Abstract
High-valent metal-hydroxide species have been implicated as key intermediates in hydroxylation chemistry catalyzed by heme monooxygenases such as the cytochrome P450s. However, in some classes of P450s, a bifurcation from the typical oxygen rebound pathway is observed, wherein the protonated FerV(OH)(porphyrin) (Cpd-Il) species carries out a net hydrogen atom transfer reaction to form alkene metabolites. In this work, we examine the hydrogen atom transfer (HAT) reactivity of FeIV(OH)(ttppc) (l), ttppc = 5,10,15-tris(2,4,6-triphenyl)-phenyl corrole, toward substituted phenol derivatives. The iron hydroxide complex 1 reacts with a series of para-substituted 2,6-di-ferf-butylphenol derivatives (4-X-2.6-DTBP; X = OMe, Me, Et, H, Ac), with second-order rate constants k2 = 3.6(1) —1.21(3) X 104M_ s_ and yielding linear Hammett and Marcus plot correlations. It is concluded that the rate-determining step for O-H cleavage occurs through a concerted HAT mechanism, based on mechanistic analyses that include a KIE = 2.9(1) and DFT calculations. Comparison of the HAT reactivity of 1 to the analogous Mn complex, MnIV(OH)(ttppc), where only the central metal ion is different, indicates a faster HAT reaction and a steeper Hammett slope for 1. The O—H bond dissociation energy (BDE) of the MnI(HO-H) complexes were estimated from a kinetic analysis to be 85 and 89 kcal mol−1 for Mn and Fe, respectively. These estimated BDEs are closely reproduced by DFT calculations and are discussed in the context of how they influence the overall H atom transfer reactivity.
Graphical Abstract
INTRODUCTION
The design and development of well-defined, observable synthetic models of highly reactive metal—oxygen intermediates has helped elucidate the mechanisms of oxidation reactions catalyzed by metalloenzymes.1‘2 For example, synthesis of biomimetic high-valent metal-oxo complexes and examination of their hydrogen atom transfer (HAT) reactivity has provided insights into the ground-state thermodynamics of related biological oxidations, as well as information for the design of synthetic catalysts.3–7 Heme enzymes are biological catalysts that perform a large variety of oxidation reactions in nature. Cytochrome P450 (CYP), in particular, is a heme monooxygenase that activates inert C—H bonds via a highvalent iron-oxo intermediate, Compound I (Cpd-I).8—13 The rate-determining step for this reaction is the HAT from the C— H substrate to Cpd-I to form Compound II (Cpd-II), an iron(lV)-hydroxide species (Scheme 1). The subsequent step, which is rapid, is the rebound of the •OH group to the incipient carbon radical to form the hydroxylated substrate. Indeed, many of the studies of HAT involving model systems have focused on the reactivity of metal-oxo complexes and the thermodynamic analysis of the O—H bond of the metal- hydroxide intermediate formed during the HAT reaction. However, CYP monooxygenases, and the related CYP peroxygenases, can perform desaturation and C—C bond cleaving reactions in addition to hydroxylation, where their pathways can bifurcate at the Cpd-II state between oxygen rebound and HAT mechanisms. Desaturation reactions result in the formation of toxic alkene metabolites (Scheme 1). Conversion of valproic acid to 4-ene valproic acid,14–16 ethyl carbamate to vinyl carbamate,17 testosterone to 17^-hydroxy- 4,6-androstadiene-3-one,18 lauric acid to 11-dodecenoic acid,19 and decarboxylation of fatty acids to terminal alkenes (specifically by OleTjE)20–25 are just some of the examples of these CYP-mediated desaturation reactions.
Scheme 1.
Divergent Oxidation Pathways for CYP Cpd-II, and Examples of Desaturation Reactions Mediated by CYPs
While substrate positioning in the active site binding pocket has been implicated to play a key role in selection of rebound hydroxylation versus HAT pathways for Cpd-II in CYPs,22‘26 the fundamental electronic and structural factors that ultimately control these pathways remain poorly understood. Ground state thermodynamic properties are often invoked as key factors in controlling HAT by high-valent metal-oxo complexes. Along these lines, the ground state thermodynamics related to the HAT reactivity of Cpd-II can be assessed from the BDE(O-H) of the expected ferric-aquo product (Fem(H2O)(porph)(cys)). Few such BDE values are available from experiment on heme enzymes, but a BDE(O-H) of 90 kcal mol−1 was recently reported in one case for the Fem(H2O) species in CYP158.27 In contrast, the Fem(H2O) form of the heme enzyme aromatic peroxygenase (APO) was characterized as having a much lower BDE of 81 kcal/mol, despite the similarity of this thiolate-ligated heme active site with that of Cys-bound CYP. The corresponding O—H bond strength in the ferric hydroxide form of horseradish peroxidase (HRP), which contains a neutral His axial ligand trans to the OH group, is 85 kcal/mol as derived from the FeIV(O)/FeIII(OH) redox potential.28 Such a wide range in similar porphyrin ferric O—H bond strengths indicates that there are likely subtle, as yet unidentified factors regarding the metal ion active site that can affect this value and ultimately influence HAT reactivity.
Given the challenges in characterizing these species in the native, enzymatic systems, we have set out to develop well- defined synthetic FeIV(OH)/FeIII(H2O) porphyrinoid complexes that should allow for an examination of their reactivity and related thermodynamics. Such compounds, in heme or nonheme systems, are extremely rare. We previously reported the synthesis of FeIV(OH)(ttppc) (ttppc = 5’10’15-tris(2’4’6- triphenyl)-phenyl corrole) (1) and its reactivity toward para-X substituted trityl radical derivatives in efforts to model the radical rebound step leading to C—O bond formation.29 Kinetic studies indicated that the radical rebound step was a concerted, charge neutral process. It should be noted that with triarylcorrole ligands, the electronic ground state of formally FeIV and MnIV corroles has been described as varying between the two canonical valence tautomers, Mm(X)(corrole+●) and MIV(X)(corrole), depending on the axial ligand X.30,31 Conclusive characterization of the ground electronic state of these compounds is challenging, and complex 1 and the analogous Mn(OH)(ttppc) will be described as metal(IV) hydroxide complexes throughout this work for simplicity. Since our original report, the complex FeIV(OHn)(TAML) (TAML = tetra-amido macrocyclic ligand) was recently reported,32 as well as a nonheme FeIII(OH) complex that demonstrated both radical rebound and hydrogen atom transfer reactivity.33
Herein, the H atom transfer reactivity of FeIV(OH)(ttppc) (1) with a series of phenol derivatives is described, providing a mechanistic analysis of HAT reactivity. The synthesis and structural characterization of Fem(H2O)(ttppc) (2), the corresponding HAT product, was also accomplished. Hammett and Marcus plot analyses, together with kinetic isotope effect (KIE) measurements, provide detailed insight into how the HAT reaction proceeds. A direct comparison with the reactivity of MnIV(OH)(ttppc), combined with computational analyses of the reaction trajectories of these two high-valent metal hydroxide complexes, enabled assessment of how the central metal ion (Mn vs Fe) tunes the H atom transfer reactivity. The experimental and computational studies on 1, along with work on the analogous Mn complex MnIV(OH)- (ttppc), provide novel insights into the thermodynamic properties that govern the reactivity of metal hydroxide species toward H atom donors.
RESULTS AND DISCUSSION
Synthesis and Structural Characterization of an FeIII(H2O) Complex
The synthesis of Fem(H2O)(ttppc) (2), the expected H atom abstraction product of FeIV(OH)- (ttppc) (1), was prepared by ligand exchange of the previously synthesized FeIII(OEt2)2(ttppc)29 with H2O, followed by fluorobenzene/n-pentane vapor diffusion to give single crystals for X-ray diffraction. The molecular structure of FeIII(H2O)- (ttppc) (2) is shown in Figure 1. The structure of 2 shows an Fe—O bond distance of 2.126(2) Å, which is much longer as compared to that of 1 (Fe—O = 1.857(3) Å, Table S1). The Fe—O bond distance for 2 is similar to other well-characterized mononuclear Fem(H2O) units in porphyrin (2.012(2) — 2.084(4) Å)34—36 and tetra-amido macrocyclic ligand scaffolds (TAML) (2.097(2)—2.1102(18) Å).37,38 The X-ray structure of 2 is the first example of an H2O-bound iron(III) corrole, which are typically isolated with NO, Et2O, or pyridine axial ligands.10—12 The frozen solution EPR spectrum of 2 shows an intense signal at g = 4.29 consistent with an intermediate spin (S = 3/2) Fem metal center, similar to Fem(OEt2)2(ttppc) in toluene (Figure S1)29 and the related Fem(OEt2)2(tpfc) (tpfc = 5,10,15-tris(pentafluorophenyl) corrole).39,40 An Evans method measurement of 2 gives μeff = 3.6(1) μB, (spin-only μeff = 3.87 μB for S = 3/2), further supporting the spin state assignment. It should be noted that we are not able to unambiguously determine the extent to which H2O remains coordinated for the EPR and Evans method measurements in toluene. For comparison, the analogous MnШ(H2O)(ttppc)41 has a much longer metal—oxygen bond distance of 2.2645(18) A, likely due to Jahn—Teller distortion for the high-spin MnШ ion42.
Figure 1.
Displacement ellipsoid plot (35% probability level) for FeIII(H2O)(ttppc) (2) at 110(2) K. H atoms (except for those attached to O1) are omitted for clarity.
Reaction of FeIV(OH)(ttppc) (1) with H Atom Transfer Reagents
Initial studies on the HAT reactivity of 1 began with C—H substrates. Reaction of 1 with 9,10-dihydroanthracene (DHA), a common H atom donor with a relatively weak C—H bond (BDE = 76.3 kcal mol−1),43 was carried out. Preliminary UV—visible spectroscopy measurements showed that 1 reacts with DHA in toluene at 23 °C, although the reaction is very slow (<5% decay to Fem(H2O)(ttppc) (2) over 5 h). In contrast, reaction of 1 with 2,4-di-ferf-butylphenol (2,4-DTBP) (toluene, 23 °C) results in rapid formation of 2, as observed by UV—vis spectroscopy. Product analysis by gas chromatography (GC-FID) led to detection of the bis(phenol) dimer product 3,3’,5,5’-tetra-ferf-butyl-(1,1’-biphenyl)-2,2’- diol in good yield (67(1) %), consistent with a 1:1 stoichiometry for oxidation of the phenol substrate by the iron complex 1.
Having demonstrated the proficiency of 1 in cleaving the phenolic O—H bond in 2,4-DTBP, we employed a series of para-substituted 2,6-di-tert-butylphenol derivatives (4-X-2,6-di- t-Bu-C6H2OH; X = OMe, tBu, Et, Me, H, Ac) as H atom transfer substrates. Variation of the para substituents allowed us to measure the influence of the electronic and thermodynamic properties of the phenol derivatives on the H atom transfer reactivity (Figure 2a). A direct comparison with the reactivity of MnIV(OH)(ttppc) can also be made, providing an opportunity to assess how the central metal ion (Mn vs Fe) affects the reactivity toward hydrogen atom donors. Reaction of 1 with a series of para-substituted phenols were carried out similarly to the reaction of 1 with 2,4-DTBP. A solution of 1 and excess 4-methoxy-2,6-di-tert-butylphenol (4-OMe-2,6- DTBP) (5.4 equiv) was analyzed by X -band electron paramagnetic resonance (EPR) spectroscopy (16 K) and gave EPR signals corresponding to Fem(H2O)(ttppc) (2) (g = 4.29) and phenoxyl radical (g = 2.0) (Figure S3a). An additional signal (g = 5.3) appears in this reaction mixture, which can also be observed in the EPR spectrum of Fem(H2O)(ttppc) (2) and excess 4-OMe-2,6-DTBP (possibly due to formation of an FeIII-phenol adduct) (Figure S3b). These data showed that 1 reacts with 4-OMe-2,6-DTBP in accordance to the scheme in Figure 2a. The reaction kinetics of 1 with 4-OMe-2,6-DTBP was monitored by stopped-flow UV—vis spectroscopy, showing isosbestic conversion of 1 to 2 within 1 s (Figure 2b). Pseudo-first-order conditions (>10-fold equiv substrate) were employed and led to single exponential curves that were fit to up to five half-lives. The resulting observed rate constant (kobs) varied linearly with phenol concentration, yielding a second-order rate constant (k2) of 1.21(3) X 104 M—1 s—1. A kinetic isotope effect (KIE) was determined by reaction of 1 with the ArOD isotopologue of 4- OMe-2,6-DTBP. A significant KIE of 2.9(1) was measured, indicating that the rate-determining step involves O—H cleavage.
Figure 2.
(a) Reaction of FeIV(OH)(ttppc) with para-X-substituted 2,6-di-ieri-butylphenol derivatives. (b) Time-resolved UV-vis spectral changes observed in the reaction of 1 (15 μM) with 4-OMe-2,6-DTBP (0.52 mM) in toluene at 23 °C. Inset: changes in absorbance vs time for the growth of 2 (575 nm) (green circles) with the best fit line (black). (c) Plot of pseudo-first-order rate constants (kobs) vs [4-OMe-2,6-DTBP] for the OH (black) and OD (blue) phenol isotopologues.
The second-order rate constants (k2) for the other para- substituted 2,6-DTBP derivatives were obtained in a similar fashion (Figures S4–S9) and are listed in Table 1 together with their thermodynamic parameters (Hammett constants c+p, redox potentials, and BDEs). It should be noted that Eox values listed in Table 1 were obtained in acetonitrile. However, we still obtain a good correlation between our measured k2 values in toluene and the Eox values from CH3CN. In each case, isosbestic conversion of 1 to the FeIII(H2O) product 2 was rapid and varied over two orders of magnitude (k2 = 3.6(l)-1.21(3) X 104M−1 s−1).
Table 1.
Second-Order Rate Constants for the Reaction Between FeIV(OH)(ttppc) (1) and para-X-Substituted Phenol Derivatives, Hammett Constants (σ+p), Redox Potentials (Eox), and BDEs of 4-X-2,6-DTBP
Comparison of the Reactivity of FeIV(OH)(ttppc) (1) and MnIV(OH)(ttppc) with Phenol Derivatives
Hammett analysis of the kinetic data for 1 (Figure 3a, red circles) showed a decrease in the rate constants with the more electron- deficient phenol, with a slope of p+ = −2.6(8). This slope is higher as compared to that of MnIV(OH)(ttppc) (p+ = −1.4(2)) (Figure 3a, blue circles), implying a relatively greater charge separation in the HAT transition state for the Fe versus the Mn complex. A Marcus plot analysis, which provides information on the involvement of electron transfer in the rate-determining step of the reaction, was also performed. A slope of approximately −0.5 is predicted for the plot of (RT/F) ln k2 versus Eox (where k2 is the second-order rate constant and Eox is the redox potential of the substrate being oxidized) for reactions with a rate-determining electron transfer step, provided that ΔG°ET/λ is close to zero.47 A slope of −0.19(6) was obtained for 1 (Figure 3b, red squares), close to the values obtained for HAT reactions of a Cu-superoxo complex (slope = −0.29)45 and MnIV(OH)(ttppc) (slope = −0.12) (Figure 3b), in which a concerted H atom transfer is invoked. Plotting the log k2 for 1 versus the bond dissociation energy (BDE) of the O-H bond of the phenol substrate (Figure 3c) also shows a good correlation, with a slope of −0.58(1). This slope is larger than that for MnIV(OH)(ttppc) (slope = −0.25(3)), suggesting that the FeIV(OH) complex is more sensitive to the changes in BDEs of the O-H bond being broken. However, it must be noted that the range of BDEs for the O-H substrates is small, and therefore, caution must be used in the interpretation of these data.
Figure 3.
Comparison of (a) Hammett, (b) Marcus, and (c) log k2 vi BDE plots for the reaction of FeIV(OH)(ttppc) (1) (red circles/ squares, in toluene) or MnIV(OH)(ttppc) (blue circles/squares, ir benzene) with 4-X-2,6-DTBP. Kinetic data for MnIV(OH)(ttppc) from ref 41.
This difference in slope implies that MnIV(OH)(ttppc) will exhibit larger rate constants than 1 when reacting with O-H substrates having sufficiently high BDEs (>ca. 85 kcal mol−1). A similar reversal in relative rate of HAT for a nonheme MnIV(O) versus an analogous FeIV(O) complex was recently described.55–57 Taken together, the data suggest a rate-limiting H atom transfer reaction for complex 1 with phenolic (O-H bond) substrates. It is reasonable to conclude that the HAT reaction proceeds in a concerted manner but with a partia transfer of charge, rather than stepwise PT/ET or ET/PT pathways.
Comparison of HAT Reactivity with Other Oxidants and Estimation of BDEs from the Kinetics Data
The rate constants for the oxidation of 4-X-2,6-DTBP (X = OMe, Me H) by FeIV(OH)(ttppc) (1) and MnIV(OH)(ttppc) can be compared to the previously reported rate constants for substituted phenol oxidation by various metal-oxo and metal hydroxo complexes, as well as a photolytically generated t BuO· radical (Table 2).48‘50‘58 In general, the rate constants for 1 are orders of magnitude larger than those for Mn- and Fe· oxo and hydroxo complexes. It should be noted that the rate constants in Table 2 are taken from reactions in differen solvents and temperatures, and thus close comparisons of the values are not warranted. However, in general, the rate constants for 1 are orders of magnitude larger than those for Mn- and Fe-oxo and hydroxo complexes. Although the high-valent FeIV(oxo) complexes in Table 2 may be among the weaker FeIV(O) oxidants, it is still interesting to note that 1, a terminal hydroxide complex, is a significantly more reactive oxidant for the electron-rich phenols.
Table 2.
Comparison of Second-Order Rate Constants (M−1 s−1) and Hammett Slopes (ρ+) for the Oxidation of Substituted Phenols, 4-X-2,6-DTBP (X = OMe, Me, H), by M(O(H)) Complexes and t-BuO•, and the Corresponding O–H BDEs of the Reduced Complex
oxidant | –OMe | –Me | –H | ρ+ | solvent | reduced complex | BDE (O–H) (kcal mol−1) |
---|---|---|---|---|---|---|---|
FeIV(OH)(ttppc) (1) | 1.21 × 104 | 2.0 × 103 | 83.0 | −2.6 | toluene, 23 °C | FeIII(HO–H)(ttppc) | 89a |
MnIV(OH)(ttppc)b | 510 | 150 | 36 | −1.4 | benzene, 23 °C | MnIII(HO–H)(ttppc) | 85a |
MnIII(OH)(dpaq)+c | 1.44 | 1.20 | 0.75 | −0.88 | CH3CN, 50 °C | MnII(HO–H)(dpaq)+ | 64d |
FeIV(O)(TMC)(NCCH3)+e | 11.4 | 0.18 | 0.04 | −3.20 | CH3CN, 25 °C | FeIII(O–H)(TMC)(NCCH3)+ | 78e |
FeIV(O)(TMC)(CF3CO2)e | 26.4 | 1.84 | 0.46 | −2.30 | CH3CN, 25 °C | FeIII(O–H)(TMC)(CF3COO) | 76.4e |
FeIV(O)(TMC)(N3)e | 29.8 | 1.84 | 0.68 | −1.50 | CH3CN, 25 °C | FeIII(O–H)(TMC)(N3) | 77.4e |
MnIV(OH)(salen)f | – | 412 | 37.3 | – | CH2Cl2, −70 °C | MnIII(HO–H)(salen) | – |
[RuVI(L)(O)2]2+g | 2.56 × 104 | 26 | 0.58 | −2.90 | CH3CN, 25 °C | [RuV(L)(O)(O–H)]2+ | 82.8g |
t-BuO•h | – | 1.2 × 108 | 4.3 × 107 | – | 1:2 C6H6/C8H18O2, 22 °C | t-BuO–H | 105i |
BDE values extrapolated from the correlation in Figure 4. BDE(O–H) refers to the bond dissociation energy of the reduced complex.
Rate constants from ref 41.
Rate constants calculated from kobs (s−1) in ref 48.
Ref 49.
Rate constants calculated from kobs (s−1) values in ref 50.
Rate constants calculated from kobs (s−1) values ref 51.
Ref 52.
ref 53.
ref 54.
An Evans—Polanyi plot (Figure 4) shows good correlation of the 4-Me-2,6-DTBP reaction rates of these complexes with the BDE(O—H) of the HAT reaction product. To calibrate this relationship, a line was fit to the plot of rate constants (log k2) of similar compounds whose BDE(O—H) values are determined experimentally or calculated by DFT methods. Based on this plot (Figure 4), BDEs of 85 and 89 kcal mol—1 are predicted for MnIII(HO—H)(ttppc) and FeIII(HO—H)- (ttppc) (2), respectively. A similar linear correlation was obtained when the log k2 versus BDE for 4-H-2,6-DTBP was plotted (Figure S10); however, a slightly lower BDE(O—H) value (~87 kcal mol—1) was predicted for FeIII(HO—H)(ttppc) (2) using this substrate.
Figure 4.
Plot of log k2 for the reaction of metal-oxo/hydroxo complexes and tBuO· with 4-Me-2,6-t-Bu-phenol (22—25 °C) vs BDE(O—H) of the O—H bond formed.
Computational Studies on the HAT Mechanism
Density functional theory (DFT) calculations were done using extensively tested and benchmarked methods, which have been shown to reproduce experimental BDEs within two kcal mol−1.59–61 We began by calculating the reaction mechanism of FeIV(OH)(ttppc) (1) and MnIV(OH)(ttppc) with the 4-H-2,6-DTBP substrate. Starting with the reactants, both complexes are charge neutral and were studied with the complete ttppc ligand included with odd (for Fe) or even (for Mn) multiplicity. The [FeIV(OH)(ttppc)] complex is in a triplet spin ground state with the quintet and singlet spin states of ΔG = 3.7 and 15.4 kcal mol−1 higher in energy (Table S8). With manganese as the central metal ion, the ground state is the quartet spin state, while the doublet and sextet spin states are ΔE + ZPE = 21.3 and 8.0 kcal mol−1 higher (Table S10). The group spin densities of 3[Fe(OH)(ttppc)] indicate approximately three unpaired electrons on the iron (pFeOH = 2.79) antiferromagnetically coupled to one unpaired electron on the ttppc ligand (ptppcc = −0.80) in an a2u type orbital. These spin densities are more consistent with the 3[Fem(OH)(ttppc+●)] valence tautomer. This configuration differs from cytochrome P450s, which, in almost all cases, is calculated to exhibit a 3[FeIV(OH)(heme)] as the ground state.22,62 We attempted to calculate the energy difference between the Fem(OH)(ttppc+●) and the FeIV(OH)(ttppc) structures by swapping the molecular orbitals to obtain a structure with triplet spin configuration Sx2_y22 n*xz1 n*yz1 c*z2° a1u2 a2u2, designated as 3[FeIV(OH)(ttppc)]. Unfortunately, however, during the SCF convergence the state relaxed back to the 3[Fem(OH)(ttppc+·)] state. The ground state calculations for the manganese complex gives spin densities consistent with a similar description, i.e., 4[Mnm(OH)(ttppc+·)]. The bond distances from the DFT optimized structures for 3[Fem(OH)- (ttppc+●)] (Figure S11, Table S2) and 4[Mnm(OH)(ttppc+·)] (Figure S12, Table S3), as well as the reaction products 4[Fem(H2O)(ttppc)] (Figure S13, Table S4) and 3[MnIII(H2O)(ttppc)] (Figure S14, Table S5), are all in good agreement with the experimentally obtained bond distances from the X-ray crystal structures of these complexes. While the DFT-optimized structures for 1 and the Mn analog suggest a metal(III)(corrole+●) structure, a definitive electronic structure assignment cannot be made from the available structural and spectroscopic data.
The hydrogen atom transfer from the para-H 2,6-DTBP to FeIV(OH)(ttppc) and MnIV(OH)(ttppc) complexes was investigated, and the transition state structures are given in Figure 5. The reaction is concerted with a single hydrogen atom abstraction to form water and phenoxyl radical. We find; hydrogen atom abstraction barriers of AE + ZPE* = 4.3 and 5.7 kcal mol−1 for the FeIVOH and MnIVOH complexes,’ respectively (Figure S15a, Tables S8 and S10). These barriers are low and hence predict a fast chemical reaction. The difference in HAT barrier height for the FeIV(OH) versus MnIV(OH) complexes implicates a rate enhancement of about a factor of 10 for iron with respect to manganese from; transition state theory at 298 K, which is reasonably close to the experimentally reported rate differences for this phenol l· (Table 2).
Figure 5.
UB3LYP/BS2 optimized transition state geometries in solvent for the hydrogen atom abstraction from para-H-2,6-DTBP by FeIV(OH)(ttppc) and MnIV(OH)(ttppc). Bond lengths are in angstroms, and energies contain zero-point energy and solvent corrections.
Geometrically, the transition states are relatively central with almost equal FeO-H versus H—OPh and MnO-H versus H’ OPh distances. In particular, short FeO-H/MnO-H distances ! of 1.158/1.183 A are found, while the H-O distance with the i phenol is 1.267/1.242 A. Consequently, the iron and manganese structures are very similar. The transition states are early and have an electronic structure close to the reactants,; with only about 20% charge-transfer (pSub = −0.18/−0.20 in i 3TSHAFe/4TSHAMn). These results match the Hammett and Marcus plots reported above that excluded a separation of proton and electron transfer through either ET/PT or PT/ET pathways. Indeed, the electron transfer is simultaneous with proton transfer and indicates an HAT mechanism for the reaction. These spin and charge distributions in the transition states match those calculated previously for desaturation i reactions by CYPs.26,63
To examine the contribution of the secondary coordination; sphere effects of the sterically hindered ligand ttppc (5,10,15- i tris(2,4,6-triphenyl)phenyl corrole) on the HAT reaction, we also ran transition state calculations with a hypothetical bare; corrole, in which the (2,4,6-triphenyl)phenyl substituents on the 5,10,15- positions were removed and replaced with hydrogen atoms. Based on the potential energy landscape for the HAT reaction (Tables S12 and S14), the barrier drops considerably by 3.6 kcal mol−1 for 3[Fe(OH)(corrole)] and by : 5.3 kcal mol−1 for 4[Mn(OH)(corrole)] (Figure S15b). The : second-coordination sphere effect of the bulky phenyl groups appears to increase the HAT barrier and facilitate the direct observation of these kinetically transient chemical steps.
Calculation of Bond Dissociation Energies (BDEs) from Density Functional Theory (DFT)
Bond dissociation; energies (BDEOH) were calculated from the following reaction of Mih(H2O)(l) → MIV(OH)(L) + H●+ BDEOH and taken as : the energy difference of the electron affinity (EA) of the : metal(lV)-hydroxide complex, the experimental ionization; energy of a hydrogen atom (IEH),64 and the acidity (AGacid) : of the metal(lIl)-aquo complex (Figure S16) in their; respective solvents (toluene for Fe and benzene for Mn).; Our calculated BDEOH values (with ZPE included) match the predicted values well, with BDEOH (MnIII(H2O)(ttppc)) = 83.7 kcal mol−1 (in benzene) and BDEOH (FeIII(H2O)(ttppc)) : = 88.5 kcal mol−1 (in toluene) (Tables S19 and S20). The : difference between these two values comes from a larger electron affinity of the iron(lV)-hydroxo complex (by 12.1 kcal : mol−1), which is partially canceled by a larger acidity (by 7.3 : kcal mol−1) of the iron(lIl)-aquo complex. This relative < reactivity for directly analogous Fe versus Mn complexes has been similarly observed in comparisons of FeIV(O) versus MnIV(O) nonheme complexes, where FeIV(O) species are much more reactive toward HAT substrates, due to their more positive redox potentials.65,66 Similarly, for porphyrin and porphyrin-type complexes, FerV(O)(porphyrin^+) species have < been shown to be much more reactive toward HAT reactions versus the analogous MnV(O) species in the same ligand < environment.3 In contrast, previous thermodynamic analysis of H atom transfer reactions of MnV(O) corrolazines revealed that the basicity of the MIV(O) (M = Mn, Cr) intermediate controls the H atom transfer reactivity for these reactions.67 In the case of the H atom transfer reactions of MIV(OH)(ttppc) (M = Fe, Mn) studied here, the < larger calculated redox potential of the FeIV(OH) complex seems to have a significant effect on the faster H atom transfer reactivity in comparison to MnIV(OH)(ttppc).
CONCLUSION
The hydrogen atom transfer reactivity of FeIV(OH)(ttppc) (l), a Cpd-II analog, was described. It is capable of the rapid < oxidation of phenol derivatives by H atom abstraction of the O-H bond and is more reactive than the other high-valent metal-hydroxo oxidants and the weakly oxidizing FeIV(O)- (TMC) complexes given in Table 2 as tested against the same substrates. It also reacts more rapidly than the analogous MnIV(OH)(ttppc) complex, providing a direct comparison of the oxidizing power of Fe(OH) versus Mn(OH) species at the same oxidation level and in identical ligand environments. The final HAT product, FeIII(H2O)(ttppc), was also definitively characterized by XRD.
The kinetic analyses and computational studies provide strong support for a mechanism of phenol oxidation by 1, and its Mn analog, that involves a concerted, rate-determining HAT step. The calculations suggest a modest amount of charge transfer in the transition state, and this finding is in line with the weak Hammett and Marcus plot slopes. A plot of BDE(O- < H) values for M(O(H)) complexes versus log for the oxidation of 4-Me-2,6-di-fert-butyl phenol is linear and provides a means to estimate the BDE(O-H) values for the FeIII(H2O) and MnIII(H2O) ttppc complexes from kinetics data. These estimated values are BDE(Fem(HO-H)) = ~89 kcal mol−1 and BDE(Mnm(HO-H)) = ~85 kcal mol−1, and DFT calculations fully support these values. The estimated BDE(O-H) for Fem(H2O)(ttppc) is strikingly similar to the BDe(o-h) recently reported for the Fem(H2O) form of CYP158 (90 kcal mol−1).27 However, it should be noted that the BDE(O-H) calculations here are ±2 kcal mol−1 at best from the error in DFT calculations and kinetic measurements. Taken together, these data suggest that 1 is a viable model for protonated Cpd-II in CYP from a structural and thermodynamic perspective, although the electronic structure of 1 may be closer to Fem(oH)(corrole+·), as opposed to FeIV(OH)- (corrole).
The BDE(O-H) values for the FeIII(H2O) and MnIII(H2O) complexes align with the relative HAT reactivity exhibited by 1 and the Mn analog and suggest that it is the overall thermodynamic driving force for these reactions, determined by the BDE(O-H) values, that is a key factor in controlling the reaction rates of phenol oxidation. However, examination of the log fc2 versus the phenol BDE(O-H) plots for 1 and the Mn analog show that 1 is much more reactive toward the electron-rich phenols but is only modestly more reactive toward the stronger O-H bond substrates. In fact, the relative reactivities for Fe versus Mn should invert with substrates that have O-H bonds much stronger than the 83 kcal mol−1 reported for the para-H derivative. This hypothetical switch in relative reactivities, as expressed in the different slopes of log fc2 versus BDE(OH) for Fe and Mn, suggest that the BDE(O-H) of the metal-bound water ligand, and hence the thermodynamics of the overall HAT reaction, are not the only predictors of efficient H atom transfer reactivity for a high-valent metal- oxo/hydroxo species.
The FeIV(OH) complex 1 is capable of H atom abstraction, as shown with the phenols in this study, and is also capable of hydroxyl radical transfer to carbon radicals to give hydroxy- lated product as seen in the model “rebound” reactions described previously.29 These findings may give some insight regarding the selection of iron, as opposed to other redox- active, earth-abundant metals such as manganese, as the metal of choice in heme monoooxygenases. The iron center appears to be an efficient and versatile oxidant in a variety of oxidation and protonation states.
EXPERIMENTAL SECTION
General Materials and Methods
All chemicals were purchased from commercial sources and used without further purification unless otherwise stated. Reactions involving inert atmosphere were performed under Ar using standard Schlenk techniques or in an N2- filled drybox. Toluene, dichloromethane, acetonitrile, and diethyl ether were purified via a Pure-Solv solvent purification system from Innovative Technologies, Inc. Benzene, ethyl acetate, and fluoroben- zene were obtained from commercial sources. Deuterated solvents for NMR were purchased from Cambridge Isotope Laboratories, Inc. (Tewksbury, MA). The 4-X-2,6-di-tert-butyl-phenol derivatives were recrystallized twice in ethanol (X = OMe, Me), acetonitrile (X = Et, Ac), or n-pentane (X = H) and dried under vacuum. The monodeuterated 4-OMe-2,6-di-tert-butylphenol was synthesized using a previously published procedure.68 FeIV(OH)(ttppc) was synthesized and purified as previously reported.29 The dark-red complex FeIII(OEt2)2(ttppc) was synthesized by reduction of FeIV(Cl)(ttppc), following a previously published report.29
Instrumentation
Kinetics and other UV-vis measurements were performed on a Hewlett-Packard Agilent 8453 diode-array spectrophotometer with a 3.5 mL quartz cuvette (path length = 1 cm). For reactions with total reaction time of <10 s, stopped-flow experiments were carried out using a HiTech SHU-61SX2 (TgK Scientific Ltd.) stopped-flow spectrophotometer with a xenon light source and Kinetic Studio software. LH NMR spectra were recorded on a Bruker Avance 400 MHz NMR spectrometer at 298 K and referenced against residual solvent proton signals. Electron paramagnetic resonance (EPR) spectra were recorded with a Bruker EMX spectrometer equipped with a Bruker ER 041 X G microwave bridge and a continuous-flow liquid helium cryostat (ESR900) coupled to an Oxford Instruments TC503 temperature controller for low temperature data collection. Laser desorption ionization mass spectrometry (LDI-MS) was conducted on a Bruker Autoflex III MALDI ToF/ToF instrument (Billerica, MA) equipped with a nitrogen laser at 335 nm using an MTP 384 ground steel target plate. The instrument was calibrated using peptide standards of known molecular weights. Gas chromatography (GC-FID) was carried out on an Agilent 6890N gas chromatograph fitted with a DB-5 5% phenylmethyl siloxane capillary column (30 m X 0.32 mm X 0.25 ^m) and equipped with a flame- ionization detector. Elemental analysis was performed at Atlantic Microlab, Inc., Norcross, GA. Cyclic voltammetry was performed on an EG&G Princeton Applied Research potentiostat/galvanostat model 263A with a three-electrode system consisting of a glassy carbon working electrode, a Ag/AgNO3 nonaqueous reference electrode (0.01 M AgNO3 with 0.1 M B^NPFg in CH3CN), and a platinum wire counter electrode. Potentials were referenced using an external ferrocene standard. Scans were run under an Ar or N2 atmosphere at 23 °C using B^NPF6 (0.1 M) as the supporting electrolyte.
Fem(H2O)(Uppc) (2)
Distilled H2O (0.5 mL) was added to a solution of FeIII(OEt2)2(ttppc) in fluorobenzene (2 mL). Vapor diffusion of pentane to this biphasic solution led to the slow formation of dark red blocks (55 mg, 56% yield) after 1 week. UV-vis (C6H6) ¿mal, nm (e X 104 M−1 cm−1): 344 (3.29), 429 (7.27), 578 (1.35), 764 (0.24). 1H NMR (400 MHz, benzene-d6): 35.1 (s, br), 12.4 (s, br), 12.3 (s, br), 10.8 (s, br), 8.3 (s, br), 6.0 (s, br), 4.98 (s, br), 4.66 (s, br), −19.0 (s, br), −35.1 (s, br), −61.6 (s, br), −69.0 (s, br) S (ppm). IR (KBr): 3580 (w, O-H), 3512 (vw, O-h), 3055 (m), 3028 (m), 1593 (s), 1556 (w), 1491 (s), 1443 (m), 1424 (w), 1394 (w), 1362 (w), 1332 (m), 1296 (m), 1214 (s), 1180 (w), 1154 (m), 1075 (m), 1045 (sh), 1020 (s), 984 (s), 916 (w), 885 (s), 851 (w), 821 (w), 794 (m), 750 (vs), 695 (vs), 637 (m), 609 (w), 574 (m), 537 (m), 499 (m), 443 (w). LDI-MS (m/z): isotopic cluster centered at 1264.6 ([M - H2O]+). Anal. calcd for C91H59N4FeH2O-2C6H5F: C, 81.33; H, 5.07; N, 3.91. Found: C, 81.53; H, 4.80; N, 4.05. μeff = 3.6(1) μB (Evan’s method). EPR: g = 4.29 (16 K).
Reaction of FeIV(OH)(ttppc) with 2,4-di-tert-Butylphenol: Product Analysis
Under an inert atmosphere, a solution of FeIV(OH)(ttppc) in benzene (2.1 mM, 500 μL) was combined with 2,4-di-tert-butylphenol (0.132 mmol, 126 equiv), and 3.4 mg of eicosane (6 mM) as an internal standard. The reaction mixture was stirred and monitored by UV-vis spectroscopy, which showed complete, instantaneous conversion of FeIV(OH) (ttppc) (1) to FeIII(H2O)(ttppc) (2). An aliquot of the reaction mixture was injected directly onto the GC for analysis. The phenol oxidation product 3,3/,5,5/-tetra-£er£-butyl-(1,1/-biphenyl)-2,2’-diol was identified by GC in comparison with an authentic sample and quantified by integration of the peak and comparison with a calibration curve constructed with eicosane. The analysis was performed in triplicate. The average yield was equal to 67(1)% based on the reaction stoichiometry of 0.5 equiv of phenol oxidation product per 1.0 equiv of FeIV(OH)(ttppc).
Kinetics
To a solution of FeIV(OH)(ttppc) (15 μM) in toluene, varying amounts of 4-X-2,6-di-tert-butylphenol (X = OMe, Me, Et, H, Ac) (0.1–0.6 mM) in toluene were added to start the reaction. The spectral changes showed isosbestic conversion of FeIV(OH)(ttppc) to FeIII(H2O)(ttppc). The pseudo-first-order rate constants, fcobs, for these reactions were obtained through nonlinear least-squares fitting of the plots of absorbance at 575 nm (AbsJ versus time (t) over 5 half-lives according to the equation Abst = Absf + (Abs0 - Absf) exp(−fcobst), where Abs0 and Absf are the initial and final absorbance, respectively. Second-order rate constants (kj) were obtained from the slope of the best-fit line from a plot of kobs versus phenol concentration.
X-ray Crystallography
All reflection intensities for 2 were measured at 110(2) K using a SuperNova diffractometer (equipped with an Atlas detector) with Cu Ka radiation (λ = 1.54178 A) under the program CrysAlisPro (version 1.171.39.29c, Rigaku OD, 2017). The same program was used to refine the cell dimensions and for data reduction. The structure was solved with the program SHELXS-2014/7 and refined on F2 with SHELXL-2014/7.69 The temperature of the data collection was controlled using the system Cryojet (manufactured by Oxford Instruments). An analytical numeric absorption correction method was used involving a multifaceted crystal model based on expressions derived elsewhere.70 The H atoms were placed at calculated positions using the instructions AFIX 43 with isotropic displacement parameters having values 1.2 Ueq of the attached C atoms. The crystal lattice contains disordered and/or partially occupied lattice solvent molecules (C6H5F and C5H12). Their contributions were removed from the final refinement using the SQUEEZE program. The crystallographic data for FeIII(H2O)(ttppc) are summarized in Table S6.
Computational Modeling
Calculations on the complexes FeIV(OH)(ttppc) and MnIV(OH)(ttppc) and their reactions with para-H-2,6-DTBP included all atoms for both metal complex and phenol. Geometry optimizations, constraint geometry scans, and vibrational frequencies were calculated with DFT in the Gaussian-09 software package.71 We utilized the unrestricted B3LYP hybrid density functional method72,73 in combination with an LANL2DZ basis set74,75 on iron/manganese (with core potential) and 6–31G on the rest of the atoms (H, C, N, O), e.g., basis set BS1. Single points using the LACV3P+ (with core potential) on Fe/Mn and 6–311+G* on the rest of the atoms were done to correct the energies, e.g., basis set BS2. The continuum polarized conductor model with a dielectric constant mimicking toluene was applied to the system during the geometry optimizations and frequencies.76 Single point energy calculations were done in benzene for the BDE determination of MnIII(H2O)(ttppc) to match the experimental conditions (Table S20). We initially optimized a reactant complex of substrate and metal(IV)-hydroxo species in close proximity and followed the pathway for H transfer through a constrained geometry scan. Subsequently, the maximum in energy of these scans was subjected to a full transition state search and the obtained transition state was characterized with a frequency calculation that gave an imaginary frequency for the correct mode. These transition states were shown to be connected to reactants and products through an intrinsic reaction coordinate (IRC) scan. As in some cases, the transition states energies were lower than those of the original reactant complexes, so the latter were reoptimized from the final points of the IRCs, which gave slightly more stable isomers. These methods were used in previous studies and were shown to predict the correct spin-state orderings and barrier heights of reaction mechanisms.26,77,78 The calculations of the BDE(O-H) values are described in the Supporting Information (Tables S19 and S20).
Supplementary Material
ACKNOWLEDGMENTS
The authors would like to thank the NIH (GM101153) (D.P.G.) for the financial support provided for this research. M.Q.E.M. thanks the Government of Malaysia for a studentship. J.P.T.Z. thanks JHU for the Glen E. Meyer ‘39 Fellowship.
Appendix
We calculated reaction energies for the following processes:
-
1
[MIII(H2O)] [MIV(OH)] + H•
ΔHreaction 1 = BDEOH
-
2
[MIV(OH)] + e‒ [MIII(OH)]‒
ΔHreaction 2 = EA
-
3
[MIII(H2O)] [MIII(OH)]‒ + H+
ΔHreaction 3 = ΔGacid
-
4
H H+ + e‒
ΔHreaction 4 = IEH
The BDEOH was calculated from: BDEOH = ‒EA –IEH + ΔGacid
Footnotes
The authors declare no competing financial interest.
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.9b02923.
EPR spectra, kinetics data, single-crystal X-ray crystallography analysis, DFT optimized structures, calculation of BDEs, computational tables with group spin densities, charges, and absolute and relative energies of all structures, as well as Cartesian coordinates of optimized geometries (PDF)
Accession Codes
CCDC 1836730 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
REFERENCES
- (1).Zaragoza JPT; Goldberg DP Dioxygen Binding and Activation Mediated by Transition Metal Porphyrinoid Complexes. Dioxygen-dependent Heme Enzymes 2018, 1–36. [Google Scholar]
- (2).Sahu S; Goldberg DP Activation of Dioxygen by Iron and Manganese Complexes: A Heme and Nonheme Perspective. J. Am. Chem. Soe. 2016, 138, 11410–11428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Baglia RA; Zaragoza JPT; Goldberg DP Biomimetic Reactivity of Oxygen-Derived Manganese and Iron Porphyrinoid Complexes. Chem. Rev. 2017, 117, 13320–13352. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Sacramento JJD; Goldberg DP Factors Affecting Hydrogen Atom Transfer Reactivity of Metal-Oxo Porphyrinoid Complexes. Aee. Chem. Res. 2018, 51, 2641–2652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Engelmann X; Monte-Perez I; Ray K Oxidation Reactions with Bioinspired Mononuclear Non-Heme Metal-Oxo Complexes. Angew. Chem., Int. Ed. 2016, 55, 7632–7649. [DOI] [PubMed] [Google Scholar]
- (6).Nam W; Lee Y-M; Fukuzumi S Tuning Reactivity and Mechanism in Oxidation Reactions by Mononuclear Nonheme Iron(IV)-Oxo Complexes. Aee. Chem. Res. 2014, 47, 1146–1154. [DOI] [PubMed] [Google Scholar]
- (7).McDonald AR; Que L Jr. High-valent nonheme iron-oxo complexes: Synthesis, structure, and spectroscopy. Coord. Chem. Rev. 2013, 257, 414–428. [Google Scholar]
- (8).Ortiz de Montellano PR Hydrocarbon Hydroxylation by Cytochrome P450 Enzymes. Chem. Rev. 2010, 110, 932–948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Poulos TL Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 3919–3962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Rittle J; Green MT Cytochrome P450 Compound I: Capture, Characterization, and C-H Bond Activation Kinetics. Seienee 2010, 330, 933–937. [DOI] [PubMed] [Google Scholar]
- (11).Huang X; Groves JT Beyond ferryl-mediated hydroxylation: 40 years of the rebound mechanism and C-H activation. JBIC, J. Biol. Inorg. Chem. 2017, 22, 185–207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).Makris TM; Koenig K.v.; Schlichting I.; Sligar SG The status of high-valent metal oxo complexes in the P450 cytochromes. J. Inorg. Bioehem. 2006, 100, 507–518. [DOI] [PubMed] [Google Scholar]
- (13).Denisov IG; Makris TM; Sligar SG; Schlichting I Structure and Chemistry of Cytochrome P450. Chem. Rev. 2005, 105, 2253–2278. [DOI] [PubMed] [Google Scholar]
- (14).Rettie AE; Sheffels PR; Korzekwa KR; Gonzalez FJ; Philpot RM; Baillie TA CYP4 Isoenzyme Specificity and the Relationship between ®-Hydroxylation and Terminal Desaturation of Valproic Acid. Biochemistry 1995, 34, 7889–7895. [DOI] [PubMed] [Google Scholar]
- (15).Rettie AE; Boberg M; Rettenmeier AW; Baillie TA Cytochrome P-450-catalyzed desaturation of valproic acid in vitro. Species differences, induction effects, and mechanistic studies. J. Biol. Chem. 1988, 263 (27), 13733–13738. [PubMed] [Google Scholar]
- (16).Rettie AE; Rettenmeier AW; Howald WN; Baillie TA Cytochrome P-450-catalyzed formation of A4-VPA, a toxic metabolite of valproic acid. Science 1987, 235, 890–893. [DOI] [PubMed] [Google Scholar]
- (17).Guengerich FP; Kim DH Enzymic oxidation of ethyl carbamate to vinyl carbamate and its role as an intermediate in the formation of 1,N6-ethenoadenosine. Chem. Res. Toxicol. 1991, 4, 413421. [DOI] [PubMed] [Google Scholar]
- (18).Nagata K; Liberato DJ; Gillette JR; Sasame HA An unusual metabolite of testosterone. 17 ^-Hydroxy-4,6-androstadiene- 3-one. Drug Metab. Dispos. 1986, 14 (5), 559–565. [PubMed] [Google Scholar]
- (19).Guan X; Fisher MB; Lang DH; Zheng Y-M; Koop DR; Rettie AE Cytochrome P450-dependent desaturation of lauric acid: isoform selectivity and mechanism of formation of 11- dodecenoic acid. Chem.-Biol. Interact 1998, 110, 103–121. [DOI] [PubMed] [Google Scholar]
- (20).Rude MA; Baron TS; Brubaker S; Alibhai M; Del Cardayre SB; Schirmer A Terminal Olefin (1-Alkene) Biosynthesis by a Novel P450 Fatty Acid Decarboxylase from Jeotgalicoccus Species. Appl. Environ. Microbiol. 2011, 77, 1718–1727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).Liu Y; Wang C; Yan J; Zhang W; Guan W; Lu X; Li S Hydrogen peroxide-independent production of a-alkenes by OleTJE P450 fatty acid decarboxylase. Biotechnol Biofuels 2014, 7, 28–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Faponle AS; Quesne MG; de Visser SP Origin of the Regioselective Fatty-Acid Hydroxylation versus Decarboxylation by a Cytochrome P450 Peroxygenase: What Drives the Reaction to Biofuel Production? Chem. - Eur. J 2016, 22, 5478–5483. [DOI] [PubMed] [Google Scholar]
- (23).Belcher J; McLean KJ; Matthews S; Woodward LS; Fisher K; Rigby SEJ; Nelson DR; Potts D; Baynham MT; Parker DA; Leys D; Munro AW Structure and Biochemical Properties of the Alkene Producing Cytochrome P450 OleTJE (CYP152L1) from the Jeotgalicoccus sp. 8456 Bacterium. J. Biol. Chem. 2014, 289, 6535–6550. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Grant JL; Mitchell ME; Makris TM Catalytic strategy for carbon-carbon bond scission by the cytochrome P450 OleT. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 10049–10054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Hsieh CH; Huang X; Amaya JA; Rutland CD; Keys CL; Groves JT; Austin RN; Makris TM The Enigmatic P450 Decarboxylase OleT Is Capable of, but Evolved To Frustrate, Oxygen Rebound Chemistry. Biochemistry 2017, 56, 3347–3357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Pickl M; Kurakin S; Cantu Reinhard FG; Schmid P; Pocheim A; Winkler CK; Kroutil W; de Visser SP; Faber K Mechanistic Studies of Fatty Acid Activation by CYP152 Peroxygenases Reveal Unexpected Desaturase Activity. ACS Catal 2019, 9, 565–577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).Mittra K; Green MT Reduction Potentials of P450 Compounds I and II: Insight into the Thermodynamics of C-H Bond Activation. J. Am. Chem. Soc. 2019, 141, 5504–5510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Hayashi Y; Yamazaki I The oxidation-reduction potentials of compound I/compound II and compound II/ferric couples of horseradish peroxidases A2 and C. J. Biol. Chem. 1979, 254 (18), 9101–9106. [PubMed] [Google Scholar]
- (29).Zaragoza JPT; Yosca TH; Siegler MA; Moenne-Loccoz P; Green MT; Goldberg DP Direct Observation of Oxygen Rebound with an Iron-Hydroxide Complex. J. Am. Chem. Soc. 2017, 139, 13640–13643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Ganguly S; Giles LJ; Thomas KE; Sarangi R; Ghosh A Ligand Noninnocence in Iron Corroles: Insights from Optical and X- ray Absorption Spectroscopies and Electrochemical Redox Potentials. Chem. - Eur. J 2017, 23, 15098–15106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Ganguly S; McCormick LJ; Conradie J; Gagnon KJ; Sarangi R; Ghosh A Electronic Structure of Manganese Corroles Revisited: X-ray Structures, Optical and X-ray Absorption Spectroscopies, and Electrochemistry as Probes of Ligand Noninnocence. Inorg. Chem. 2018, 57 (16), 9656–9669. [DOI] [PubMed] [Google Scholar]
- (32).Weitz AC; Mills MR; Ryabov AD; Collins TJ; Guo Y; Bominaar EL; Hendrich MP A Synthetically Generated LFeIVOHn Complex. Inorg. Chem. 2019, 58, 2099–2108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Drummond MJ; Ford CL; Gray DL; Popescu CV; Fout AR Radical Rebound Hydroxylation Versus H-Atom Transfer in Non-Heme Iron(III)-Hydroxo Complexes: Reactivity and Structural Differentiation. J. Am. Chem. Soc. 2019, 141, 6639–6650. [DOI] [PubMed] [Google Scholar]
- (34).Xu N; Powell DR; Richter-Addo GB Nitrosylation in a Crystal: Remarkable Movements of Iron Porphyrins Upon Binding of Nitric Oxide. Angew. Chem., Int. Ed. 2011, 50, 9694–9696. [DOI] [PubMed] [Google Scholar]
- (35).Cheng B; Safo MK; Orosz RD; Reed CA; Debrunner PG; Scheidt WR Synthesis, Structure, and Characterization of Five-Coordinate Aquo(octaethylporphinato)iron(III) Perchlorate. Inorg. Chem 1994, 33, 1319–1324. [Google Scholar]
- (36).Ohgo Y; Chiba Y; Hashizume D; Uekusa H; Ozeki T; Nakamura M Novel spin transition between S = 5/2 and S = 3/2 in highly saddled iron(III) porphyrin complexes at extremely low temperatures. Chem. Commun. 2006, 1935–1937. [DOI] [PubMed] [Google Scholar]
- (37).Ellis WC; Tran CT; Roy R; Rusten M; Fischer A; Ryabov AD; Blumberg B; Collins TJ Designing Green Oxidation Catalysts for Purifying Environmental Waters. J. Am. Chem. Soc. 2010, 132, 9774–9781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Ghosh A; Ryabov AD; Mayer SM; Horner DC; Prasuhn DE; Sen Gupta S; Vuocolo L; Culver C; Hendrich MP; Rickard CEF; Norman RE; Horwitz CP; Collins TJ Understanding the Mechanism of H+-Induced Demetalation as a Design Strategy for Robust Iron(III) Peroxide-Activating Catalysts. J. Am. Chem. Soc. 2003, 125 (41), 12378–12379. [DOI] [PubMed] [Google Scholar]
- (39).Simkhovich L; Goldberg I; Gross Z Iron(III) and Iron(IV) Corroles: Synthesis, Spectroscopy, Structures, and No Indications for Corrole Radicals. Inorg. Chem. 2002, 41, 5433–5439. [DOI] [PubMed] [Google Scholar]
- (40).Simkhovich L; Mahammed A; Goldberg I; Gross Z Synthesis and Characterization of Germanium, Tin, Phosphorus, Iron, and Rhodium Complexes of Tris(pentafluorophenyl)corrole, and the Utilization of the Iron and Rhodium Corroles as Cyclopropanation Catalysts. Chem. - Eur. J 2001, 7, 1041–1055. [DOI] [PubMed] [Google Scholar]
- (41).Zaragoza JPT; Siegler MA; Goldberg DP A Reactive Manganese(IV)-Hydroxide Complex: A Missing Intermediate in Hydrogen Atom Transfer by High-Valent Metal-Oxo Porphyrinoid Compounds. J. Am. Chem. Soc. 2018, 140, 4380–4390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (42).Blomberg MRA; Siegbahn PEM A comparative study of high-spin manganese and iron complexes. Theor. Chem. Acc. 1997, 97, 72–80. [Google Scholar]
- (43).Stein SE; Brown RL Prediction of carbon-hydrogen bond dissociation energies for polycyclic aromatic hydrocarbons of arbitrary size. J. Am. Chem. Soc. 1991, 113, 787–793. [Google Scholar]
- (44).Hansch C; Leo A; Taft RW A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar]
- (45).Lee JY; Peterson RL; Ohkubo K; Garcia-Bosch I; Himes RA; Woertink J; Moore CD; Solomon EI; Fukuzumi S; Karlin KD Mechanistic Insights into the Oxidation of Substituted Phenols via Hydrogen Atom Abstraction by a Cupric-Superoxo Complex. J. Am. Chem. Soc. 2014, 136, 9925–9937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Lucarini M; Pedrielli P; Pedulli GF; Cabiddu S; Fattuoni C Bond Dissociation Energies of O-H Bonds in Substituted Phenols from Equilibration Studies. J. Org. Chem. 1996, 61, 9259–9263. [Google Scholar]
- (47).Marcus RA; Sutin N Electron transfers in chemistry and biology. Biochim. Biophys. Acta, Rev. Bioenerg 1985, 811, 265–322. [Google Scholar]
- (48).Wijeratne GB; Corzine B; Day VW; Jackson TA Saturation Kinetics in Phenolic O-H Bond Oxidation by a Mononuclear Mn(III)-OH Complex Derived from Dioxygen. Inorg. Chem. 2014, 53, 7622–7634. [DOI] [PubMed] [Google Scholar]
- (49).Rice DB; Wijeratne GB; Burr AD; Parham JD; Day VW; Jackson TA Steric and Electronic Influence on Proton-Coupled Electron-Transfer Reactivity of a Mononuclear Mn(III)-Hydroxo Complex. Inorg. Chem. 2016, 55, 8110–8120. [DOI] [PubMed] [Google Scholar]
- (50).Sastri CV; Lee J; Oh K; Lee YJ; Lee J; Jackson TA; Ray K; Hirao H; Shin W; Halfen JA; Kim J; Que L Jr.; Shaik S; Nam W Axial ligand tuning of a nonheme iron(IV)-oxo unit for hydrogen atom abstraction. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 19181–19186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Kurahashi T; Kikuchi A; Shiro Y; Hada M; Fujii H Unique Properties and Reactivity of High-Valent Manganese-Oxo versus Manganese-Hydroxo in the Salen Platform. Inorg. Chem. 2010, 49, 6664–6672. [DOI] [PubMed] [Google Scholar]
- (52).Yiu DTY; Lee MFW; Lam WWY; Lau T-C Kinetics and Mechanisms of the Oxidation of Phenols by a trans- Dioxoruthenium(VI) Complex. Inorg. Chem. 2003, 42, 1225–1232. [DOI] [PubMed] [Google Scholar]
- (53).Das PK; Encinas MV; Steenken S; Scaiano JC Reaction of tert-butoxy radicals with phenols. Comparison with the reactions of carbonyl triplets. J. Am. Chem. Soc. 1981, 103, 4162–4166. [Google Scholar]
- (54).Finn M; Friedline R; Suleman NK; Wohl CJ; Tanko JM Chemistry of the t-Butoxyl Radical: Evidence that Most Hydrogen Abstractions from Carbon are Entropy-Controlled. J. Am. Chem. Soc. 2004, 126, 7578–7584. [DOI] [PubMed] [Google Scholar]
- (55).Massie AA; Sinha A; Parham JD; Nordlander E; Jackson TA Relationship between Hydrogen-Atom Transfer Driving Force and Reaction Rates for an Oxomanganese(IV) Adduct. Inorg. Chem. 2018, 57, 8253–8263. [DOI] [PubMed] [Google Scholar]
- (56).Kaizer J; Klinker EJ; Oh NY; Rohde J-U; Song WJ; Stubna A; Kim J; Munck E; Nam W; Que L Jr. Nonheme FeIVO Complexes That Can Oxidize the C-H Bonds of Cyclohexane at Room Temperature. J. Am. Chem. Soc. 2004, 126, 472–473. [DOI] [PubMed] [Google Scholar]
- (57).Wu X; Seo MS; Davis KM; Lee Y-M; Chen J; Cho KB; Pushkar YN; Nam W A Highly Reactive Mononuclear NonHeme Manganese(IV)-Oxo Complex That Can Activate the Strong C-H Bonds of Alkanes. J. Am. Chem. Soc. 2011, 133, 20088–20091. [DOI] [PubMed] [Google Scholar]
- (58).Lansky DE; Goldberg DP Hydrogen Atom Abstraction by a High-Valent Manganese(V)-Oxo Corrolazine. Inorg. Chem. 2006, 45, 5119–5125. [DOI] [PubMed] [Google Scholar]
- (59).de Visser SP Trends in Substrate Hydroxylation Reactions by Heme and Nonheme Iron(IV)-Oxo Oxidants Give Correlations between Intrinsic Properties of the Oxidant with Barrier Height. J. Am. Chem. Soc. 2010, 132, 1087–1097. [DOI] [PubMed] [Google Scholar]
- (60).Kumar D; Karamzadeh B; Sastry GN; de Visser SP What Factors Influence the Rate Constant of Substrate Epoxidation by Compound I of Cytochrome P450 and Analogous Iron(IV)-Oxo Oxidants? J. Am. Chem. Soc. 2010, 132, 7656–7667. [DOI] [PubMed] [Google Scholar]
- (61).Cantu Reinhard FG; Faponle AS; de Visser SP Substrate Sulfoxidation by an Iron(IV)-Oxo Complex: Benchmarking Computationally Calculated Barrier Heights to Experiment. J. Phys. Chem. A 2016, 120, 9805–9814. [DOI] [PubMed] [Google Scholar]
- (62).de Visser SP; Ogliaro F; Sharma PK; Shaik S What Factors Affect the Regioselectivity of Oxidation by Cytochrome P450? A DFT Study of Allylic Hydroxylation and Double Bond Epoxidation in a Model Reaction. J. Am. Chem. Soc. 2002, 124, 11809–11826. [DOI] [PubMed] [Google Scholar]
- (63).Li X-X; Postils V; Sun W; Faponle AS; Sola M; Wang Y; Nam W; de Visser SP Reactivity Patterns of (Protonated) Compound II and Compound I of Cytochrome P450: Which is the Better Oxidant? Chem. - Eur. J 2017, 23, 6406–6418. [DOI] [PubMed] [Google Scholar]
- (64).Lias SG Ionization Energy Evaluation in the NIST Chemistry WebBook In NIST Standard Reference Database Number 69; Linstrom PJ, Mallard WG, Eds.; National Institute of Standards and Technology: Gaithersburg, MD, 2005; pp 20899. [Google Scholar]
- (65).Chen J; Cho K-B; Lee Y-M; Kwon YH; Nam W Mononuclear nonheme iron(IV)-oxo and manganese(IV)-oxo complexes in oxidation reactions: experimental results prove theoretical prediction. Chem. Commun. 2015, 51, 13094–13097. [DOI] [PubMed] [Google Scholar]
- (66).Denler MC; Massie AA; Singh R; Stewart-Jones E; Sinha A; Day VW; Nordlander E; Jackson TA MnIV-Oxo complex of a bis(benzimidazolyl)-containing N5 ligand reveals different reactivity trends for MnIV-oxo than FeIV-oxo species. Dalton Trans. 2019, 48, 5007–5021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (67).Baglia RA; Prokop-Prigge KA; Neu HM; Siegler MA; Goldberg DP Mn(V)(O) versus Cr(V)(O) Porphyrinoid Complexes: Structural Characterization and Implications for Basicity Controlling H-Atom Abstraction. J. Am. Chem. Soc. 2015, 137, 9 10874–10877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (68).Kundu S; Miceli E; Farquhar ER; Ray K Mechanism of phenol oxidation by heterodinuclear Ni Cu bis(^-oxo) complexes involving nucleophilic oxo groups. Dalton Trans. 2014, 43, 4264–9 4267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (69).Sheldrick G Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C 2015, 71 (1), 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (70).Clark RC; Reid JS The analytical calculation of absorption in multifaceted crystals. Acta Crystallogr., Sect. A: Found. Crystallogr. 1995, 51, 887–897. [Google Scholar]
- (71).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Mennucci B.; Petersson GA; Nakatsuji H; Caricato M; Li X; Hratchian HP; Izmaylov AF.; Bloino J; Zheng G; Sonnenberg JL; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark M; Heyd JJ; Brothers E; Kudin KN; Staroverov VN; Kobayashi R;; Normand J; Raghavachari K; Rendell A; Burant JC; Iyengar SS; Tomasi J; Cossi M.; Rega N; Millam JM.; Klene M; Knox JE; Cross JB;Bakken V; Adamo C; Jaramillo J; Gomperts R; Stratmann RE; Yazyev O.; Austin AJ.; Cammi R; Pomelli C; Ochterski JW; Martin RL.; Morokuma K; Zakrzewski VG;Voth GA; Salvador P; Dannenberg JJ; Dapprich S; Daniels AD; Farkas O; Foresman JB; Ortiz JV; Cioslowski J; Fox DJ Gaussian 09, revision A.1; Gaussian, Inc.: Wallingford, CT, 2009. [Google Scholar]
- (72).Becke AD Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar]
- (73).Lee C; Yang W; Parr RG Development of the Colle- Salvetti correlation-energy formula into a functional of the electron 9 density. Phys. Rev. B: Condens. Matter Mater. Phys 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- (74).Hay PJ; Wadt WR Abinitio Effective Core Potentials for Molecular Calculations - Potentials for the Transition-Metal Atoms Sc 9 to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar]
- (75).Francl MM; Pietro WJ; Hehre WJ; Binkley JS; Gordon MS; DeFrees DJ; Pople JA Self-consistent molecular 9 orbital methods. XXIII. A polarization-type basis set for second-row 9 elements. J. Chem. Phys. 1982, 77 (7), 3654. [Google Scholar]
- (76).Tomasi J; Mennucci B; Cammi R Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [DOI] [PubMed] [Google Scholar]
- (77).Kumar S; Faponle AS; Barman P; Vardhaman AK; Sastri V; Kumar D; de Visser SP Long-Range Electron Transfer Triggers Mechanistic Differences between Iron(IV)-Oxo and Iron- (IV)-Imido Oxidants. J. Am. Chem. Soc. 2014, 136, 17102–17115. [DOI] [PubMed] [Google Scholar]
- (78).Yang T; Quesne MG; Neu HM; Cantu Reinhard FG; Goldberg DP; de Visser SP Singlet versus Triplet Reactivity in an 9 Mn(V)-Oxo Species: Testing Theoretical Predictions Against 9 Experimental Evidence. J. Am. Chem. Soc. 2016, 138, 12375–12386. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.