J.; Choi S. K.; Ding Z.; Gennis R. B.; Solomon E. I. The three-spin intermediate at the O–O cleavage and proton-pumping junction in heme-Cu oxidases. Science 2021, 373, 1225–1229. 10.1126/science.abh3209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Trammell R.; Rajabimoghadam K.; Garcia-Bosch I. Copper-Promoted Functionalization of Organic Molecules: from Biologically Relevant Cu/O2 Model Systems to Organometallic Transformations. Chem. Rev. 2019, 119, 2954–3031. 10.1021/acs.chemrev.8b00368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Liu J. J.; Diaz D. E.; Quist D. A.; Karlin K. D. Copper(I)-Dioxygen Adducts and Copper Enzyme Mechanisms. Isr. J. Chem. 2016, 56, 738–755. 10.1002/ijch.201600025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Sen K.; Hackett J. C. Peroxo–Iron Mediated Deformylation in Sterol 14α-Demethylase Catalysis. J. Am. Chem. Soc. 2010, 132, 10293–10305. 10.1021/ja906192b. [DOI] [PubMed] [Google Scholar]
  58. Barman P.; Cantú Reinhard F. G.; Bagha U. K.; Kumar D.; Sastri C. V.; de Visser S. P. Hydrogen by Deuterium Substitution in an Aldehyde Tunes the Regioselectivity by a Nonheme Manganese(III)–Peroxo Complex. Angew. Chem., Int. Ed. 2019, 131, 10749–10753. 10.1002/ange.201905416. [DOI] [PubMed] [Google Scholar]
  59. Cho D.; Choi S.; Cho J.; Baik M.-H. Peroxocobalt(iii) species activates nitriles via a superoxocobalt(ii) diradical state. Dalton Trans. 2020, 49, 2819–2826. 10.1039/D0DT00042F. [DOI] [PubMed] [Google Scholar]
  60. Zhao R.; Guo J.; Zhang C.; Lu Y.; Dagnaw W. M.; Wang Z.-X. DFT Mechanistic Insight into the Dioxygenase-like Reactivity of a CoIII-peroxo Complex: O–O Bond Cleavage via a [1,3]-Sigmatropic Rearrangement-like Mechanism. Inorg. Chem. 2020, 59, 2051–2061. 10.1021/acs.inorgchem.9b03470. [DOI] [PubMed] [Google Scholar]
  61. Noh H.; Jeong D.; Ohta T.; Ogura T.; Valentine J. S.; Cho J. Distinct Reactivity of a Mononuclear Peroxocobalt(III) Species toward Activation of Nitriles. J. Am. Chem. Soc. 2017, 139, 10960–10963. 10.1021/jacs.7b04479. [DOI] [PubMed] [Google Scholar]
  62. Tomasi J.; Mennucci B.; Cammi R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
  63. For calculation of the dielectronic constant of the solvent mixture, see:; Moore W. The use of an approximate dielectric constant to blend solvent systems. J. Am. Pharm. Assoc., Sci. Ed. 1958, 47, 855. 10.1002/jps.3030471208. [DOI] [PubMed] [Google Scholar]
  64. Zhao Y.; Truhlar D. G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101 10.1063/1.2370993. [DOI] [PubMed] [Google Scholar]
  65. Benson S. W.The Foundations of Chemical Kinetics; Krieger: Malabar, FL, 1982. [Google Scholar]
  66. Yuan R.; Lin Z. Mechanistic Insight into the Gold Catalyzed Carboxylative Cyclization of Propargylamines. ACS Catal. 2015, 5, 2866–2872. 10.1021/acscatal.5b00252. [DOI] [Google Scholar]
  67. Zhou Q.; Li Y. The Real Role of NHeterocyclic Carbene in Reductive Functionalization of CO2: An Alternative Understanding from Density Functional Theory Study. J.Am. Chem. Soc. 2015, 137, 10182–10189. 10.1021/jacs.5b03651. [DOI] [PubMed] [Google Scholar]
  68. Jiang Y.-Y.; Liu T.-T.; Zhang R.-X.; Xu Z.-Y.; Sun X.; Bi S. Mechanism and Rate-Determining Factors of Amide Bond Formation through Acyl Transfer of Mixed Carboxylic–Carbamic Anhydrides: A Computational Study. J. Org. Chem. 2018, 83, 2676–2685. 10.1021/acs.joc.7b03107. [DOI] [PubMed] [Google Scholar]
  69. Schoenebeck F.; Houk K. N. Ligand-Controlled Regioselectivity in Palladium-Catalyzed Cross Coupling Reactions. J. Am. Chem. Soc. 2010, 132, 2496–2497. 10.1021/ja9077528. [DOI] [PubMed] [Google Scholar]
  70. 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]
  71. Zhao Y.; Truhlar D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. 10.1021/ar700111a. [DOI] [PubMed] [Google Scholar]
  72. Zhao Y.; Truhlar D. G. Benchmark Energetic Data in a Model System for Grubbs II Metathesis Catalysis and Their Use for the Development, Assessment, and Validation of Electronic Structure Methods. J. Chem. Theory Comput. 2009, 5, 324–333. 10.1021/ct800386d. [DOI] [PubMed] [Google Scholar]
  73. Kulkarni A. D.; Truhlar D. G. Performance of Density Functional Theory and Møller–Plesset Second-Order Perturbation Theory for Structural Parameters in Complexes of Ru. J. Chem. Theory Comput. 2011, 7, 2325–2332. 10.1021/ct200188n. [DOI] [PubMed] [Google Scholar]
  74. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H–Pu. J. Chem. Phys. 2010, 132, 154104–154119. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  75. Yu H. S.; He X.; Li S. L.; Truhlar D. G. MN15: A Kohn–Sham global-hybrid exchange–correlation density functional with broad accuracy for multi-reference and single-reference systems and noncovalent interactions. Chem. Sci. 2016, 7, 5032–5051. 10.1039/C6SC00705H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Feller D. The role of databases in support of computational chemistry calculations. J. Comput. Chem. 1996, 17, 1571–1586. . [DOI] [Google Scholar]
  77. Schuchardt K. L.; Didier B. T.; Elsethagen T.; Sun L.; Gurumoorthi V.; Chase J.; Li J.; Windus T. L. Basis Set Exchange: A Community Database for Computational Sciences. J. Chem. Inf. Model. 2007, 47, 1045–1052. 10.1021/ci600510j. [DOI] [PubMed] [Google Scholar]
  78. Weigend F.; Furche F.; Ahlrichs R. Gaussian basis sets of quadruple zeta valence quality for atoms H-Kr. J. Chem. Phys. 2003, 119, 12753–12762. 10.1063/1.1627293. [DOI] [Google Scholar]
  79. Weigend F.; Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
  80. Glendening E. D.; Badenhoop J. K.; Reed A. E.; Carpenter J. E.; Bohmann J. A.; Morales C. M.; Landis C. R.; Weinhold F.. NBO, 6.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, 2013.
  81. Wiberg K. B. Application of the pople-santry-segal CNDO method to the cyclopropylcarbinyl and cyclobutyl cation and to bicyclobutane. Tetrahedron 1968, 24, 1083–1096. 10.1016/0040-4020(68)88057-3. [DOI] [Google Scholar]
  82. Reed A. E.; Curtiss L. A.; Weinhold F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. 10.1021/cr00088a005. [DOI] [Google Scholar]
  83. Weinhold F. Natural bond orbital analysis: A critical overview of relationships to alternative bonding perspectives. J. Comput. Chem. 2012, 33, 2363–2379. 10.1002/jcc.23060. [DOI] [PubMed] [Google Scholar]
  84. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam N. J.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09, revision A.01; Gaussian, Inc.: Wallingford, CT, 2009. [Google Scholar]
  85. Schröder D.; Shaik S.; Schwarz H. Two-State Reactivity as a New Concept in Organometallic Chemistry. Acc. Chem. Res. 2000, 33, 139–145. 10.1021/ar990028j. [DOI] [PubMed] [Google Scholar]
  86. Son Y.; Kim K.; Kim S.; Tripodi G. L.; Pereverzev A.; Roithová J.; Cho J. Spectroscopic Evidence for a Cobalt-Bound Peroxyhemiacetal Intermediate. JACS Au 2021, 1, 1594–1600. 10.1021/jacsau.1c00166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Aboelella N. W.; Kryatov S. V.; Gherman B. F.; Brennessel W. W.; Young V. G.; Sarangi R.; Rybak-Akimova E. V.; Hodgson K. O.; Hedman B.; Solomon E. I.; Cramer C. J.; Tolman W. B. Dioxygen Activation at a Single Copper Site:Structure, Bonding, and Mechanism of Formation of 1:1 Cu–O2 Adducts. J. Am. Chem. Soc. 2004, 126, 16896–16911. 10.1021/ja045678j. [DOI] [PubMed] [Google Scholar]
  88. Henson M. J.; Mukherjee P.; Root D. E.; Stack T. D. P.; Solomon E. I. Spectroscopic and Electronic Structural Studies of the Cu(III)2 Bis-μ-oxo Core and Its Relation to the Side-On Peroxo-Bridged Dimer. J. Am. Chem. Soc. 1999, 121, 10332–10345. 10.1021/ja9918425. [DOI] [Google Scholar]
  89. Liu W. J.; Huang H. H.; Ouyang T.; Jiang L.; Zhong D. C.; Zhang W.; Lu T. B. A Copper (II) molecular catalyst for efficient and selective photochemical reduction of CO2 to CO in a water-containing system. Chem.—Eur. J. 2018, 24, 4503–4508. 10.1002/chem.201705566. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

au2c00014_si_001.pdf (1.6MB, pdf)

Articles from JACS Au are provided here courtesy of American Chemical Society

RESOURCES