Abstract
The transfer of OH• from metal-hydroxo species to carbon radicals (R•) to give hydroxylated products (ROH) is a fundamental process in metal-mediated heme and nonheme C–H bond oxidations. This step, often referred to as the hydroxyl “rebound” step, is typically very fast, making direct study of this process challenging if not impossible. In this report, we describe the reactions of the synthetic models M(OH)(ttppc) (M = Fe (1), Mn (3); ttppc = 5,10,15-tris(2,4,6-triphenyl)phenyl corrolato3−) with a series of triphenylmethyl carbon radical (R•) derivatives ((4-X-C6H4)3C•; X = OMe, tBu, Ph, Cl, CN) to give the one-electron reduced MIII(ttppc) complexes and ROH products. Rate constants for 3 for the different radicals ranged from 11.4(1) to 58.4(2) M−1 s−1, as compared to those for 1, which fall between 0.74(2) and 357(4) M−1 s−1. Linear correlations for Hammett and Marcus plots for both Mn and Fe were observed, and the small magnitudes of the slopes for both correlations imply a concerted OH• transfer reaction for both metals. Eyring analyses of reactions for 1 and 3 with (4-X-C6H4)3C• (X = tBu, CN) also give good linear correlations, and a comparison of the resulting activation parameters highlight the importance of entropy in these OH• transfer reactions. Density functional theory calculations of the reaction profiles show a concerted process with one transition state for all radicals investigated and help to explain the electronic features of the OH rebound process.
Graphical Abstract

INTRODUCTION
Many biological catalysts that perform oxidation reactions utilize a heme ligand scaffold. One notable example is the enzyme Cytochrome P450 (CYP), which is a heme monooxygenase that can activate inert C–H bonds for hydroxylation with a high-valent iron oxo intermediate (Compound I) (Scheme 1).1–6 The rate-determining step in this process is the hydrogen atom transfer (HAT)7 from the C–H substrate to Compound I, forming an iron hydroxide species (protonated Compound II). Characterization of protonated Compound II was recently accomplished in the absence of substrate,8 but observation of this species during catalysis is essentially nonviable because of the rapid transfer of the OH group to the nascent carbon radical and the corresponding transient nature of the Fe(OH) intermediate. Computational studies on this OH “rebound” step are consistent with small reaction barriers to form alcohol products.9–12 However, direct experimental examination of the OH group transfer step has not been possible in an enzymatic setting. Our group has recently focused on examining synthetic model reactions of these rebound processes in order to obtain mechanistic information on this critical product-determining pathway.13,14 Our first success in this area came with the synthesis and structural characterization of a protonated Compound II model complex, Fe(OH)(ttppc) (1) (ttppc = 5,10,15-tris(2,4,6-triphenyl)-phenylcorrolato3−), which was stabilized by the use of a sterically encumbered corrole ligand. Reaction of this complex with substituted triphenylmethyl carbon radicals (R•) led to OH transfer to give hydroxylated product (ROH) and the one-electron reduced FeIII(ttppc) (2). Kinetic analyses of these reactions pointed to a concerted transfer of OH• from the metal to R•, as opposed to a stepwise process such as electron-transfer followed by cation-transfer (ET/CT).15 Related studies in nonheme chemistry have been carried out recently by us and others for Fe, Cu, and Ni.16–23
Scheme 1.

Hydroxyl Transfer Performed by CYP Enzymes
Synthetic, heme-based catalysts are also known to proceed through a mechanism similar to that shown in Scheme 1. In the synthetic systems, metals other than iron can be employed, and manganese porphyrins were employed successfully to carry out challenging C–H hydroxylation and halogenation reactions.24–30 There has been some effort to examine the mechanisms of these reactions, including a study with an MnV(O) porphyrin complex that examined the kinetics of rebound versus radical cage escape, which are divergent pathways following initial HAT. Because the putative MnIV(OH) intermediate could not be isolated, product distributions were employed to assess relative rates of cage escape and rebound reaction pathways.30 Divergent pathways for the rebound step in CYP enzymes may also occur, such as hydroxylation versus alkane desaturation in olefin fatty acid decarboxylase (OleT) and related CYP152 enzymes.31–35 The structural and electronic characteristics of both the active sites and substrates that may favor hydroxylation versus desaturation pathways remain subjects of significant research.36,37
In this work, we report a combined experimental and computational study on the reactivity of M(OH)(ttppc) (M = Fe (1), Mn (3)) toward triphenylmethyl radical derivatives. These complexes are formally in the +4 oxidation state, although corroles are well-known to behave as noninnocent ligands, making metal oxidation state assignments ambiguous. However, the overall oxidation level for both the Mn and Fe complexes is the same as that of the protonated Compound II intermediate in heme enzymes. The synthesis and structure of complex 3 was reported previously,38 and it is essentially isostructural with iron-containing 1. Together with expanded, new studies on 1, we provide a direct comparison of OH• transfer reactivity for Mn(OH) and Fe(OH) complexes. This comparative analysis has led to significant progress in our goal of understanding the structural, thermodynamic, and kinetic factors that contribute to OH• transfer, and related “rebound” processes, in heme enzymes and catalysts.
PRODUCT ANALYSIS
Examination of the reactivity of the Mn(OH) complex 3 with carbon radicals was initiated with triphenylmethyl radical, as shown in Scheme 2. The stable, crystalline compound Ph2C=C6H4–CPh3 is in equilibrium with ~2% of the radical form, Ph3C•, in toluene.39,40 Addition of six equivalents of Ph2C=C6H4–CPh3 to 3 in toluene at 23 °C led to a color change from brown to green over a period of 2 h. Monitoring the reaction by UV–vis spectroscopy revealed isosbestic conversion of the spectrum for 3 to a spectrum that corresponded to that of the reduced MnIII(ttppc) (4) (λmax = 417, 435, 450, 660 nm).38 One-electron reduction of 3 is consistent with formal OH• transfer to the triphenylmethyl radical and production of the alcohol, Ph3COH. The UV–vis spectrum for MnIII(Et2O)(ttppc) was reported previously (λmax = 413, 446, 660 nm)41 and was therefore used as a standard to assess the yield of the MnIII complex. Addition of excess Et2O to the final reaction mixture resulted in small changes in the UV–vis spectrum consistent with formation of the MnIII(Et2O) adduct. Quantitation of the absorbance at 660 nm gave a yield of 88 ± 1% (average of 3 runs) for MnIII(ttppc) (Figure S3).
Scheme 2.

Reaction of Mn(OH)(ttppc) with (4-X-C6H4)C• (X = H, tBu)
Formation of the expected alcohol, (Ph3COH), was measured by 1H NMR spectroscopy. Reaction of 3 and 6 equiv of Ph2C=C6H4–CPh3 in toluene-d8 led to a change in the 1H NMR spectrum. The aromatic region is complicated by overlapping peaks from the starting Gomberg’s dimer, but formation of a well-separated singlet at 2.28 ppm can be assigned to the hydroxyl group of Ph3COH based on comparison with an independent sample. The yield of alcohol was quantified by integration of the peak at 2.28 ppm and comparison against an internal standard (Figure S1), giving a yield of 90 ± 1% (average of three runs). This yield is an excellent agreement with the yield of reduced MnIII complex.
Addition of the para-tert-butyl-substituted radical derivative (4-tBu-C6H4)3C• to 3 led to similar reactivity. This radical is generated in situ from treatment of the corresponding triphenyl methyl halide with Cu0,42 giving a stable radical in solution that is blocked through the para substituent from the dimerization seen for Ph2C=C6H4–CPh3. Addition of (4-tBu-C6H4)3C• to 3 in toluene-d8 led to the expected color change from brown to green. The monitoring of this reaction by 1H NMR spectroscopy revealed production of (4-tBu-C6H4)3COH, which exhibits a unique singlet at 1.21 ppm corresponding to the para-tert-butyl substituents. Integration of this peak and comparison with an internal standard afforded a yield of 90 ± 1% (Figure S2), matching that found for the unsubstituted triphenylmethyl radical.
KINETICS
Substitution of the para-phenyl positions leads to radical derivatives that remain monomeric in solution, providing a convenient series of radicals for kinetics measurements as shown previously for Fe(OH)(ttppc) with (4-X-C6H4)3C• (X = OMe, tBu, Cl, Ph) derivatives.15 The kinetics of the reaction of 3 with the radical derivatives (4-X-C6H4)3C• (X = OMe, tBu, Cl, Ph, CN) in toluene was monitored by UV–vis spectroscopy. Addition of excess radical (pseudo-first-order) leads to the isosbestic conversion of 3 to 4. These changes are shown in Figure 1 for the tBu derivative. The peak for the radical at 525 nm remains unchanged because it is in large excess, while the Soret band for 3 at 420 nm disappears at the same time that the Q-band for 4 grows in at 660 nm. A plot of the absorbance at 660 nm versus time is given in the inset of Figure 1b, showing single exponential behavior that was fit to give a pseudo-first-order rate constant kobs = 6.67 × 10−3 s−1. Varying the concentration of the tBu radical derivative led to a linear plot of kobs versus [(4-tBu-C6H4)3C•], and the slope of the best-fit line gave a second-order rate constant, k2 = 30.5(1) M−1 s−1, which is similar in value to the 49(1) M−1 s−1 value recorded for this reaction with the analogous Fe complex 1.
Figure 1.

(a) Overlay of UV–vis spectra for 3 (red, dashed), 4 (blue, dotted), and (4-tBu-C6H4)3C• (green, solid). (b) UV–vis spectra for the reaction of 3 (18 μM) with (4-tBu-C6H4)3C• (20 equiv) in toluene-d8 from 0 (red) to 400 (blue) s at 50 s intervals (gray). Inset: plot of absorbance at 660 nm versus time (open green circles) showing the growth of the MnIII complex, together with the best fit line.
Reactions of 3 with the other (4-X-C6H4)3C• derivatives were carried out similarly to the tBu derivative, as shown in the scheme in Figure 2. Reaction kinetics were monitored by UV–vis spectroscopy, and each radical exhibited good pseudo-first-order kinetics, leading to second order rate constants from 11.4(1) to 58.4(1) M−1 s−1 for the different radical derivatives (Figures S6–S11, Supporting Information). The rate constants are listed in Figure 2 together with the sum of the Hammett parameters (3σ+) for each of the para substituents, with lower σ+ values corresponding to the more electron-donating substituents.43 The rate constants vary by a relatively modest factor of ~5, despite a relatively large change in Hammett values. A Hammett plot, i.e., log(k2) versus 3σ+, for 3 is shown in Figure 3, together with the same plot for the analogous Fe(OH) complex 1.15 Both compounds exhibit a linear correlation, but the line for 3 exhibits a significantly smaller Hammett slope (ρ = −0.15) than for 1 (ρ = −0.62). These trends indicate that the rates of OH• transfer from Mn(OH)(ttppc) to the radical derivatives are much less sensitive to the electron-rich nature of the carbon radical center and imply that there is less charge-transfer in the transition state for the Mn(OH) as compared to the Fe(OH) complex.15,44 A crossover in relative rates of reaction for Mn versus Fe is also observed because of the difference in ρ values in Figure 3, with the crossover point close to σ+ = 0. This change in reactivity leads to the Mn complex being ~15 times more reactive toward the electron-poor para-CN substrate than the Fe analogue.
Figure 2.

Reaction of M(OH)(ttppc) (M = Fe, Mn) with para-X-substituted triphenylmethyl radicals (X = OMe, tBu, Ph, Cl, CN). Hammett parameters (3σ+) and redox potentials (Eox versus Fc/Fc+, MeCN) for the different radicals,43,45 and the corresponding second order rate constants for reactions with the Fe(OH) and Mn(OH) complexes are given below each radical structure.
Figure 3.

Hammett plot for the reactions of Mn(OH)(ttppc) (blue triangles) and Fe(OH)(ttppc) (red diamonds) with (4-X-C6H4)3C• (X = OMe, tBu, Ph, Cl, CN).
SPECTROELECTROCHEMISTRY
The redox potential of 3 was measured by a combination of cyclic voltammetry (CV) and spectroelectrochemistry in order to gain further insight into the relative reactivity of 3 versus 1 for the different substrates. The CV of 3 is shown in Figure 4a together with that previously reported for 1. The CV of 3 in PhCN exhibits two reversible waves at −0.07 and +0.60 V versus ferrocene/ferrocenium (Fc+/Fc) and a third, irreversible peak at −1.04 V. Comparison with the CV for 4 suggests that the reversible couple at E1/2 = −0.07 V is the formal [MnIV(OH)]0/[MnIII(OH)]− reduction potential, with the analogous couple at −0.11 V for 1 assigned previously to the formal [FeIV(OH)] 0/[FeIII(OH)]− couple.15 These assignments were corroborated by spectroelectrochemical measurements. Controlled potential electrolysis (CPE) of 3 in PhCN at −0.30 V was accompanied by isosbestic changes in the UV–vis spectrum that indicated the one-electron reduction of 3 (λmax = 425 nm) to [MnIII(ttppc) (λmax = 450, 514, 675 nm) (Figure 4c). The same solution was reoxidized by CPE at Eappl = +0.10 V, leading to the restoration of the spectrum for 3 (97% yield) (Figure S4). The CPE measurements on 1 in PhCN led to similar one-electron reversible behavior (Figure 4d) at applied potentials bracketing the −0.11 V reversible wave in Figure 4a and confirmed that this couple corresponds to the reversible one-electron reduction of the Fe(OH)(ttppc) complex (Figure S5).15
Figure 4.

(A) Cyclic voltammograms of Fe(OH)(ttppc) (red) and Mn(OH)(ttppc) (blue) in benzonitrile at 23 °C, with 0.1 M Bu4NPF6 supporting electrolyte. Electrochemical data for 1 reproduced from ref 15. (B) Marcus plots for the reactions of 3 (blue triangles) and 1 (red diamonds) with (4-X-C6H4)3C• (X = OMe, tBu, Ph, Cl, CN). C–D) UV–vis spectral changes for 1 and 3, respectively, at Eappl = −0.30 V, in benzonitrile at 23 °C.
The two values are similar and also indicate that electron transfer (ET) reactions with these radical substrates will be exergonic, with the exceptions of the p-Cl and presumably p-CN radicals. Thus, to determine the role that ET may play in these reactions, a Marcus plot was produced (Figure 4b) by plotting (RT/F) ln(k2) vs the Eox of the radicals used.45 For reactions where the thermodynamic driving force (−ΔG°et) is significantly less than the reorganization energy (λ), a Marcus plot with slope = −0.5 indicates that a simple electron transfer is involved in the rate limiting step of the reaction.46 The Marcus plot for the reaction of the Mn complex 3 reveals a linear relationship (R2 = 0.957) with a slope of −0.05, which is close to the zero slope expected for a reaction that does not involve charge separation in the transition state.47,48 These data are consistent with a concerted group transfer of OH•, rather than a stepwise process involving reduction of the metallocorrole followed by the resulting para-substituted triphenylmethyl cation bonding with the OH− ligand (an electron-transfer/cation-transfer (ET/CT) mechanism).
EYRING ANALYSIS
We examined the temperature dependence of the reaction rates for the MnOH and FeOH complexes with the para-tBu substituted triphenylmethyl derivative (Figures S12, S13) between −5 and 35 °C for Mn and −5 and 23 °C for Fe. The resulting Eyring plots are linear and are shown in Figure 5a. The activation parameters ΔH‡ and ΔS‡ were obtained from the best-fit lines of the data for both the Mn and Fe complexes and are also given in Figure 5a. The results show that the activation enthalpies for MnOH and FeOH are similar (within ~2 kcal mol−1), but the entropies differ by approximately a factor of 3, with the MnOH complex exhibiting the more negative ΔS‡. For both complexes, the –TΔS‡ term remains smaller in magnitude than ΔH‡ throughout the experimental temperature range, indicating that these reactions remain under enthalpy control. However, the significantly larger magnitude of the ΔS‡ for the MnOH complex suggests a more ordered transition state (TS) for this complex. The weaker slopes in the Hammett and Marcus plots for MnOH compared to FeOH are also in good agreement with this relative ordering of activation entropies, since they suggest a more concerted reaction for MnOH and thus a more ordered TS.
Figure 5.

Eyring plots for the reactions of Mn(OH)(ttppc) (blue triangles) and Fe(OH)(ttppc) (red squares) with (A) (4-tBu-C6H4)3C• and (B) (4-CN-C6H4)3C•.
These trends were even more pronounced when we examined the temperature dependence for the reactions with the para-CN substituted radical (Figures S14, S15). As shown in Figure 5b, the data remain linear for the relatively electron-poor para-CN derivative, although the temperature ranges were changed to +15 to +55 °C for Mn and +23 to +65 °C for Fe, because these reactions were relatively slow and higher temperatures were employed to obtain reasonable reaction times. Fitting of the data resulted in the activation parameters for the MnOH and FeOH complexes listed in Figure 5b. The enthalpy terms are slightly smaller than those obtained for the para-tBu derivative, but the most dramatic change occurs in the ΔS‡ terms, which are significantly larger in magnitude than those obtained for the para-tBu derivative. The activation entropy for the FeOH complex is still within the expected range for a bimolecular reaction, but the activation entropy for the MnOH complex is significantly more negative. It is closer to that seen, for instance, for a Diels–Alder reaction, with a highly concerted and ordered transition state.49 The –TΔS‡ term for the MnOH complex at 298 K is 12.0 kcal mol−1, which is ~3-fold larger than the ΔH‡ value for this complex, and indicates this reaction is entropy-controlled near room temperature. A graphical representation of entropic versus enthalpic control of these reactions is shown in Figure S16. The aforementioned result is consistent with a highly concerted reaction in which good orbital overlap is required between the carbon radical and the acceptor orbital on the MnOH unit in order for productive OH· transfer to occur. Such a mechanism is also in agreement with the electron-poor nature of the carbon radical in this case, which strongly inhibits the involvement of stepwise electron-transfer and instead favors a concerted radical rebound process. Entropy-controlled reactions are not common50–55 but do include examples of radical recombination reactions in which the radicals must enter a specific orientation in a highly ordered TS.54,56 It is interesting to note the parallels between a radical–radical coupling reaction and the radical rebound reactions described here, where the MnOH unit is “coupling” with the incoming carbon radical in a specific, productive orientation to make the new C–O bond.
COMPUTATIONAL STUDIES
We started the computational modeling with calculations of the reactant species, i.e., [FeIV(OH)(ttppc)] and [MnIV(OH)-(ttppc)]. Our overall computational findings are similar to those found in a previous study examining these complexes in reactions with substituted phenol substrates.57 Geometry optimizations were performed for the iron complex in the lowest energy singlet, triplet and quintet spin state structures and for the manganese system in the doublet, quartet and sextet spin states. The calculations on the iron-hydroxo complex give a triplet spin ground state (31), with the singlet and quintet states higher-lying at ΔE+ZPE = 13.3 and 5.8 kcal mol−1, respectively. The manganese-hydroxo complex gives a quartet spin ground state (43). The optimized geometry for 43 is shown in Figure 6, together with selected data for 31. The Fe–O distance in 31 is 1.86 Å, while the corresponding manganese complex 43 has an Mn–O distance of 1.87 Å. The other spin state structures (coordinates given in SI) give similar distances. These distances match the experimentally determined single crystal X-ray diffraction results, with reported values of Fe–O = 1.857(3) Å and Mn–O = 1.881(2) Å.15,38 The triplet spin structure (31) gives a calculated ν(Fe–O) = 548 cm−1, which compares well with the experimental value of 576 cm−1 from resonance Raman spectroscopy. The computational methods give good agreement with the experimentally observed structural and vibrational features, helping to validate the overall computational results.
Figure 6.

UB3LYP/BS1 optimized geometry for 43 as obtained in Gaussian-09, together with relevant parameters for 31. Bond lengths (Fe–O, Mn–O) are in angstroms and spin densities (ρ) in atomic units. C atoms in brown or orange, N atoms in blue, Fe atom in purple, O atom in red, H atoms in gray. Values for 31 are given in parentheses.
Group spin densities of both 31 and 43 give a dominant spin contribution on the metal, although a significant amount of unpaired spin density is also seen on the corrole ligand: ρttppc = −0.7 for Fe and −0.6 for Mn. These spin densities implicate an electronic configuration with an unpaired down-spin electron on the corrole coupled to three/four up-spin electrons in metal-type orbitals for iron/manganese. The electronic configurations indicated by the calculations are δx2-y21 π*xz1 π*yz1 σ*z21 a2u1 for 43 and δx2-y22 π*xz1 π*yz1 σ*z21 a2u1 for 31. These computed configurations can be described as 4[MnIII(OH)(ttppc+*)] and 3[FeIII(OH)(ttppc+*)]. This type of electronic structure is different from that suggested by DFT calculations for the analogous high-valent species in porphyrins or heme enzymes, which usually converge to FeIV(OH)-(porph) structures with no spin on the macrocycle.58,59 However, calculations on the analogous metallocorroles commonly show spin density on the macrocycle, suggesting π-radical-cation character. Attempts were made to interchange molecular orbitals and obtain an alternative 31’ structure with configuration δx2-y22 π*xz1 π*yz1 σ*z20 a2u2. However, this state relaxed back to the [FeIII(OH)(ttppc+*)] state during the SCF convergence.
The mechanism of reaction between 1,3,51 or 2,4,63 and the triphenylmethyl substituted radical substrates (4-X-C6H4)3C• (SubX, where X = H, tBu, OMe, Cl, Ph, and CN) was examined by DFT. The lowest energy reaction barriers, however, are for the overall quartet spin transition state for iron and quintet spin transition state for manganese and start from the interaction of 31 with 2SubX or 43 with 2SubX. Note that the electronic configurations are retained in the reactant complexes with substrate and metal-hydroxo complex in close proximity, indicating that no long-range electron transfer takes place. No spin-state crossing was observed for any of the calculated reactions for 31 or 43 (calculations for the other spin states are included in the Supporting Information). The reactions start from a reactant complex composed of M(OH)(ttppc) and SubX in close proximity (structure ReM,X) and are followed by a single transition state (TSreb,M,X) that separates them from alcohol product complexes (PM,X). A concerted, one-step reaction mechanism is found for all pathways, and no long-range electron transfer is predicted.
Representative transition states 5TSreb,Mn,X and 4TSreb,Fe,X are given in Figure 7 and Figure S17, respectively, with the imaginary mode highlighted as a vector diagram. The activation enthalpies calculated at ΔE‡+ZPE level of theory (UB3LYP/BS2//UB3LYP/BS1 energies with solvent and zero-point corrections included) are given in Table 1. The ΔE‡+ZPE value for 5TSreb,Mn,tBu is calculated to be higher in energy by 5.9 kcal mol−1 versus that of 5TSreb,Mn,CN, which is close to the difference in the activation enthalpy terms (ΔΔH‡ = 7.0 kcal mol−1) obtained from the Eyring plots for the same substrates (Figure 5). A similar trend is seen for 5TSreb,Fe,tBu versus 5TSreb,Fe,CN in which the difference in the ΔE‡+ZPE values for the two substrates give a ΔΔH‡ = 5.0 kcal mol−1, compared to the experimental ΔΔH‡ = 4.3 kcal mol−1. Although the calculations overestimate the absolute values for ΔH‡ by 3–4 kcal mol−1, the trends clearly indicate that the enthalpic contribution to the barrier is higher for the tBu substrate than that for the CN substrate. Previous validation studies of experimental rate constants against our computational methods showed that the correct trend in reactivity and chemoselectivity was obtained despite a systematic error.60,61 Given the second order rate constants for the p-tBu substrate reaction with 1 and 3 are faster than those for the p-CN (Figure 2), these calculated ΔE‡+ZPE values imply that the ΔS‡ for the hydroxyl radical transfer reaction with the p-CN substrate must be greater in magnitude than that for the p-tBu substrate. This finding further supports our experimental observations of larger ΔS‡ for the p-CN substrate (Figure 5). However, the calculated ΔS‡ and their associated ΔG‡ values do not match those found experimentally (Tables S11, S13). These deviations in calculated versus experimentally observed ΔS‡ values may be due to the limitations of using ideal gas assumptions to estimate the translational, rotational, and vibrational entropies. In particular, the computational entropies do not take into consideration environmental effects such as solvent participation and the involvement of a solvent cage around the cluster. The large difference in dipole moment in the various transition states could be responsible for differences in local environment and the solvent cage. Recent studies on enzymatic reaction mechanisms have shown that induced dipole moment interactions in the protein can dramatically affect barrier heights and rate constants and even guide a selectivity preference.62–64 As such, differences in dipole moment will affect the free energies of activation through solvent reorganization and stabilization of the transition state structures.
Figure 7.

UB3LYP/BS1 optimized geometry (displayed as a vector diagram) of 4TSreb,Mn,Cl as obtained in Gaussian ian-09.
Table 1.
UB3LYP/BS1 Optimized Transition State Parameters (Referring to TSreb,M,X) for Both the M(OH)(ttppc) (M = Mn and (Fe)) Complexes, in Their Reactions with (4-X-C6H4)3C• (X = OMe, tBu, Ph, H, Cl, or CN)
| M = Mn and (Fe), X: | OMe | tBu | Ph | H | Cl | CN |
|---|---|---|---|---|---|---|
| ΔE‡+ZPE (kcal mol−1)a | 10.8 (12.7) | 15.6 (17.1) | 14.6 (16.2) | 15.6 (17.3) | 14.7 (16.6) | 9.7 (12.1) |
| R(M–O) (Å) | 2.051 (2.017) | 1.980 (1.954) | 1.983 (1.960) | 1.978 (1.956) | 1.994 (1.973) | 2.066 (2.032) |
| R(O–C) (Å) | 2.567 (2.428) | 2.663 (2.550) | 2.580 (2.534) | 2.589 (2.484) | 2.431 (2.360) | 2.266 (2.230) |
| Imag. Freq. (cm−1) | i69 (i117) | i89 (i125) | i131 (i131) | i146 (i176) | i241 (i278) | i396 (i465) |
| ΔμDipole(Debye)b | 7.8 (6.1) | 3.7 (3.3) | 3.1 (4.9) | 1.1 (2.3) | −3.2 (−3.2) | −2.7 (−2.0) |
| QCT (C)c | 0.67 (0.65) | 0.44 (0.46) | 0.41 (0.50) | 0.38 (0.41) | 0.27 (0.30) | 0.21 (0.25) |
Change in enthalpy + zero point energy for the transition state.
Change in dipole moment, comparing the reactant complex and transition state complex.
Charge transfer in the TS, measured as the number of electrons transferred to the metallocorrole from (4-X-C6H4)3C•.
The Fe–O bond length (Table 1) for the different SubX reactions is elongated by a similar amount (Δ(Fe–O) = 0.09 Å for 4TSreb,Fe,tBu and Δ(Fe–O) = 0.17 Å for 4TSreb,Fe,CN, which were the smallest and largest changes, respectively). The C–O distances in the transition states range from 2.230 Å for 4TSreb,Fe,CN to 2.550 Å for 4TSreb,Fe,tBu, which are long for a C–O bond and are consistent with an early TS. The C–O bond in the TS for the p-CN derivative is significantly shorter than those for the other derivatives, which indicates it has a tighter transition state than the other substrates. To put these values in perspective, the C–O distances are longer than those reported for the rebound step in cycloheptatriene hydroxylation by an iron(IV)(oxo)(pentafluorophenyl-porphyrin) complex, which found doublet and quartet spin transition states with C–O distances of 1.831 and 2.053 Å.65 Calculations for the small aliphatic substrates methane and propene and iron(IV)(oxo)(porphine) lead to much longer C–O distances of 2.556 and 2.427 Å, respectively, in the rebound transition states.12 Long C–O distances were also found for the rebound transition states during methane hydroxylation by μ-nitrido-bridged diiron(IV)-oxo porphyrinoid complexes.66 A large variation in geometric structure of rebound transition states is seen in the former examples, although there appears to be some dependence on the size of the substrate. The data in Table 1 show no clear trends in geometric features of the transition states for hydroxyl transfer for the different substrates.
In contrast to the structural parameters, a trend is seen in the imaginary frequencies for the transition states in Table 1, ranging for iron from i117 cm−1 for 4TSreb,Fe,OMe to i465 cm−1 for 4TSreb,Fe,CN, and for manganese from i69 to i396 cm−1. These imaginary frequencies identify a maximum along the potential energy surface for the transition state, and their magnitudes reflect the breadth of the potential energy curve around the transition state. Thus, shallow barriers have low imaginary frequencies, whereas steep barriers tend to have large imaginary frequencies. This trend means that the curvature around the barrier gets steeper as the electron-withdrawing character of the substituent increases. Stronger electron-donating substituents have broader potential energy surfaces, as seen from the smaller imaginary frequencies.
To understand the observed reactivity trends and distinguish between radical and charge-transfer pathways, a valence bond diagram was generated for the Fe complex (Scheme 3). This type of diagram give insight into the possible electron transfers and orbital reorganizations that happen during the transition state and have, for instance, been used previously to explain reactivity trends of hydrogen atom abstractions and regioselectivities.66,67 The differences in orbital-breaking and orbital-formation processes associated with a concerted, radical-type OH transfer versus long-range electron transfer were examined. The valence configuration of reactants and products are shown, helping to illustrate the electron transfer processes as well as the molecular orbital changes that occur during the reaction. The calculated electronic ground state for 1 has orbital occupation δx2-y22 π*xz1 π*yz1 σ*z21 a2u1, and these valence electrons are identified with a dot in Scheme 3. A molecular orbital is depicted as a line between two dots. The δx2-y2, π*xz, and σ*z2 orbitals are occupied with four electrons, have dominant contributions on the iron, and can be considered as atomic orbitals. The πyz π*yz pair of orbitals forms a two-center-three-electron bond along the Fe–O unit, while there is an unpaired electron on the substrate (in φSub) and one on the corrole ligand (in a2u).
Scheme 3.

Concerted Radical Transfer Versus Electron Transfer Pathways for Reaction of Triphenylmethyl Radical with Fe(OH)(ttppc)
If we consider a concerted reaction process for the OH• transfer (downward pathway in Scheme 3), then the OH• transfer to the SubX radical will form a C–O bond (σCO) by pairing up the radical on the substrate (from φSub) and one electron on the hydroxo group. The latter comes from the πyz π*yz two-center-three-electron bond that is split back into atomic orbitals: one electron forms the σCO bond, one electron remains as 3dyz on iron, and the third electron is promoted into the a2u orbital. In valence bond theory, the product electronic configuration in the geometry of the reactants determines the height of the transition state, and therefore, for a concerted OH• transfer, the barrier should correlate with the energy to form the σCO bond (Eσ(CO)), the energy to split the two-center-three-electron πyz/π*yz orbitals back into atomic orbitals (Eπ/π*yz) and the electron affinity of the corrole (EAcor).
By contrast, an initial outer-sphere electron-transfer (ET) will ionize the substrate and fill the corrole a2u orbital with a second electron. A reaction starting with outer-sphere ET should have a transition state energy that correlates with the ionization energy of the substrate (IESubX) and the reduction potential or electron affinity of the metal corrole complex (EAcor). The IESubX and Eσ(CO) values for all substrates SubX were calculated and are plotted in Figure 8 versus 3σ+ Hammett parameter. Both trends seen for IESubX and Eσ(CO) are consistent with the second order rate constants observed experimentally. The decreasing EC–O with 3σ+ should lower the overall driving force of the reaction, which is consistent with the observed slower reaction rates for the more electron-withdrawing σ+ values. However, the changes in the calculated EC–O bond energies is relatively small, ranging from 26.4 kcal mol−1 for the (4-OMe-C6H4)3C-OH to 18.6 kcal mol−1 for the (4-CN-C6H4)3C-OH substrates. The differences in IE values are larger, from 105 to 132 kcal mol−1 over the substrate range. This trend in substrate IE is also consistent with the degree of charge transfer calculated for the transition states of these hydroxyl transfer reactions (Table 1, QCT), which decreases as Hammett parameter increases, from 0.67 Coulombs for the OMe reagent to 0.23 Coulombs for the CN reagent. These trends, taken together with the VB analysis, support the notion that these reactions occur through a concerted •OH transfer with varying degrees of charge-transfer, and imply, more broadly, that the reaction pathway is influenced by the electronic characteristics of the substrate being modified.
Figure 8.

Ionization energy of the substrate (IESubX) and energy of the sigma C–O bond formed by hydroxyl transfer (Eσ(CO)) to (4-X-C6H4)3C• (X = H, OMe, tBu, Ph, Cl, CN).
CONCLUSIONS
This work provides the first study of reactions involving OH• transfer from a well-defined, Mn(OH) complex to a carbon radical. These reactions are analogous to the “rebound” step in C-H bond hydroxylation carried out by high-valent metal-oxo porphyrins, including the hydroxylation mediated by heme monooxygenases such as Cytochrome P450. The reactivity of the closely related Fe(OH) complex, which was previously described in preliminary form,15 was expanded here and provided a direct comparison between Mn(OH) and Fe(OH) species. The kinetic data obtained from a series of substituted triphenylmethyl radical derivatives show that both Mn(OH) and Fe(OH) complexes give linear Hammett and Marcus plot correlations that suggest the electronic nature of the tertiary carbon radicals plays a relatively minor role in the overall reaction rates, and point to a concerted, as opposed to stepwise (ET/CT), mechanism for OH• transfer. However, the Fe(OH) complex does have a significantly larger Hammett ρ value than the Mn(OH) complex, causing the Hammett plots for the two species to intersect near σ+ ~ 0. Similar trends in Hammett slopes were seen for 1 and 3 in their reactions with para-substituted phenol HAT substrates.57 These data hint at a greater selectivity for Fe(OH) versus Mn(OH) rebound toward electronically differentiated substrates, which may be advantageous for controlling product distribution in the native, iron-containing biological systems. A study on porphyrin Compound-I analogs provides support for this idea in which differences in free energies of activation, resulting from the entropies of activation, for epoxidation versus hydroxylation help determine product selectivity.51
Temperature-dependent kinetic studies showed that for the tBu-substituted triphenylmethyl radical, both Fe(OH) and Mn(OH) reactions are under enthalpy control, although the Mn complex exhibits a larger entropic factor. In comparison, the entropy terms for the electron-poor CN-substituted derivative are much larger, and in the case of Mn(OH), leads to the more rare situation of an entropy-controlled reaction at 298 K (−TΔS‡ ≫ ΔH‡). This finding is consistent with a concerted reaction involving a highly ordered transition state that requires good orbital overlap between reaction partners. These mechanistic features can be found in orbitally directed Diels–Alder reactions or in radical–radical coupling reactions, both of which provide other examples of entropy-controlled processes.49,54,56
The DFT calculations of the reaction trajectories for these model rebound reactions provide good support for the experimentally derived mechanistic scenario: a concerted process with a single transition state is computed for both Mn(OH) and Fe(OH) with all of the substrate derivatives. The electronic structures of the transition states show relatively small amounts of charge transfer, also consistent with the experimental Hammett and Marcus plots. The overall energetics of the different calculated reaction barriers do not match the experimental kinetics, but this discrepancy in fact provides further support for the importance of the entropic factors in these reactions. Whether the major entropic contributions come from an ordering of the reaction partners in the T.S. or from environmental effects (e.g., solvent) remains an open question but suggests that entropy may play a key role in determining the ultimate outcome of the rebound step in heme enzymes.
EXPERIMENTAL SECTION
Materials.
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 and acetonitrile were purified via a Pure-Solv solvent purification system from Innovative Technologies, Inc. Benzene and benzonitrile were obtained from commercial sources. Deuterated solvents for NMR were purchased from Cambridge Isotope Laboratories, Inc. (Tewksbury, MA). The (4-tBu-C6H4)3CBr,68 (4-Ph-C6H4)3CBr,69,70 (4-Cl-C6H4)3CBr,15 and (4-CN-C6H4)3CBr71 were synthesized following previously reported procedures. (4-OMe-C6H4)3CCl was obtained from Alfa-Aesar and used as received. All substituted triphenylmethyl radicals were generated by reduction of their corresponding halide with excess copper powder, as previously reported.15,16 The unsubstituted triphenylmethyl radical source, Gomberg’s dimer (Ph2C=C6H4–CPh3), was synthesized following literature procedures.39,40 Fe(OH)(ttppc), FeIII(ttppc), Mn(OH)-(ttppc), and MnIII(ttppc) were synthesized and purified as previously reported.15,38,41 The salt (Bu4N)PF6 was purchased from Sigma-Aldrich, and recrystallized from ethanol twice before use.
Instrumentation.
Kinetics and other UV–vis spectroscopic measurements were performed on a Hewlett-Packard Agilent 8453 diode-array spectrophotometer with a 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. For all UV–vis spectroscopy measurements acquired above or below 23 °C, and for spectro-electrochemical analysis, data collections were performed on a Varian Cary 60 Bio spectrophotometer. 1H NMR spectra were recorded on a Bruker Avance 400 MHz NMR spectrometer at 298 K, and referenced against residual solvent proton signals. 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 Bu4NPF6 in CH3CN), and a platinum wire counter electrode. Spectro-electrochemical cyclic voltammograms or controlled potential electrolysis measurements were collected on a 6-channel Ivium-n-Stat with a gold honeycomb card (Pine Research part AB01STC1 AU) working and counter electrode and a Ag/AgNO3 nonaqueous reference electrode (0.01 M AgNO3 with 0.1 M Bu4NPF6 in CH3CN), with samples placed in a 1.7 mm path length cuvette under an inert Ar atmosphere. Potentials were referenced using an external ferrocene standard. Scans were run at 23 °C using Bu4NPF6 (0.1 M) as the supporting electrolyte.
Reaction of Mn(OH)(ttppc) (3) with Triphenylmethyl Radical. Product Analysis.
To a solution of 3 in benzene-d6 (1 mM, 1 mL) were added Ph2C=C6H4–CPh3 (1.5 mg, 6.2 μmol, 6 equiv) and the trimethylphenylsilane (TMPS) (1.0 mM) as an internal standard. The solution was stirred for 2 h and loaded into an NMR tube. A 1H NMR spectrum was collected, and a peak at 2.28 ppm assigned to the OH proton of Ph3COH was integrated and compared with the peak from the TMPS standard. An average yield for Ph3COH of 90 ± 1% was found (average of 3 runs). This reaction was repeated in nondeuterated solvent by the addition of Ph2C=C6H4–CPh3 (0.15 mM, 6 equiv) to a solution of Mn(OH)-(ttppc) (25 μM, λmax = 420 nm) in toluene. The inorganic product, MnIII(ttppc) λmax = 445, 660 nm) was quantified by the extinction coefficient (2.7 × 103 M−1 cm−1) at 660 nm. An average yield of MnIII(ttppc) of 88 ± 1% was found (average of 3 runs).
Reaction of Mn(OH)(ttppc) with 4-tBu-C6H4)3C•. Product Analysis.
To a solution of (4-tBu-Ph)3CBr (6.4 mM, 2 mL) in toluene-d8 was added excess Cu powder (~200 mg), and the reaction was stirred in the dark for 4 h at 23 °C. A color change to yellow was observed, indicating formation of (4-tBu-C6H4)3C•. The solution was filtered through Celite to remove solid Cu, and an aliquot was obtained to determine the concentration by UV–vis spectroscopy (5.3 mM, 83% yield). The freshly prepared radical solution (500 μL) was then added to 3 in toluene-d8 (500 μL) to give a final reaction mixture of 3 (2.35 mM) and (4-tBu-C6H4)3C• (2.65 mM, 1.1 equiv). The internal standard trimethylphenylsilane (TMPS) (8 mM) was added, and the solution was stirred for 24 h and then loaded into an NMR tube. A 1H NMR spectrum was collected, and a singlet at 1.21 ppm was observed corresponding to the tBu groups of the alcohol (4-tBu-C6H4)3COH. This peak was integrated and compared to the TMPS standard, giving a yield of 90 ± 1% (average yield of 3 runs).
Kinetics.
In an N2-filled glovebox under low light conditions, the radical derivatives (4-X-C6H4)3C• (X = OMe, tBu, Ph, Cl, or CN) were freshly prepared in toluene. Radical concentrations were determined by UV–vis spectroscopy. Varying amounts of radical stock solution (50–1000 μL) were immediately added to a solution of the metal complex in toluene (18 to 40 μM, 1.00 to 1.95 mL) to start the reaction. The spectral changes showed isosbestic conversion of Mn(OH)(ttppc) λmax = 420 nm) to MnIII(ttppc) (λmax = 445, 660 nm) for the manganese system and Fe(OH)(ttppc) (λmax = 405 nm) to FeIII(ttppc) λmax = 420, 575, 765 nm) for the Fe system. The pseudo-first-order rate constants, kobs, for these reactions were obtained from plots of absorbance versus time at the wavelengths 420 or 660 nm for the Mn complex, 465 nm for the Fe complex with the p-CN radical derivative, and 575 nm for the Fe complex with the p-tBu radical derivative. These plots were fit to the equation Abst = Absf + (Abs0 – Absf) exp(–kobst), where Abst is the absorbance at time t, kobs is the pseudo-first-order rate constant, and Abs0 and Absf are the initial and final absorbance values, respectively. Second order rate constants (k2) were obtained from the slope of the best-fit line from a plot of kobs versus radical concentration.
Computational Modeling.
Calculations were done using density functional theory methods as implemented in Gaussian-09,72 and methods and procedures were followed as reported and tested previously on analogous systems.57,61 Calculations were carried out on the complete metallocorrole structures without any truncation of the peripheral ligand substituents. Reaction profiles were calculated for the reaction of 1 or 3 with (4-X-C6H4)3C• (X = H, tBu, OMe, Cl, Ph, CN). The computed structures ranged in size from 191 atoms for X = H/Cl to 221 atoms for X = Ph.
The unrestricted B3LYP hybrid density functional theory method was used for geometry optimizations, constraint geometry scans, and frequency calculations.73,74 ' Due to the size of the chemical system, all structures were optimized without constraints with an LanL2DZ basis set on iron/manganese (with core potential) and 6-31G on the rest of the atoms (H, C, N, O, Cl): basis set BS1.75,76 Single points using a triple-ζ quality basis set on iron/manganese (with core potential), i.e., LACV3P+, and 6-311+G* on the rest of the atoms were done to correct the energies: basis set BS2. Geometries were optimized with solvent corrections included through the conductor-like polarized continuum model (CPCM) with a dielectric constant mimicking toluene.77 All free energy calculations were done at 298 K. Test calculations with dispersion-corrected DFT reproduced the trends well (Supporting Information); hence dispersion is not critical here.
Supplementary Material
ACKNOWLEDGMENTS
The authors would like to thank the NIH (GM101153) (D.P.G.) for the financial support of this research. J.G.A. is thankful for a Johns Hopkins University Greer Undergraduate Research Award. M.Q.E.M. thanks the Government of Malaysia for a studentship.
Footnotes
Complete contact information is available at: https://pubs.acs.org/10.1021/acs.inorgchem.0c02640
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c02640.
Kinetics data, additional spectro-electrochemical data, DFT optimized structures, calculation of C–O bond energies from DFT calculations, computational tables with group spin densities, charges and absolute and relative energies of all structures, Cartesian coordinates of all optimized geometries (PDF)
The authors declare no competing financial interest.
Contributor Information
Daniel C. Cummins, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States
Jessica G. Alvarado, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States
Jan Paulo T. Zaragoza, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States.
Muhammad Qadri Effendy Mubarak, Manchester Institute of Biotechnology and Department of Chemical Engineering and Analytical Science, The University of Manchester, Manchester M1 7DN, United Kingdom.
Yen-Ting Lin, Manchester Institute of Biotechnology and Department of Chemical Engineering and Analytical Science, The University of Manchester, Manchester M1 7DN, United Kingdom.
Sam P. de Visser, Manchester Institute of Biotechnology and Department of Chemical Engineering and Analytical Science, The University of Manchester, Manchester M1 7DN, United Kingdom.
David P. Goldberg, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States.
REFERENCES
- (1).Ortiz de Montellano PR Hydrocarbon Hydroxylation by Cytochrome P450 Enzymes. Chem. Rev 2010, 110, 932–948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (2).Poulos TL Heme Enzyme Structure and Function. Chem. Rev 2014, 114, 3919–3962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Rittle J; Green MT Cytochrome P450 Compound I: Capture, Characterization, and C-H Bond Activation Kinetics. Science 2010, 330, 933. [DOI] [PubMed] [Google Scholar]
- (4).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]
- (5).Denisov IG; Makris TM; Sligar SG; Schlichting I Structure and Chemistry of Cytochrome P450. Chem. Rev 2005, 105, 2253–2278. [DOI] [PubMed] [Google Scholar]
- (6).Makris TM; von Koenig K; Schlichting I; Sligar SG The status of high-valent metal oxo complexes in the P450 cytochromes. J. Inorg. Biochem 2006, 100, 507–518. [DOI] [PubMed] [Google Scholar]
- (7).Sacramento JJD; Goldberg DP Factors Affecting Hydrogen Atom Transfer Reactivity of Metal–Oxo Porphyrinoid Complexes. Acc. Chem. Res 2018, 51, 2641–2652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Yosca TH; Ledray AP; Ngo J; Green MT A new look at the role of thiolate ligation in cytochrome P450. JBIC, J. Biol. Inorg. Chem 2017, 22, 209–220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Shaik S; Kumar D; de Visser SP; Altun A; Thiel W Theoretical Perspective on the Structure and Mechanism of Cytochrome P450 Enzymes. Chem. Rev 2005, 105, 2279–2328. [DOI] [PubMed] [Google Scholar]
- (10).Li D; Wang Y; Han K Recent density functional theory model calculations of drug metabolism by cytochrome P450. Coord. Chem. Rev 2012, 256, 1137–1150. [Google Scholar]
- (11).Blomberg MRA; Borowski T; Himo F; Liao R-Z; Siegbahn PEM Quantum Chemical Studies of Mechanisms for Metalloenzymes. Chem. Rev 2014, 114, 3601–3658. [DOI] [PubMed] [Google Scholar]
- (12).Shaik S; Cohen S; de Visser SP; Sharma PK; Kumar D; Kozuch S; Ogliaro F; Danovich D The “Rebound Controversy”: An Overview and Theoretical Modeling of the Rebound Step in C-H Hydroxylation by Cytochrome P450. Eur. J. Inorg. Chem 2004, 2004, 207–226. [Google Scholar]
- (13).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]
- (14).Zaragoza JPT; Goldberg DP, Dioxygen Binding and Activation Mediated by Transition Metal Porphyrinoid Complexes. In Dioxygen-dependent Heme Enzymes; The Royal Society of Chemistry: 2019; pp 1–36. [Google Scholar]
- (15).Zaragoza JPT; Yosca TH; Siegler MA; Moënne-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]
- (16).Pangia TM; Yadav V; Gérard EF; Lin Y-T; de Visser SP; Jameson GNL; Goldberg DP Mechanistic Investigation of Oxygen Rebound in a Mononuclear Nonheme Iron Complex. Inorg. Chem 2019, 58, 9557–9561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).Yadav V; Rodriguez RJ; Siegler MA; Goldberg DP Determining the Inherent Selectivity for Carbon Radical Hydroxylation versus Halogenation with FeIII(OH)(X) Complexes: Relevance to the Rebound Step in Non-heme Iron Halogenases. J. Am. Chem. Soc 2020, 142, 7259–7264. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).Yadav V; Gordon JB; Siegler MA; Goldberg DP Dioxygen-Derived Nonheme Mononuclear FeIII (OH) Complex and Its Reactivity with Carbon Radicals. J. Am. Chem. Soc 2019, 141, 10148–10153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Pangia TM; Davies CG; Prendergast JR; Gordon JB; Siegler MA; Jameson GNL; Goldberg DP Observation of Radical Rebound in a Mononuclear Nonheme Iron Model Complex. J. Am. Chem. Soc 2018, 140, 4191–4194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).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]
- (21).Kleinlein C; Bendelsmith AJ; Zheng S-L; Betley TAC-H Activation from Iron(II)-Nitroxido Complexes. Angew. Chem., Int. Ed 2017, 56, 12197–12201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Bower JK; Cypcar AD; Henriquez B; Stieber SCE; Zhang S C(sp3)–H Fluorination with a Copper(II)/(III) Redox Couple. J. Am. Chem. Soc 2020, 142, 8514–8521. [DOI] [PubMed] [Google Scholar]
- (23).Bakhoda AG; Wiese S; Greene C; Figula BC; Bertke JA; Warren TH Radical Capture at Nickel(II) Complexes: C–C, C–N, and C–O Bond Formation. Organometallics 2020, 39, 1710–1718. [Google Scholar]
- (24).Guo M; Seo MS; Lee Y-M; Fukuzumi S; Nam W Highly Reactive Manganese(IV)-Oxo Porphyrins Showing Temperature-Dependent Reversed Electronic Effect in C–H Bond Activation Reactions. J. Am. Chem. Soc 2019, 141, 12187–12191. [DOI] [PubMed] [Google Scholar]
- (25).Liu W; Groves JT Manganese Porphyrins Catalyze Selective C-H Bond Halogenations. J. Am. Chem. Soc 2010, 132, 12847–12849. [DOI] [PubMed] [Google Scholar]
- (26).Liu W; Groves JT Manganese Catalyzed C–H Halogenation. Acc. Chem. Res 2015, 48, 1727–1735. [DOI] [PubMed] [Google Scholar]
- (27).Srour H; Maux PL; Simonneaux G Enantioselective Manganese-Porphyrin-Catalyzed Epoxidation and C–H Hydroxylation with Hydrogen Peroxide in Water/Methanol Solutions. Inorg. Chem 2012, 51, 5850–5856. [DOI] [PubMed] [Google Scholar]
- (28).Song WJ; Seo MS; George SD; Ohta T; Song R; Kang M-J; Tosha T; Kitagawa T; Solomon EI; Nam W Synthesis, characterization, and reactivities of manganese(V)-oxo porphyrin complexes. J. Am. Chem. Soc 2007, 129, 1268–1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Carrier M-N; Scheer C; Gouvine P; Bartoli J-F; Battioni P; Mansuy D Biomimetic hydroxylation of aromatic compounds: Hydrogen peroxide and manganese-polyhalogenated porphyrins as a particularly good system. Tetrahedron Lett. 1990, 31, 6645–6648. [Google Scholar]
- (30).Liu W; Cheng M-J; Nielsen RJ; Goddard WA; Groves JT Probing the C–O Bond-Formation Step in Metalloporphyrin-Catalyzed C–H Oxygenation Reactions. ACS Catal. 2017, 7, 4182–4188. [Google Scholar]
- (31).Rettie AE; Sheffels PR; Korzekwa KR; Gonzalez FJ; Philpot RM; Baillie TA CYP4 Isoenzyme Specificity and the Relationship between. omega.-Hydroxylation and Terminal Desaturation of Valproic Acid. Biochemistry 1995, 34, 7889–7895. [DOI] [PubMed] [Google Scholar]
- (32).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. Science 1987, 263, 13733–13738. [PubMed] [Google Scholar]
- (33).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]
- (34).Wise CE; Hsieh CH; Poplin NL; Makris TM Dioxygen Activation by the Biofuel-Generating Cytochrome P450 OleT. ACS Catal. 2018, 8, 9342–9352. [Google Scholar]
- (35).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]
- (36).Bharadwaj VS; Kim S; Guarnieri MT; Crowley MF Different Behaviors of a Substrate in P450 Decarboxylase and Hydroxylase Reveal Reactivity-Enabling Actors. Sci. Rep 2018, 8, 12826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (37).Pickl M; Kurakin S; Cantú Reinhard FG; Schmid P; Pöcheim 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]
- (38).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]
- (39).Gomberg M An Instance of Trivalent Carbon: Triphenylmethyl. J. Am. Chem. Soc 1900, 22, 757–771. [Google Scholar]
- (40).Jang ES; McMullin CL; Käßß M; Meyer K; Cundari TR; Warren TH Copper(II) Anilides in sp3 C-H Amination. J. Am. Chem. Soc 2014, 136, 10930–10940. [DOI] [PubMed] [Google Scholar]
- (41).Liu H-Y; Yam F; Xie Y-T; Li X-Y; Chang CK A Bulky Bis-Pocket Manganese(V)-Oxo Corrole Complex: Observation of Oxygen Atom Transfer between Triply Bonded MnV-O and Alkene. J. Am. Chem. Soc 2009, 131, 12890–12891. [DOI] [PubMed] [Google Scholar]
- (42).Dünnebacke D; Neumann WP; Penenory A; Stewen U Über sterisch gehinderte freie Radikale, XIX. Stabile 4,4′,4″-trisubstituierte Triphenylmethyl-Radikale. Chem. Ber 1989, 122, 533–535. [Google Scholar]
- (43).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]
- (44).Colclough N; Smith JRL A mechanistic study of the oxidation of phenols in aqueous solution by oxoiron(IV) tetra(N-methylpyridyl)porphyrins. A model for horseradish peroxidase compound II? J. Chem. Soc., Perkin Trans 2 1994, 1139–1149. [Google Scholar]
- (45).Volz H; Lotsch W Polarographische reduktion von triarylmethylkationen. Tetrahedron Lett. 1969, 10, 2275–2278. [Google Scholar]
- (46).Marcus RA; Sutin N Electron Transfer in Chemistry and Biology. Biochim. Biophys. Acta, Rev. Bioenerg 1985, 811, 811. [Google Scholar]
- (47).Osako T; Ohkubo K; Taki M; Tachi Y; Fukuzumi S; Itoh S Oxidation Mechanism of Phenols by Dicopper-Dioxygen (Cu2/O2) Complexes. J. Am. Chem. Soc 2003, 125, 11027–11033. [DOI] [PubMed] [Google Scholar]
- (48).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]
- (49).Andrews LJ; Keefer RM A Kinetic Study of the Diels-Alder Reaction of Various Anthracene and Maleic Anhydride Derivatives. J. Am. Chem. Soc 1955, 77, 6284–6289. [Google Scholar]
- (50).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]
- (51).Takahashi A; Kurahashi T; Fujii H Activation Parameters for Cyclohexene Oxygenation by an Oxoiron(IV) Porphyrin π-Cation Radical Complex: Entropy Control of an Allylic Hydroxylation Reaction. Inorg. Chem 2007, 46, 6227–6229. [DOI] [PubMed] [Google Scholar]
- (52).Moss RA; Lawrynowicz W; Turro NJ; Gould IR; Cha Y Activation parameters for the additions of arylhalocarbenes to alkenes. J. Am. Chem. Soc 1986, 108, 7028–7032. [Google Scholar]
- (53).Houk KN; Rondan NG Origin of negative activation energies and entropy control of halocarbene cycloadditions and related fast reactions. J. Am. Chem. Soc 1984, 106, 4293–4294. [Google Scholar]
- (54).Sobek J; Martschke R; Fischer H Entropy Control of the Cross-Reaction between Carbon-Centered and Nitroxide Radicals. J. Am. Chem. Soc 2001, 123, 2849–2857. [DOI] [PubMed] [Google Scholar]
- (55).Inoue Y; Ikeda H; Kaneda M; Sumimura T; Everitt SRL; Wada T Entropy-Controlled Asymmetric Photochemistry: Switching of Product Chirality by Solvent. J. Am. Chem. Soc 2000, 122, 406–407. [Google Scholar]
- (56).Hatano S; Abe J Activation Parameters for the Recombination Reaction of Intramolecular Radical Pairs Generated from the Radical Diffusion-Inhibited HABI Derivative. J. Phys. Chem. A 2008, 112, 6098–6103. [DOI] [PubMed] [Google Scholar]
- (57).Zaragoza JPT; Cummins DC; Mubarak MQE; Siegler MA; de Visser SP; Goldberg DP Hydrogen Atom Abstraction by High-Valent Fe(OH) versus Mn(OH) Porphyrinoid Complexes: Mechanistic Insights from Experimental and Computational Studies. Inorg. Chem 2019, 58, 16761–16770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).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]
- (59).Ji L; Faponle AS; Quesne MG; Sainna MA; Zhang J; Franke A; Kumar D; van Eldik R; Liu W; de Visser SP Drug Metabolism by Cytochrome P450 Enzymes: What Distinguishes the Pathways Leading to Substrate Hydroxylation Over Desaturation? Chem. - Eur. J 2015, 21, 9083–9092. [DOI] [PubMed] [Google Scholar]
- (60).Cantú 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]
- (61).Cantú Reinhard FG; Sainna MA; Upadhyay P; Balan GA; Kumar D; Fornarini S; Crestoni ME; de Visser SP A Systematic Account on Aromatic Hydroxylation by a Cytochrome P450 Model Compound I: A Low-Pressure Mass Spectrometry and Computational Study. Chem. - Eur. J 2016, 22, 18608–18619. [DOI] [PubMed] [Google Scholar]
- (62).de Visser SP Second-Coordination Sphere Effects on Selectivity and Specificity of Heme and Nonheme Iron Enzymes. Chem. - Eur. J 2020, 26, 5308–5327. [DOI] [PubMed] [Google Scholar]
- (63).Ghafoor S; Mansha A; de Visser SP Selective Hydrogen Atom Abstraction from Dihydroflavonol by a Nonheme Iron Center Is the Key Step in the Enzymatic Flavonol Synthesis and Avoids Byproducts. J. Am. Chem. Soc 2019, 141, 20278–20292. [DOI] [PubMed] [Google Scholar]
- (64).Louka S; Barry SM; Heyes DJ; Mubarak MQE; Ali HS; Alkhalaf LM; Munro AW; Scrutton NS; Challis GL; de Visser SP The catalytic mechanism of aromatic nitration by cytochrome P450 TxtE: Involvement of a ferric-peroxynitrite intermediate. J. Am. Chem. Soc 2020, 142, 15764 accepted.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (65).Cantú Reinhard FG; Fornarini S; Crestoni ME; de Visser SP Hydrogen Atom vs. Hydride Transfer in Cytochrome P450 Oxidations: A Combined Mass Spectrometry and Computational Study. Eur. J. Inorg. Chem 2018, 2018, 1854–1865. [Google Scholar]
- (66).Quesne MG; Senthilnathan D; Singh D; Kumar D; Maldivi P; Sorokin AB; de Visser SP Origin of the Enhanced Reactivity of μ-Nitrido-Bridged Diiron(IV)-Oxo Porphyrinoid Complexes over Cytochrome P450 Compound I. ACS Catal. 2016, 6, 2230–2243. [Google Scholar]
- (67).Shaik S; Kumar D; de Visser SP A Valence Bond Modeling of Trends in Hydrogen Abstraction Barriers and Transition States of Hydroxylation Reactions Catalyzed by Cytochrome P450 Enzymes. J. Am. Chem. Soc 2008, 130, 10128–10140. [DOI] [PubMed] [Google Scholar]
- (68).Li G; Han A; Pulling ME; Estes DP; Norton JR Evidence for Formation of a Co–H Bond from (H2O)2Co(dmgBF2)2 under H2: Application to Radical Cyclizations. J. Am. Chem. Soc 2012, 134, 14662–14665. [DOI] [PubMed] [Google Scholar]
- (69).Saito S; Nakazono K; Takahashi E Template Synthesis of [2]Rotaxanes with Large Ring Components and Tris(biphenyl)-methyl Group as the Blocking Group. The Relationship between the Ring Size and the Stability of the Rotaxanes. J. Org. Chem 2006, 71, 7477–7480. [DOI] [PubMed] [Google Scholar]
- (70).Broser W; Niemeier HKW Über paramagnetische dimerenkomplexe des substituierten triphenylmethylradikals. Tetrahedron 1976, 32, 1183–1187. [Google Scholar]
- (71).Belzner J; Bunz U; Semmler K; Szeimies G; Opitz K; Schluter A-D Concerning the Synthesis of [1.1.1]Propellane. Chem. Ber 1989, 122, 397–398. [Google Scholar]
- (72).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H; Li X; Caricato M; Marenich AV; Bloino J; Janesko BG; Gomperts R; Mennucci B; Hratchian HP; Ortiz JV; Izmaylov AF; Sonnenberg JL; Williams Ding F; Lipparini F; Egidi F; Goings J; Peng B; Petrone A; Henderson T; Ranasinghe D; Zakrzewski VG; Gao J; Rega N; Zheng G; Liang W; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Throssell K; Montgomery JA Jr., Peralta JE; Ogliaro F; Bearpark MJ; Heyd JJ; Brothers EN; Kudin KN; Staroverov VN; Keith TA; Kobayashi R; Normand J; Raghavachari K; Rendell AP; Burant JC; Iyengar SS; Tomasi J; Cossi M; Millam JM; Klene M; Adamo C; Cammi R; Ochterski JW; Martin RL; Morokuma K; Farkas O; Foresman JB; Fox DJ Gaussian 16 Rev. C.01 Wallingford, CT, 2016. [Google Scholar]
- (73).Lee C; Yang W; Parr RG Development of the ColleSalvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- (74).Becke AD Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]
- (75).Hay PJ; Wadt WR Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys 1985, 82, 270–283. [Google Scholar]
- (76).Ditchfield R; Hehre WJ; Pople JA Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys 1971, 54, 724–728. [Google Scholar]
- (77).Tomasi J; Mennucci B; Cammi R Quantum Mechanical Continuum Solvation Models. Chem. Rev 2005, 105, 2999–3094. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
