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

Reactivity of a heterobinuclear heme–peroxo–Cu complex with para-substituted catechols shows a pKa-dependent change in mechanism

Sanjib Panda a, Suzanne M. Adam a, Hai Phan a, Patrick J. Rogler a, Pradip Kumar Hota a, Joshua R. Helms b, Brad S. Pierce b, Gayan B. Wijeratne *b and Kenneth D. Karlin *a
aDepartment of Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, USA. E-mail: karlin@jhu.edu
bDepartment of Chemistry & Biochemistry, The University of Alabama, Tuscaloosa, Alabama 35487, USA. E-mail: gwijeratne@ua.edu

Received 21st August 2024 , Accepted 19th December 2024

First published on 30th December 2024


Abstract

In biological systems, heme–copper oxidase (HCO) enzymes play a crucial role in the oxygen reduction reaction (ORR), where the pivotal O–O bond cleavage of the (heme)FeIII–peroxo–CuII intermediate is facilitated by active-site (peroxo core) hydrogen bonding followed by proton-coupled electron transfer (PCET) from a nearby (phenolic) tyrosine residue. A useful approach to comprehend the fundamental relationships among H-bonding/proton/H-atom donors and their abilities to induce O–O bond homolysis involves the investigation of synthetic, bioinspired model systems where the exogenous substrate properties (such as pKa and bond dissociation energy (BDE)) can be systematically altered. This report details the reactivity of a heme–peroxo–copper HCO model complex (LS-4DCHIm) toward a series of substituted catechol substrates that span a range of pKa and O–H bond BDE values, exhibiting different reaction mechanisms. Considering their interactions with the bridging peroxo ligand in LS-4DCHIm, the catechol substrates are importantly capable of one or two (i) H-bonds, (ii) proton transfers, and/or (iii) net H-atom transfers, thereby making them attractive, yet complex candidates for studying the redox chemistry of the metal-bound peroxide. A combination of spectroscopic studies and kinetic analysis implies that the suitable modulation of pKa and O–H bond BDE values of catechols result in either double proton transfer with the release of H2O2 or double PCET resulting in reductive O–O bond rupture. The distinguishing role of substrate properties in directing the mechanism and outcome of O2 protonation/reduction reactions is discussed in terms of designing O2-reduction catalysts based on biological inspiration.


Introduction

The controlled movement of protons and electrons in biological systems is of great fundamental and functional importance, as these chemical processes can occur in a variety of ways and are often the cornerstone of redox enzyme transformations. During the final step of cellular respiration, O–O bond reductive cleavage is effected at the heterobinuclear active site of heme–copper oxidases (HCOs), where two electrons from the heme a3 (FeII → FeIV[double bond, length as m-dash]O), one electron from the CuB site (CuI → CuII–OH) and a single proton-coupled electron transfer (PCET) process from a juxtaposed tyrosine (Tyr → Tyr˙) residue are involved.1–4 Due to the importance of understanding O2 ⇄ H2O redox interconversion in the context of biology, as well as for fuel cell applications or catalytic oxidations,5–12 the mechanistic subtleties of how protons and electrons transfer most efficiently to accomplish O–O bond reductive cleavage (i.e., concerted or sequentially) must be deciphered. In doing so, the acid strength (pKa), reduction potential and bond dissociation energy (BDE) of the relevant substrates should be balanced to achieve net H-atom donation to an O2-derived moiety. In thermodynamic terms, this can occur through various pathways under the broader scope of the proton-coupled electron transfer (PCET) umbrella, including PTET, ETPT or HAT (PT: proton transfer; ET: electron transfer; HAT: hydrogen atom transfer), which continues to attract significant interest from many research groups.13–22

As relevant for this report, which describes PCET-mediated O–O bond cleavage of a (heme)FeIII–peroxo–CuII(L) coordination complex, the heme–Cu binuclear center (BNC) in HCOs activates and reduces O2 by 4e (to H2O, Fig. 1).2,3 The most commonly studied HCOs are cytochrome c oxidases (CcOs), which receive reducing equivalents from cytochrome c via a dicopper cofactor, CuA, and a low-spin heme site (Fig. 1, blue pathway). Most importantly, due to the fast kinetics of the O–O bond homolysis step in CcO enzymatic cycle, the fundamental details regarding the properties of putative metal-peroxide species (IP, documented to form before O–O bond cleavage) remain obscured.23 Extensive investigations using synthetic, spectroscopic, and computational models have provided substantial evidence for the presence of such an intermediate.2,23–29 Moreover, synthetic accessibility of (heme)FeIII–peroxo–CuII complexes (resembling IP) has resulted in their use as CcO models, which suggest that the O–O bond rupture in CcO most likely proceeds through a PT followed by ET from the Tyr.2,23–26


image file: d4sc05623j-f1.tif
Fig. 1 The flow of electrons in two types of HCOs. Quinol oxidases utilise a protein solubilised ubiquinol-8 (structure shown in green) as a source of two electrons, whereas CcOs receive reducing equivalents from cytochrome c via a dicopper site, CuA. Both types of HCOs contain a low-spin heme cofactor, which shuttles electrons to the heme–copper active site where dioxygen is bound and reduced.

Another subfamily of HCOs, quinol oxidases (QOs), instead utilise a protein-solubilised ubiquinol (2,3-dimethoxy-5-methyl-6-polyprenyl-1,4-benzoquinol, Fig. 1, green) molecule as an electron source,30 where the two protons released following (PC)ET to the BNC are pumped to the intermembrane space (Fig. 1).31,32 In CcO and QO native enzymes, as well as in biomimetic systems from Collman33 and other research groups carrying out electrocatalytic O2-reduction of synthetic compounds,2,28 it has been shown that this process is controlled by the rate of electron flux to the active site, but also requires a network of H-bonding interactions to facilitate the PCET from a proximal phenolic moiety.34–36 Therefore, a fundamental understanding of the functional relationships between transition metal-O2 species and phenolic-type substrates remains to be of high importance.

Hydroquinones (para-dihydroxybenzenes) or catechols (ortho-dihydroxybenzenes) act as redox substrates in several biochemical/metalloenzyme reactions, most famously in quinol oxidase37–40 and catechol oxidase,41–43 respectively, where the reducing equivalents are directed toward O–O bond cleavage chemistry. In both of these cases, the hydroquinone or catechol undergoes 2-electron oxidation to form the corresponding quinone, most likely via a PCET-type process. Redox reactions of catechols in synthetic systems can be exceedingly complex to interpret because, depending on the catecholic ring substitution(s), a library of such substrates can span a wide range of pKa's, O–H BDEs, and ionization potentials (IP) (for the first and second OH group; Fig. 2). Other mechanistic routes include only proton or only electron donation (although the latter is unlikely because oxidations without accompanying proton transfers yield high-energy species),14 as well as “overoxidation” to produce a hydroxy-quinone or ring-opened diols (Fig. 2). The overall PCET square scheme for 4-substituted catechols is shown in Fig. 2 and importantly includes potential internal H-bonding in certain deprotonated formulations.


image file: d4sc05623j-f2.tif
Fig. 2 Square scheme and other oxidation pathways of o-catechols. The greyed cationic species are unlikely to be involved in PCET reactions due to their high-energy nature.4

The bio-inspired model complex utilised in this study, LS-4DCHIm ([(DCHIm)(F8)FeIII–(O22−)–CuII(DCHIm)4]+; F8 = tetrakis(2,6-difluorophenyl)porphyrinate; DCHIm = 1,5-dicyclohexylimidazole) (Fig. 3), contains a low-spin heme-FeIII bridged by a (trans-μ-1,2-O22−) ligand to a CuII ion,44 where the axial ligand on the heme-iron and the four additional Cu ligands are monodentate imidazole (DCHIm) donors. The copper(II) ion in LS-4DCHIm adopts a distorted square pyramidal geometry, with a DCHIm ligand occupying the axial position. Importantly, the independent nature of the monodentate ligands seemingly imparts a certain degree of flexibility in Cu–ligand geometry, allowing small rearrangements to accommodate substrates approaching the bridging peroxo to assist in activation and/or cleavage.45 Our lab has recently evaluated the reactivity of LS-4DCHIm toward separate H+/H-bond and electron sources,45 where a weak phenolic acid, 4-NO2-phenol, H-bonded to the copper-bound O-atom of the bridging peroxo moiety in LS-4DCHIm to generate an adduct with an activated O–O bond (UV-vis, rR, and density functional theory (DFT) evidence).45 Unlike the parent peroxo complex, the H-bonded adduct was observed to be reactive toward exogenous decamethylferrocene reductant (Fc*), and undergoes 2e reductive O–O bond cleavage.45 Therefore, the overall peroxo reduction process was essentially following a PT-ET mechanism which proceeded via a H-bonded adduct.


image file: d4sc05623j-f3.tif
Fig. 3 Reactions of the model heme–peroxo–copper complex (LS-4DCHIm) with substituted catechols reveal distinct mechanistic pathways. As dependent on the catechol substitution (with varying pKa and BDE), H+ transfer occurs when R = NO2, net H-atom transfer when R = CN, CF3, Cl2, Cl, and no reaction when R = H, CH3, OCH3.

In this report, we expand our understanding of O2 activation and reduction by heme–copper systems related to biological O2-reduction. Our overarching interests lie in gaining a deeper understanding of the details of PCET-type reactions involving reductive O–O cleavage of dioxygen. Thus, by utilizing catechols, which can potentially H-bond in this system and provide one or two protons and/or electrons, we can monitor the effects of subtle changes in pKa, BDE, IP and H-bonding ability on the reaction outcomes with the LS-4DCHIm peroxo model system. Therefore, we have tested the reactivity of LS-4DCHIm toward a series of substituted catechols that span a range of pKa and O–H BDE values (Fig. 3). This demonstrates how the pKa of the catechol influences the reaction mechanism, such as: (i) for catechols with strongly electron-withdrawing NO2 group, the reaction proceeds via metal–oxygen bond cleavage (releasing H2O2); (ii) in contrast, catechols with moderately withdrawing groups (CN, CF3, Cl2, and Cl) favour O–O bond rupture; and (iii) catechols without substitution or with strongly electron-donating groups (OCH3 and CH3) do not react at all.

The results reported herein, garnered from a variety of spectroscopic and analytical methods, provide insights into the favourable substrate properties that lead to O–O bond cleavage. They also highlight the fact that subtle changes in chemical properties can have significant impacts on the mechanism and outcome of protonation/reduction reactions. Kinetic investigations and determination of the fate of the peroxo O-atoms (i.e., giving H2O2 or water) following the catechol reactions allow for distinguishing between 2PT vs. PTET mechanistic proposals (Fig. 3). Relating the reaction outcomes to trends in thermodynamic parameters of the catechol substrates allows for semi-quantitation; these, we observe a pKa-dependent mechanism, rather than a BDE-dependent mechanism (i.e., a rate-limiting proton transfer process), which leads us to offer new insights into the favourable conditions for O–O bond cleavage in such an environment.

Results and discussion

Reaction of LS-4DCHIm with 4-NO2-catechol (2PT)

Perhaps the most suitable way to monitor reactions of the LS-4DCHIm complex is via UV-visible spectroscopy, which allows detection of even slight changes in the absorption profile at the heme Soret and Q-bands, as well as in the low energy features, and in some cases, allows for observation of organic reaction products. Addition of one equivalent of 4-NO2-catechol to a solution of LS-4DCHIm (λmax = 423, 537, 845 nm, black spectra in Fig. 4B and C) at −90 °C induces a fast reaction as monitored by UV-vis spectroscopy (∼60 s) to yield the product spectra shown in red in Fig. 4B and C. The absorbance changes include a Q-band shift to 542 nm and isosbestic conversion of the Soret band to 413 nm. In addition, we observe a rapid decrease in intensity of the low energy features associated with the peroxo-to-Fe charge transfer transitions for heme–peroxo–Cu complexes in a low-spin iron(III) environment, based on prior rR spectroscopy and DFT calculations (Fig. 4B and C).44,46 These observations suggest fragmentation of the heme–peroxo–Cu core structure and formation of the low-spin bis-imidazolyl F8FeIII(DCHIm)2 product (Fig. 4A), which has been separately generated and characterised (λmax = 413, 542 nm).45 Additionally, the EPR spectrum of the product mixture (Fig. S1) shows the individual oxidised metal complexes (Fig. 4A), CuII(DCHIm)4 and F8FeIII(DCHIm)2. Authentic samples of both of these species can be generated, therefore spectral addition using calibration curves allows for semi-quantitation; these FeIII and CuII products were both obtained in ∼75% yield (Fig. S1).
image file: d4sc05623j-f4.tif
Fig. 4 (A) Reaction scheme and (B) UV-vis spectral changes at the Q-band and low energy regions (0.1 mM, l = 1 cm) and the Soret band (C) (0.1 mM, l = 2 mm) which occur when 1 equiv. of 4-NO2-catechol is added to LS-4DCHIm (black spectra) to yield the products depicted in the top scheme (red spectra), including F8FeIII(DCHIm)2, CuII(DCHIm)4 and 4-NO2-catecholate (see ESI) (l = cuvette path length).

Indeed, the absorption features of the bis-imidazolyl heme product (i.e., F8FeIII(DCHIm)2; where the DCHIm ligands are coordinated trans-axially) dominate the UV-vis spectrum of the final product mixture; however, we are aware that different mechanisms can lead to the same heme product formation,45 often preventing direct UV-vis observation of organic reaction products (catecholate or quinone) due to their low molar absorptivity (<4000 M−1 cm−1) and positioning of λmax (400–440 nm, i.e., underneath the intense heme Soret band). Nevertheless, in this case, an additional increase in absorbance is observed at ∼440 nm (appearing as a shoulder on the heme Soret band, indicated by a red arrow in Fig. 4C) in the UV-visible spectrum, which can be attributed to the 4-NO2-catecholate species which has been generated separately (Fig. 4C, and S2). This absorbance increase corresponds to the formation of approximately one equivalent of 4-NO2-catecholate as determined by spectral matching to a 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture of the products: F8FeIII(DCHIm)2, CuII(DCHIm)4, and [4-NO2-catecholate]2− (Fig. S2). Therefore, the spectroscopic characterization of the products by UV-vis and EPR support the formation of the F8FeIII(DCHIm)2, CuII(DCHIm)4, and catecholate products depicted in Fig. 4A. Consistent with these results, the ESI(+)-MS of the reaction mixture (Fig. 5 and S3) exhibits a peak at m/z = 1276.57 corresponding to [F8FeIII(DCHIm)2]+. Additionally, the peak at 527.36 can be attributed to the stable reduced form of [CuII(DCHIm)4]2+, i.e., [CuI(DCHIm)2]+, which is always observed due to the inherent reducing environment within the ESI-MS experiment. Also, the peak at 679.37 may suggest the partial formation of [(DCHIm)2CuII(4-NO2-catecholate)]+, detected as [M−H]+ in the ESI-MS (the expected m/z for [M]+ is 680.38; see the ESI for our further explanation) (Fig. 5). In conjunction with the UV-vis spectrum of free 4-NO2-catecholate (vide supra), the presence of a CuII-catecholate moiety in the mass spectrum provides indirect evidence of catecholate formation as a by-product.


image file: d4sc05623j-f5.tif
Fig. 5 Experimental ESI(+)-MS (segment) of product mixture following the reaction of LS-4DCHIm with 4-NO2-catechol. Masses (m/z) at 527.36 ([Cu(DCHIm)2]+), 679.37 ([Cu(DCHIm)2(4-NO2-catecholate)]+, detected as [M–H]+), and 1276.57 ([F8Fe(DCHIm)2]+).

As further support for this conclusion, the established horseradish peroxidase (HRP) enzyme assay45,47 allows for quantification of the amount of H2O2 released in the reaction shown in Fig. 4A. This method shows that addition of one equivalent of 4-NO2-catechol to the LS-peroxo complex results in the evolution of a stoichiometric amount of H2O2 (100% yield, Table 1), implying that this acidic catechol acts as a 2H+-transfer reagent (2PT mechanism; pKa(1)[NO2-catechol] = 11.9, pKa(2)[NO2-monocatecholate]1− = 31.4, vide infra). Interestingly, upon addition of 2 equiv. strong acid [DMF·H+](CF3SO3) to the reaction product mixture, we observe the disappearance of the absorption feature (λmax = 440 nm, vide supra) attributed to the catecholate, therefore, we believe that the stronger acid re-protonates the catecholate (Fig. S4). Additionally, the generation of H2O2 in the {LS-4DCHIm + 4-NO2-catechol} reaction mixture has been supported by an iodide test (see Fig. S5).

Table 1 Quantification of H2O2 evolved from reactions of LS-4DCHIm with catechols via the horseradish peroxidase assay
Substrate H2O2 evolveda (%)
a Quantification of H2O2 was achieved by recording the intensity of the diammonium 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonate) (AzBTS-(NH4)2) peaks (see ESI).
4-NO2-catechol 100
4-CN-catechol 6
4-CF3-catechol 2
4,5-Cl2-catechol 5
4-Cl-catechol 6


It has previously been shown that, when added to various metal peroxo species, acids with sufficiently low pKa's cause release of H2O2 following acid–base metal–oxygen bond cleavage.45,48 Accordingly, a catechol in which the first and second pKa values are low enough (i.e., lower than the pKa of the conjugate acids of Fe–(O22−)–Cu and Fe⋯(HO2)⋯Cu, respectively), can act as a 2H+-donor, without providing any exogenous reducing equivalents. To this end, we conclude that 4-NO2-catechol, which is relatively acidic but difficult to oxidise (O–H BDE = 78.3 and 73.4), reacts as a 2-proton donor, cleaving metal–O bonds via acid–base chemistry and releasing H2O2, an undesirable ROS (Scheme 1).


image file: d4sc05623j-s1.tif
Scheme 1 Stepwise transfer of two protons from 4-NO2-catechol to LS-4DCHIm, cleaving the metal–O bonds and releasing H2O2 without the involvement of any reducing equivalents.

Reaction of LS-4DCHIm with 4-substituted –CN, –CF3, –Cl2, and –Cl catechols (2PTET)

The addition of one equivalent of catechols featuring moderately to slightly electron-withdrawing substituents (–CN, –CF3, –Cl2, and –Cl catechols) to LS-4DCHIm at −90 °C likewise induces a reaction. Isosbestic conversion of the Soret band from 423 to 413 nm is accompanied by a shift in the Q-band from 537 to 542 nm, and the low energy features of LS-4DCHIm disappear (Fig. 6 and S9). As in the case of 4-NO2-catechol, the final UV-vis spectrum is consistent with the formation of the heme species, F8FeIII(DCHIm)2 (λmax = 413, 542 nm, Fig. 6B, C and S9)45 in high yield. EPR spectroscopy (Fig. S6) further supports a product mixture comprising the heme FeIII and a CuII complex in a ∼1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio. In support of this conclusion, we measured the final Cu(II) concentration in the {LS-4DCHIm + 4-Cl-catechol} reaction mixture; the results confirm the quantitative formation of Cu(II) and also reveal a usual axial EPR pattern with hyperfine splitting, which is consistent with the simulated spectra (Fig. S6b). Overall, the detection of the oxidised metal centers in the EPR product spectra and the disappearance of low-energy features in the UV-vis spectra collectively indicate that the catechol reactions result in the rupture of the bridging peroxo unit. Moreover, analogous to the 4-NO2-catechol case, the ESI(+)-MS (Fig. S7) of the reaction mixtures of LS-4DCHIm with 4-Cl-catechol shows peaks at 1276.56 and 527.36, corresponding to [F8FeIII(DCHIm)2]+ and [CuI(DCHIm)2]+ (i.e., stable reduced form observed instead of [CuII(DCHIm)4]2+), respectively.
image file: d4sc05623j-f6.tif
Fig. 6 (A) Reaction scheme and (B) UV-vis spectral changes at the Q-band and low energy regions (0.1 mM, l = 1 cm) and the Soret band (C) (0.1 mM, l = 2 mm), which occur when 1 equiv. of the catechols shown in the scheme (R = –CN, –CF3, –Cl2, –Cl) are added to LS-4DCHIm (black spectra) to yield the products depicted in the top scheme (blue spectra), characteristic of the F8FeIII(DCHIm)2 product.

However, in contrast to the reactivity observed with 4-NO2-catechol, addition of one equivalent of these catechols evolves negligible amounts of H2O2 according to the HRP test (2–6% of one equivalent, Table 1). This distinction suggests that subsequent proton transfer(s) followed by electron transfer(s) (PTET) from catechol moieties lead to the reductive cleavage of the O–O core. This may proceed via the initial formation of an adduct where the peroxo unit is most likely H-bonded to the catechol (vide infra).45 Only one mol-equiv. of catechol is required to complete the reactions, thus we propose that one equivalent of o-quinone derived from the catechol substrate (Fig. 6A) is also produced, unlike the case of 4-NO2-catechol (where the catecholate is generated). The reaction stoichiometry therefore dictates that two protons and two electrons have transferred, and assuming both of the protons and electrons go to the O-atoms of the peroxo moiety, we expect that 2 moles of hydroxide (OH) are formed (Fig. 6A).

Methods aimed to detect the formation of H2O/OH have so far been unsuccessful; however, ESI(−)-MS of the reaction mixture {LS-DCHIM + 4-Cl-catechol} shows the formation of o-quinone product (Fig. S8). Additionally, the fact that exactly one equiv. of catechol is required to complete the reaction supports the proposed reaction stoichiometry. Considering the results of the HRP assay and the lack of reactivity with electron-rich catechols having weak O–H BDEs, we hypothesise that although a net two H-atom transfer reaction has occurred, the mechanism most likely incurs a series of catechol-initiated PTET events, including a rate-limiting PT step, rather than two concerted HAT processes. It has been shown in several metal-peroxide systems, including our own work with LS-4DCHIm, that while strong acids promote release of H2O2via M–O cleavage (Scheme 1), relatively weaker acids more favourably achieve reductive O–O cleavage.45,48 Comprehensive kinetic analyses for the reactions of the LS-4DCHIm peroxo complex (vide infra) provide additional mechanistic insights.

Mechanistic insights based on kinetic analyses and DFT

To further probe the reaction mechanism(s), pseudo-first order kinetics experiments were employed using increasing concentrations of the catechol substrates described above (see ESI for details and Fig. S10 for all kinetic plots). In all cases, rate saturation occurs at high catechol concentrations (>10 equiv. for all cases except 4-NO2-catechol, which reaches a maximum rate with only ∼5 equiv. added) (Fig. 7, top). This type of saturation behaviour is consistent with a mechanism in which rapid equilibrium formation of an intermediate (Keq.) precedes the rate determining step (Fig. 7, middle), and the relationship between the observed rate constant, kobs, and the physical parameters involved can be described by the equation in Fig. 7, bottom.45,49–54
image file: d4sc05623j-f7.tif
Fig. 7 (Top) kobsvs. [catechol] plots for two representative catechols (see Fig. S10 for all other plots) exemplifying the saturation behaviour at high catechol concentrations. (Middle) Scheme showing the kinetic reaction model employed. (Bottom) The equation used to fit the kobsvs. [catechol] data.

The Keq. and k1 values shown in Table 2 were determined from kobsvs. [catechol] plots (Fig. 7, top, and S10). The fit parameters support the existence of two distinct mechanisms based on the acidity of the catecholic substrate (vide infra). Although the rapid reactions prevent isolation or spectroscopic characterization of an intermediate in these cases, we propose that the fast equilibrium step corresponds to the formation of a transient [LS-4DCHIm⋯(catechol)] H-bonded adduct. This proposal is supported by results from our previous studies, including kinetics, rR spectroscopy, and DFT calculations, which demonstrated that LS-4DCHIm can physically accommodate small substrates capable of H-bonding (e.g., 4-NO2-phenol).45 Additionally, H-bonding to metal-peroxo moieties is known to be “activating” (i.e., it induces weakening of the O–O bond and strengthening of the Fe–O bond, as inferred from rR spectroscopic analysis, thereby reducing the barrier for O–O cleavage).45,54–57 In this study, the catechol binding to the LS-4DCHIm complex, likely via H-bonding to the peroxo moiety, would effectively initiate (i) a PTET cascade and O–O reductive cleavage (R = –CN, –CF3, –Cl2, –Cl) or (ii) PT and release of H2O2 (R = –NO2), depending on substrate acidity. Thus, the formation of the proposed H-bonded precursor complex prior to the rate-limiting step of the reaction (vide infra) would result in the observed saturation behaviour for the overall reaction kinetics (Fig. 7). Interestingly, the catechol pKa values do not correlate linearly with the fitted Keq. values, which is evidence against a full proton transfer during the initial equilibrium, but rather provides further support for initial formation of an intermediate resembling the H-bonded LS-4DCHIm(catechol) adduct (see Fig. 7).

Table 2 Kinetic parameters calculated from fitting kobsvs. [catechol] data plots shown in Fig. 7 (top) with the equation in Fig. 7 (bottom)
Substrate K eq. (M−1) k 1 (M−1 s−1)
4-NO2-catechol 13[thin space (1/6-em)]900 ± 1500 0.0535 ± 0.0016
4-CN-catechol 4500 ± 300 0.0679 ± 0.0027
4-CF3-catechol 1900 ± 100 0.0538 ± 0.0017
4,5-Cl2-catechol 4000 ± 400 0.0531 ± 0.0011
4-Cl-catechol 2100 ± 200 0.0291 ± 0.0011


To better understand the fundamental relationships between the reaction mechanisms (kinetics) and reactant properties (thermodynamics) and to gain insights into what factors lead to efficient, metal-ion mediated O–O reductive cleavage, it is necessary to evaluate the reactivity outcomes in terms of catechol pKa, O–H BDE, and ionization potential (IP). For the sake of internal consistency (since these values have not been experimentally measured for the scope of catechols studied herein under the relevant conditions, i.e., solvent and temperature), we report and compare our own DFT-calculated parameters (pKa, BDE, and IP for the first and second proton, H-atom, and electron, respectively) (see ESI for details),58–60 to gain a sense of relative trends across the scope of catechols employed. The computationally determined parameters are given in Tables 3 and S1.

Table 3 Thermodynamic parameters for the catechols used in this study, calculated using DFT with a THF solvent modela
Substrate O–H BDE 1 (2) kcal mol−1 pKa 1 (2) IP 1 kcal mol−1
a Shown are the O–H bond dissociation energies (BDE), pKa's, and ionization potentials (IP) for the first (and second) H-atom, H+, or e, respectively in THF. DFT calculations were performed using the B3LYP(SMD)/6-311++G(2df,p)//B3LYP(SMD)/6-31+G(d) level of theory (see ESI).
image file: d4sc05623j-u1.tif 74.2 (69.3) 11.9 (31.4) 151.6
image file: d4sc05623j-u2.tif 73.6 (68.7) 15.5 (33.5) 147.9
image file: d4sc05623j-u3.tif 72.3 (66.7) 18.1 (35.0) 146.1
image file: d4sc05623j-u4.tif 70.5 (65.6) 18.3 (33.5) 142.4
image file: d4sc05623j-u5.tif 69.7 (65.5) 19.9 (36.3) 140.2
image file: d4sc05623j-u6.tif 70.4 (65.3) 23.0 (38.5) 138.1
image file: d4sc05623j-u7.tif 67.8 (64.1) 23.0 (40.8) 127.1
image file: d4sc05623j-u8.tif 64.1 (62.7) 23.7 (39.2) 133.3


Considering the rate-determining step, k1, in relation to the calculated thermodynamic parameters, O–H BDE and pKa for the first H-atom or proton transferred, respectively, a trend is clear. The Evans–Polanyi plot14,49,61–63 depicted in Fig. 8A and S11 shows an apparently linear relationship between the k1 rate constant and the first O–H BDE for the catechols excluding 4-NO2-catechol, which corroborates the spectroscopic evidence indicating that 4-NO2-catechol follows a different mechanistic pathway from the other four catechols.25 The small value of the slope of the linear best-fit trendline in the Evans–Polanyi plot (0.08) denotes a weak dependence of rate on the catechol O–H BDE and an “early” transition state for the reaction.19,62,64 This is also consistent with our proposal that the identity of the intermediate is a “reactant-like” H-bonded adduct (Fig. 7, bottom). Furthermore, the positive directionality of the correlation in Fig. 8A reflects the fact that the rate is the slowest for the catechol with the weakest O–H BDE. This is in contrast to the expected relationship between BDE and reaction rate for an HAT reaction, consistent with the observation that electron-rich catechols having very weak O–H BDE do not react with LS-4DCHIm to reductively cleave the bridging peroxide moiety. Since the rates of oxidation of catechols do not well correlate with their bond dissociation energies, we therefore presume that the rate limiting step is a protonation process (Fig. 7 and as discussed in the next paragraph), but not an O–H bond cleavage event which drives the reaction forward.


image file: d4sc05623j-f8.tif
Fig. 8 Evans–Polanyi plot (A) and relationship between reaction rate and first pKa (B), which show linear trends where the 4-NO2-catechol is a clear outlier. The data point for 4-NO2 catechol was excluded from the fit, as the reaction mechanism is different, see the discussion in the text. Also, see the ESI.

The relationship between the k1 rate constant and the pKa of the first proton is shown in Fig. 8B, and again, a trend is evident among all the catechols with the exception of 4-NO2-catechol. While this is consistent with 4-NO2-catechol following a different reaction mechanism, it is clear that the pKa is an important parameter for predicting the mechanistic outcomes of this system. Importantly, 4-Cl-catechol, which has the highest pKa (is the least acidic), also exhibits the slowest rate of reaction (Table 2). 4-CN-catechol having pKa(1) = 15.5 defines a threshold where the acidic nature is not strong enough to break the M–Operoxo bond, thus not allowing the production of free H2O2. Since the 4-CN-catechol is the most acidic of the catechols that yield 2H+/2e O–O reductive cleavage chemistry, and also the fastest reacting, this finding is in excellent agreement with a protonation (or H-bonding)-dependent reaction rate (and mechanism), where proton transfer is rate-determining, and the pKa of the catechol substrate is the reaction-determining parameter (2PTET and O–O cleavage: 15.5 ≤ pKa(1) ≤ 19.9 vs. 2PT and H2O2 release: pKa(1) ≤ 11.9). As mentioned, previous studies with LS-4DCHIm showed formation of an H-bonded adduct with p-NO2-phenol (pKa (THF, expt.) = 18);65 however, in that case, an external electron source was required for an overall PCET reaction to occur.45

In the reactions of weakly acidic catechols (which also have a sufficiently low O–H BDE), the catecholic substrate is presumed to provide the protons and, more importantly, the electrons necessary for reductive O–O bond cleavage to produce H2O. Thus, based on H2O2 quantification and kinetic analysis (supporting a rate limiting proton transfer process), we propose that the 4-NO2-catechol follows a fast 2PT mechanism, and the 4-substituted –CN, –CF3, –Cl2, and –Cl catechols follow a 2PTET mechanism. In the cases involving 2H+/2e reactivity (net 2H˙ transfer from the catechol to the peroxo complex, giving the respective o-quinone) and resultant O–O reductive cleavage, the catechol must be acidic enough to initiate the reaction, likely by activating the peroxo moiety via H-bonding,45,54 while also having sufficiently weak BDE(s). In the grand scheme of O2-activation and reduction (Fig. 9), these findings indicate that, for successful O2-reductive cleavage to occur within a bridging heme–peroxo–Cu construct, a PTET mechanism is preferred over an HAT mechanism, although the same quantity of protons and electrons are transferred in either case. This is consistent with previous findings that H-bonding and/or protonation of a metal-bridging peroxo moiety can activate the O–O bond and lower the barrier to reduction;45,54 both of these interactions are dependent on the H-bonding ability or pKa of the substrate.


image file: d4sc05623j-f9.tif
Fig. 9 Mechanistic pathways for reactions of LS-4DCHIm with catechols of varying pKa's.

It is important to note that, although parameters have been determined for both the first and second proton and/or electron, it is reasonable to assume that the first pKa, BDE, or IP determines the reaction pathway (Fig. 9), especially because the trends are consistent between first vs. second H+, H˙, and e (Table 3). Because O–O cleavage of a formal peroxo moiety only requires one electron (plus proton(s)), one might predict that catechols with weaker O–H BDEs or lower IPs will react most favourably to afford O–O cleavage products. If, in fact, only one electron is transferred (with or without proton(s)), the resultant products are expected to be a high-valent FeIV[double bond, length as m-dash]O and semiquinone; however, these are unlikely to persist or be observable, even under the cryogenic reaction conditions due to the highly reactive nature of the compound II-type species45 and the fact that the second O–H bond in the semiquinone is significantly weaker than the first (see Fig. 9 (center), 2, and Table 3).14,60,66 This is consistent with the proposed equilibrium formation of an H-bonded intermediate, which activates the O–O bond for cleavage; in this case, it also primes the proximal o-hydroxy group of the semiquinone to transfer its proton (and electron) in a rapid cascade.67 It is important to note that in HCO enzymes and in some model studies,54 the fourth and final electron required for complete dioxygen reduction has been shown to come from the iron center, thereby generating the ferryl species, FeIV[double bond, length as m-dash]O (single electron transfers from the Fe, Cu, and Tyr moieties provide the first three reducing equivalents). In the catechol reactions, which we propose result in O–O reductive cleavage, the rapid reaction kinetics prevent observation of the intermediate heme species. Therefore, we cannot determine whether both of the catecholic electrons go to the peroxo moiety (bypassing formation of a high-valent iron-oxo species) or, rather, if the two electrons required for peroxo-cleavage originate from the catechol (1e; catechol to semiquinone) and the heme-iron center (1e; FeIII to FeIV). In the latter case (depicted in Fig. 9), we presume that the second electron from the catechol would rapidly reduce the FeIV intermediate to the more stable FeIII product observed spectroscopically, resulting in the release of hydroxide. Furthermore, no reaction is observed with the less acidic R-catechols (R = H, Me, OMe), as indicated by the absence of changes in the UV-vis spectra in the Soret and Q-band regions (given in Fig. S12, ESI), and also leads to the possibility of H-bonded adduct formation ambiguous.68

Conclusions

In this report, we have described the reactions of a low-spin heme–peroxo–Cu complex with a series of substituted catechols at low temperature (−90 °C) in MeTHF, which interestingly span two different mechanisms based on the thermodynamic parameters of the catecholic substrates, namely their O–H BDE and pKa.

Product determination, H2O2 quantification, and kinetic evaluations of the reactions of LS-4DCHIm with weakly to highly acidic catechols in combination with DFT calculations have led us to propose that an initial pKa-dependent H-bond adduct formation is critical for activating the peroxide bridge and priming the substrate for the subsequent proton and electron transfers (Fig. 9, center pathway). Indeed, such H-bonded precursor complexes have been proposed to reduce the activation barrier of the overall reaction, making the proton/electron transfer energetically more favourable.32 However, a balance of pKa and O–H BDE is necessary to avoid double PT and release of H2O2, rather than oxidation of the catechol and effective O–O cleavage. Using a series of catechols with a range of pKa's and O–H BDEs, we have shown that the NO2-catechol participates in 2PTs and H2O2 release (Fig. 9, top pathway), whereas reactions of LS-4DCHIm with CN–, CF3–, Cl2–, and Cl-catechols release negligible H2O2 and lead to O–O reductive cleavage via a PTET mechanism.

Interestingly, the LS-4DCHIm peroxo complex is unreactive toward catechols with electron-donating substitution (i.e., non-acidic, with weak O–H bonds, such as those in 4-OCH3-catechol and 4-CH3-catechol) or unsubstituted catechol, even if added in excess (up to 10 equiv.). This result has mechanistic implications, suggesting that even if catechols are competent in providing two protons and two electrons (i.e., two net H-atoms) to effect O–O bond reductive cleavage, the mechanism of transfer is stepwise. In other words, catechols are unlikely to transfer protons and electrons in a concerted manner, and therefore, weak substrate O–H BDEs alone are not sufficient to activate and reduce the heme/Cu-bridged peroxide moiety (Fig. 9, bottom pathway). This study highlights the key balance of H+ and e transfer to govern the O–O cleavage versus M–O bond rupture and ROS release, which is relevant to understanding biochemical O2-reduction during cellular respiration and for the rational design of practical catalysts and fuel cell technologies. Additionally, this work could contribute to a broader understanding of enzymatic redox processes and provide further insights into the role of PCET in biological systems.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

K. D. K., S. M. A. and G. B. W. conceived the project idea and K. D. K. supervised the investigation; S. M. A., S. P., and G. B. W. synthesized all compounds, and generated and analyzed the experimental data with the help of other co-authors; S. P. performed the theoretical calculations and S. P., S. M. A., and G. B. W. interpreted the data. J. H. and B. S. P. carried out the EPR quantitation. K. D. K., S. M. A., S. P., and G. B. W. wrote and edited the manuscript with input from all other co-authors. All authors have approved the final version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This research was supported by the U.S. National Institutes of Health (NIH) under Awards R01GM60353 (K. D. K.) and R35GM139536 (K. D. K.).

Notes and references

  1. M. Wikström, K. Krab and V. Sharma, Chem. Rev., 2018, 118, 2469–2490 CrossRef.
  2. (a) S. M. Adam, G. B. Wijeratne, P. J. Rogler, D. E. Diaz, D. A. Quist, J. J. Liu and K. D. Karlin, Chem. Rev., 2018, 118, 10840–11022 CrossRef PubMed; (b) S. Panda, H. Phan and K. D. Karlin, J. Inorg. Biochem., 2023, 249, 112367 CrossRef PubMed; (c) S. Hematian, I. Garcia-Bosch and K. D. Karlin, Acc. Chem. Res., 2015, 48, 2462–2474 CrossRef.
  3. S. Yoshikawa and A. Shimada, Chem. Rev., 2015, 115, 1936–1989 CrossRef.
  4. P. E. M. Siegbahn and M. R. A. Blomberg, Chem. Rev., 2010, 110, 7040–7061 CrossRef PubMed.
  5. N. Mano and A. de Poulpiquet, Chem. Rev., 2018, 118, 2392–2468 CrossRef PubMed.
  6. M. L. Pegis, C. F. Wise, D. J. Martin and J. M. Mayer, Chem. Rev., 2018, 118, 2340–2391 CrossRef PubMed.
  7. E. S. Dy, T. A. Roman, Y. Kubota, K. Miyamoto and H. Kasai, J. Phys.:Condens. Matter, 2007, 19, 445010 CrossRef.
  8. J. A. Cracknell, K. A. Vincent and F. A. Armstrong, Chem. Rev., 2008, 108, 2439–2461 CrossRef PubMed.
  9. (a) T. Matsumoto, K. Kim and S. Ogo, Angew. Chem., Int. Ed., 2011, 50, 11202–11205 CrossRef PubMed; (b) D. S. Su and G. Sun, Angew. Chem., Int. Ed., 2011, 50, 11570–11572 CrossRef CAS PubMed.
  10. B. Zhang, S. Wang, W. Fan, W. Ma, Z. Liang, J. Shi, S. Liao and C. Li, Angew. Chem., Int. Ed., 2016, 55, 14748–14751 CrossRef.
  11. W. Schöfberger, F. Faschinger, S. Chattopadhyay, S. Bhakta, B. Mondal, J. A. A. W. Elemans, S. Müllegger, S. Tebi, R. Koch, F. Klappenberger, D.-C. M. Paszkiewicz, J. V. Barth, E. Rauls, H. Aldahhak, W. G. Schmidt and A. Dey, Angew. Chem., Int. Ed., 2016, 55, 2350–2355 CrossRef.
  12. W. Zhang, W. Lai and R. Cao, Chem. Rev., 2017, 117, 3717–3797 CrossRef PubMed.
  13. (a) Proton-coupled electron transfers can be categorized by multiple steps: PTET = proton transfer followed by electron transfer, ETPT = electron transfer followed by proton transfer, or a single step: CPET = concerted proton-electron transfer where the proton and electron may transfer to different entities. Hydrogen atom transfer (HAT) involves a concerted transfer of an H+/e pair (H˙) to the same orbital.; (b) J. P. Layfield and S. Hammes-Schiffer, Chem. Rev., 2014, 114, 3466–3494 CrossRef PubMed.
  14. (a) J. J. Warren, T. A. Tronic and J. M. Mayer, Chem. Rev., 2010, 110, 6961–7001 CrossRef PubMed; (b) R. G. Agarwal, C. F. Wise, J. J. Warren and J. M. Mayer, Chem. Rev., 2022, 122, 1482 CrossRef PubMed; (c) R. G. Agarwal, S. C. Coste, B. D. Groff, A. M. Heuer, H. Noh, G. A. Parada, C. F. Wise, E. M. Nichols, J. J. Warren and J. M. Mayer, Chem. Rev., 2022, 122, 1–49 CrossRef; (d) M. Bergner, S. Dechert, S. Demeshko, C. Kupper, J. M. Mayer and F. Meyer, J. Am. Chem. Soc., 2017, 139, 701–707 CrossRef; (e) M. A. Bowring, L. R. Bradshaw, G. A. Parada, T. P. Pollock, R. J. Fernández-Terán, S. S. Kolmar, B. Q. Mercado, C. W. Schlenker, D. R. Gamelin and J. M. Mayer, J. Am. Chem. Soc., 2018, 140, 7449–7452 CrossRef PubMed; (f) J. M. Mayer, Acc. Chem. Res., 2011, 44, 36–46 CrossRef; (g) W. D. Morris and J. M. Mayer, J. Am. Chem. Soc., 2017, 139, 10312–10319 CrossRef PubMed.
  15. D. R. Weinberg, C. J. Gagliardi, J. F. Hull, C. F. Murphy, C. A. Kent, B. C. Westlake, A. Paul, D. H. Ess, D. G. McCafferty and T. J. Meyer, Chem. Rev., 2012, 112, 4016–4093 CrossRef.
  16. (a) J. Rosenthal and D. G. Nocera, Acc. Chem. Res., 2007, 40, 543–553 CrossRef PubMed; (b) R. I. Cukier and D. G. Nocera, Annu. Rev. Phys. Chem., 1998, 49, 337–369 CrossRef; (c) C. J. Chang, L. L. Chng and D. G. Nocera, J. Am. Chem. Soc., 2003, 125, 1866–1876 CrossRef.
  17. S. Hammes-Schiffer and A. A. Stuchebrukhov, Chem. Rev., 2010, 110, 6939–6960 CrossRef.
  18. E. C. Gentry and R. R. Knowles, Acc. Chem. Res., 2016, 49, 1546–1556 CrossRef.
  19. A. Migliore, N. F. Polizzi, M. J. Therien and D. N. Beratan, Chem. Rev., 2014, 114, 3381–3465 CrossRef.
  20. C. Costentin, M. Robert, J.-M. Savéant and C. Tard, Acc. Chem. Res., 2014, 47, 271–280 CrossRef PubMed.
  21. D. Usharani, D. Janardanan, C. Li and S. Shaik, Acc. Chem. Res., 2013, 46, 471–482 CrossRef PubMed.
  22. A. Dey, J. Dana, S. Aute, P. Maity, A. Das and H. N. Ghosh, Chem.–Eur. J., 2017, 23, 3455–3465 CrossRef.
  23. (a) F. Poiana, C. von Ballmoos, N. Gonska, M. R. A. Blomberg, P. Ädelroth and P. Brzezinski, Sci. Adv., 2017, 3, e1700279 CrossRef PubMed; (b) M. R. A. Blomberg, Chem. Soc. Rev., 2020, 49, 7301–7330 RSC; (c) M. R. A. Blomberg, Biochem., 2016, 55, 489–500 CrossRef.
  24. (a) A. W. Schaefer Jr, A. C. Roveda, A. Jose and E. I. Solomon, J. Am. Chem. Soc., 2019, 141, 10068–10081 CrossRef; (b) A. Jose, A. W. Schaefer Jr, A. C. Roveda, W. J. Transue, S. K. Choi, Z. Ding, R. B. Gennis and E. I. Solomon, Science, 2021, 373, 1225–1229 CrossRef; (c) Z. Halime, M. T. Kieber-Emmons, M. F. Qayyum, B. Mondal, T. Gandhi, S. C. Puiu, E. E. Chufan, A. A. N. Sarjeant, K. O. Hodgson, B. Hedman, E. I. Solomon and K. D. Karlin, Inorg. Chem., 2010, 49, 3629–3645 CrossRef CAS PubMed; (d) M. A. Ehudin, A. W. Schaefer, S. M. Adam, D. A. Quist, D. E. Diaz, J. A. Tang, E. I. Solomon and K. D. Karlin, Chem. Sci., 2019, 10, 2893–2905 RSC.
  25. (a) L. Noodleman, W.-G. H. Du, J. A. Fee, A. W. Götz and R. C. Walker, Inorg. Chem., 2014, 53, 6458–6472 CrossRef CAS; (b) W.-G. H. Du, A. W. Götz and L. Noodleman, Inorg. Chem., 2019, 58, 13933–13944 CrossRef CAS; (c) L. Noodleman, A. W. Götz, W.-G. H. Du and L. Hunsicker-Wang, Front. Chem., 2023, 11, 1186022,  DOI:10.3389/fchem.2023.1186022.
  26. (a) A. Shimada, T. Tsukihara and S. Yoshikawa, Front. Chem., 2023, 11, 1108190,  DOI:10.3389/fchem.2023.1108190; (b) I. Ishigami, S. Russi, A. Cohen, S.-R. Yeh and D. L. Rousseau, J. Biol. Chem., 2022, 298, 101799 CrossRef CAS PubMed.
  27. (a) M. C. Carrasco, K. J. Dezarn, F. S. T. Khan and S. Hematian, J. Inorg. Biochem., 2021, 225, 111593 CrossRef CAS; (b) K. Ladomenou, G. Charalambidis and A. G. Coutsolelos, Polyhedron, 2013, 54, 47–53 CrossRef CAS.
  28. (a) H. Kitagishi, D. Shimoji, T. Ohta, R. Kamiya, Y. Kudo, A. Onoda, T. Hayashi, J. Weiss, J. A. Wytko and K. Kano, Chem. Sci., 2018, 9, 1989–1995 RSC; (b) Z. Halime and K. D. Karlin, in Copper-Oxygen Chemistry, ed. S. Itoh and K. D. Karlin, John Wiley & Sons, Inc., Hoboken, 2011, pp. 283–319 Search PubMed; (c) J. Meng, H. Qin, H. Lei, X. Li, J. Fan, W. Zhang, U.-P. Apfel and R. Cao, Angew. Chem., Int. Ed., 2023, 62, e202312255 CrossRef CAS.
  29. (a) Y. Yu, X. Lv, J. Li, Q. Zhou, C. Cui, P. Hosseinzadeh, A. Mukherjee, M. J. Nilges, J. Wang and Y. Lu, J. Am. Chem. Soc., 2015, 137, 4594–4597 CrossRef CAS PubMed; (b) X. Liu, Y. Yu, C. Hu, W. Zhang, Y. Lu and J. Wang, Angew. Chem., Int. Ed., 2012, 51, 4312–4316 CrossRef CAS; (c) D. Kass, S. Katz, H. Özgen, S. Mebs, M. Haumann, R. García-Serres, H. Dau, P. Hildebrandt, T. Lohmiller and K. Ray, J. Am. Chem. Soc., 2024, 146, 24808 CrossRef CAS PubMed.
  30. P. F. Urban and M. Klingenberg, Eur. J. Biochem., 1969, 9, 519–525 CrossRef CAS.
  31. J. Abramson, S. Riistama, G. Larsson, A. Jasaitis, M. Svensson-Ek, L. Laakkonen, A. Puustinen, S. Iwata and M. Wikström, Nat. Struct. Biol., 2000, 7, 910–917 CrossRef CAS.
  32. M. Verkhovsky, D. A. Bloch and M. Verkhovskaya, Biochim. Biophys. Acta, Bioenerg., 2012, 1817, 1550–1556 CrossRef CAS PubMed.
  33. (a) J. P. Collman, N. K. Devaraj, R. A. Decréau, Y. Yang, Y.-L. Yan, W. Ebina, T. A. Eberspacher and C. E. D. A Chidsey, Science, 2007, 315, 1565–1568 CrossRef CAS PubMed; (b) R. A. Decreau, J. P. Collman and A. Hosseini, Chem. Soc. Rev., 2010, 39, 1291–1301 RSC; (c) J. P. Collman, A. Dey, C. J. Barile, S. Ghosh and R. A. Decréau, Inorg. Chem., 2009, 48, 10528–10534 CrossRef CAS PubMed; (d) J. P. Collman and R. A. Decréau, Chem. Commun., 2008, 5065–5076 RSC.
  34. M. R. A. Blomberg, P. E. M. Siegbahn, G. T. Babcock and M. Wikström, J. Am. Chem. Soc., 2000, 122, 12848–12858 CrossRef CAS.
  35. S. Yoshikawa, K. Muramoto, K. Shinzawa-Itoh and M. Mochizuki, Biochim. Biophys. Acta, Bioenerg., 2012, 1817, 579–589 CrossRef CAS PubMed.
  36. M. R. A. Blomberg, P. E. M. Siegbahn and M. Wikström, Inorg. Chem., 2003, 42, 5231–5243 CrossRef CAS PubMed.
  37. A. Puustinen, M. I. Verkhovsky, J. E. Morgan, N. P. Belevich and M. Wikström, Proc. Natl. Acad. Sci. U. S. A., 1996, 93, 1545–1548 CrossRef CAS PubMed.
  38. P. Lu, M. H. Heineke, A. Koul, K. Andries, G. M. Cook, H. Lill, R. van Spanning and D. Bald, Sci. Rep., 2015, 5, 10333 CrossRef PubMed.
  39. F. L. Sousa, R. J. Alves, M. A. Ribeiro, J. B. Pereira-Leal, M. Teixeira and M. M. Pereira, Biochim. Biophys. Acta, Bioenerg., 2012, 1817, 629–637 CrossRef CAS PubMed.
  40. L. Winstedt and C. von Wachenfeldt, J. Bacteriol., 2000, 182, 6557–6564 CrossRef CAS PubMed.
  41. C. Gerdemann, C. Eicken and B. Krebs, Acc. Chem. Res., 2002, 35, 183–191 CrossRef CAS PubMed.
  42. C. Gasparetti, Biochemical and Structural Characterisation of the Copper Containing Oxidoreductases Catechol Oxidase, Tyrosinase, and Laccase from Ascomycete Fungi, 2012, vol. 16 Search PubMed.
  43. E. I. Solomon, D. E. Heppner, E. M. Johnston, J. W. Ginsbach, J. Cirera, M. Qayyum, M. T. Kieber-Emmons, C. H. Kjaergaard, R. G. Hadt and L. Tian, Chem. Rev., 2014, 114, 3659–3853 CrossRef CAS.
  44. I. Garcia-Bosch, S. M. Adam, A. W. Schaefer, S. K. Sharma, R. L. Peterson, E. I. Solomon and K. D. Karlin, J. Am. Chem. Soc., 2015, 137, 1032–1035 CrossRef CAS.
  45. S. M. Adam, I. Garcia-Bosch, A. W. Schaefer, S. K. Sharma, M. A. Siegler, E. I. Solomon and K. D. Karlin, J. Am. Chem. Soc., 2017, 139, 472–481 CrossRef CAS PubMed.
  46. (a) M. T. Kieber-Emmons, Y. Li, Z. Halime, K. D. Karlin and E. I. Solomon, Inorg. Chem., 2011, 50, 11777–11786 CrossRef CAS PubMed; (b) M. T. Kieber-Emmons, M. F. Qayyum, Y. Li, Z. Halime, K. O. Hodgson, B. Hedman, K. D. Karlin and E. I. Solomon, Angew. Chem., Int. Ed. Engl., 2012, 51, 168–172 CrossRef CAS.
  47. R. L. Peterson, J. W. Ginsbach, R. E. Cowley, M. F. Qayyum, R. A. Himes, M. A. Siegler, C. D. Moore, B. Hedman, K. O. Hodgson, S. Fukuzumi, E. I. Solomon and K. D. Karlin, J. Am. Chem. Soc., 2013, 135, 16454–16467 CrossRef CAS PubMed.
  48. H. Y. V. Ching, E. Anxolabéhère-Mallart, H. E. Colmer, C. Costentin, P. Dorlet, T. A. Jackson, C. Policar and M. Robert, Chem. Sci., 2014, 5, 2304–2310 RSC.
  49. G. B. Wijeratne, B. Corzine, V. W. Day and T. A. Jackson, Inorg. Chem., 2014, 53, 7622–7634 CrossRef CAS PubMed.
  50. C. E. Elwell, N. L. Gagnon, B. D. Neisen, D. Dhar, A. D. Spaeth, G. M. Yee and W. B. Tolman, Chem. Rev., 2017, 117, 2059–2107 CrossRef CAS PubMed.
  51. T. L. Poulos, Chem. Rev., 2014, 114, 3919–3962 CrossRef CAS PubMed.
  52. D. B. Rice, G. B. Wijeratne, A. D. Burr, J. D. Parham, V. W. Day and T. A. Jackson, Inorg. Chem., 2016, 55, 8110–8120 CrossRef CAS.
  53. L. Q. Hatcher and K. D. Karlin, J. Biol. Inorg. Chem., 2004, 9, 669–683 CrossRef CAS PubMed.
  54. A. W. Schaefer, M. T. Kieber-Emmons, S. M. Adam, K. D. Karlin and E. I. Solomon, J. Am. Chem. Soc., 2017, 139, 7958–7973 CrossRef CAS PubMed.
  55. M. Gregory, P. J. Mak, S. G. Sligar and J. R. Kincaid, Angew. Chem., Int. Ed., 2013, 52, 5342–5345 CrossRef CAS PubMed.
  56. I. G. Denisov, P. J. Mak, T. M. Makris, S. G. Sligar and J. R. Kincaid, J. Phys. Chem. A, 2008, 112, 13172–13179 CrossRef CAS PubMed.
  57. P. J. Mak, W. Thammawichai, D. Wiedenhoeft and J. R. Kincaid, J. Am. Chem. Soc., 2015, 137, 349–361 CrossRef CAS PubMed.
  58. H.-Y. Zhang, Y.-M. Sun and D.-Z. Chen, Quant. Struct.-Act. Relat., 2001, 20, 148–152 CrossRef CAS.
  59. (a) X.-Q. Zhu and C.-H. Wang, J. Org. Chem., 2010, 75, 5037–5047 CrossRef CAS PubMed; (b) X.-Q. Zhu, C.-H. Wang and H. Liang, J. Org. Chem., 2010, 75, 7240–7257 CrossRef CAS PubMed.
  60. H-bonding ability has been excluded from these calculations, we consider it to be on trend with the reported pKa values..
  61. M. G. Evans and M. Polanyi, Trans. Faraday Soc., 1938, 34, 11–24 RSC.
  62. G. I. Panov, M. V. Parfenov and V. N. Parmon, in Catalysis Reviews: Science and Engineering, Taylor & Francis, 2015, vol. 57, pp. 436–477 Search PubMed.
  63. G. B. Wijeratne, V. W. Day and T. A. Jackson, Dalton Trans., 2015, 44, 3295–3306 RSC.
  64. H. Gao and J. T. Groves, J. Am. Chem. Soc., 2017, 139, 3938–3941 CrossRef CAS PubMed.
  65. E. Nam, P. E. Alokolaro, R. D. Swartz, M. C. Gleaves, J. Pikul and J. A. Kovacs, Inorg. Chem., 2011, 50, 1592–1602 CrossRef CAS PubMed.
  66. J. J. Warren and J. M. Mayer, Biochemistry, 2015, 54, 1863–1878 CrossRef CAS PubMed.
  67. Low-temperature radical trapping experiments using an established phenolic radical trap (DIPPMPO, 5-(diisopropoxyphosphoryl)-5-methyl-1-pyrroline-N-oxide) were not successful at trapping a semiquinone intermediate. This is most likely because the internal H-bond of the semiquinone significantly weakens the second O–H bond for HAT/PCET, and during those transfers, it remains in close proximity to the cleaving peroxo moiety..
  68. We do not definitively know the possibility of the H-bonding in the cases where the catechol does not react with the peroxo, since there is no good way to probe H-bonding (i.e., such as kinetics). The lack of any observable reaction (UV-vis criteria) doesn't necessarily preclude the occurrence of H-bonding. However, based on our previous studies,45 the formation of an H-bond between acidic 4-nitrophenol and the same model peroxo complex caused changes in both the UV-vis Soret band and the O–O rRaman stretch, distinguishing the H-bonded adduct from the parent peroxo complex. Thus, in the absence of any spectroscopic change for catechols with R = H, Me and OMe, we conclude that there is no such interaction. Perhaps this is because these catechols are far less acidic (cf.Table 3, and thus far less good H-bonders)..

Footnotes

Electronic supplementary information (ESI) available: Materials and methods, DFT calculations, Schemes S1, S2, Table S1, and Fig. S1–S12. See DOI: https://doi.org/10.1039/d4sc05623j
Authors are equally contributed.

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