Next Article in Journal
Protein and Polysaccharide-Based Electroactive and Conductive Materials for Biomedical Applications
Next Article in Special Issue
Biomimetic Ketone Reduction by Disulfide Radical Anion
Previous Article in Journal
Towards a Bioactive Food Packaging: Poly(Lactic Acid) Surface Functionalized by Chitosan Coating Embedding Clove and Argan Oils
Previous Article in Special Issue
The Two Faces of the Guanyl Radical: Molecular Context and Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Disproportionation of H2O2 Mediated by Diiron-Peroxo Complexes as Catalase Mimics

Research Group of Bioorganic and Biocoordination Chemistry, University of Pannonia, H-8201 Veszprém, Hungary
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(15), 4501; https://doi.org/10.3390/molecules26154501
Submission received: 7 July 2021 / Revised: 23 July 2021 / Accepted: 24 July 2021 / Published: 26 July 2021
(This article belongs to the Special Issue Biomimetic Radical Chemistry and Applications 2021)

Abstract

:
Heme iron and nonheme dimanganese catalases protect biological systems against oxidative damage caused by hydrogen peroxide. Rubrerythrins are ferritine-like nonheme diiron proteins, which are structurally and mechanistically distinct from the heme-type catalase but similar to a dimanganese KatB enzyme. In order to gain more insight into the mechanism of this curious enzyme reaction, non-heme structural and functional models were carried out by the use of mononuclear [FeII(L1–4)(solvent)3](ClO4)2 (14) (L1 = 1,3-bis(2-pyridyl-imino)isoindoline, L2 = 1,3-bis(4′-methyl-2-pyridyl-imino)isoindoline, L3 = 1,3-bis(4′-Chloro-2-pyridyl-imino)isoindoline, L4 = 1,3-bis(5′-chloro-2-pyridyl-imino)isoindoline) complexes as catalysts, where the possible reactive intermediates, diiron-perroxo [FeIII2(μ-O)(μ-1,2-O2)(L1-L4)2(Solv)2]2+ (58) complexes are known and well-characterized. All the complexes displayed catalase-like activity, which provided clear evidence for the formation of diiron-peroxo species during the catalytic cycle. We also found that the fine-tuning of iron redox states is a critical issue, both the formation rate and the reactivity of the diiron-peroxo species showed linear correlation with the FeIII/FeII redox potentials. Their stability and reactivity towards H2O2 was also investigated and based on kinetic and mechanistic studies a plausible mechanism, including a rate-determining hydrogen atom transfer between the H2O2 and diiron-peroxo species, was proposed. The present results provide one of the first examples of a nonheme diiron-peroxo complex, which shows a catalase-like reaction.

1. Introduction

Nonheme manganese catalases (MnCAT), as alternatives to the heme-containing catalases, contain a binuclear active sites and deplete hydrogen peroxide in cells through a ping-pong mechanism interconventing between Mn2(II,II) and Mn2(III,III) states in a two-electron catalytic cycle [1,2]. Manganese catalase enzymes such as Lactobacillus plantarum (LPC) [3,4], Thermus thermophilus (TTC) [5,6], Thermoleophilium album (TAC) [7], and Pyrobaculum calidifontis VA1 (PCC) [8] have been isolated and characterized from both bacteria and archae. Despite their large numbers and importance, few catalase enzyme structures are known and they can also be sub-divided into two groups, each with a different active site.
Among them both LPC and TTC have a dimanganese catalytic core with Glu3His2 ligands coupled by two single-atom solvent bridges (Scheme 1A) [4,9]. In contrast to the above structure, the cyanobacterial manganese catalase (KatB) from Anabaena shows a markedly different active site with canonical Glu4His2 coordination geometry and two terminal water molecules (Scheme 1B). However, it closely resembles the active site of the ferritin-like ruberythrin (RBR), which is a nonheme diiron protein [10]. The Mn-Mn distances (~3.8 Å) in KatB are similar to that was observed for the iron-containing ribonucleotide reductase R2 (3.6–3.7 Å), but much shorter distances have been found for LPC and TTC (3.0–3.1 Å). The H2O2 binds terminally to one of the Mn center in LPC/TTC, while in KatB and RBR, the symmetrical active sites and the relatively long M-M distances favor its μ-1,2 binding mode [11]. These enzymes, coupled with superoxide dimutates (SOD), are responsible for the full detoxification of O2•─ through the formation of H2O2 and its redox disproportionation into water and dioxygen. H2O2 can also be classified as a ROS particle, since it readily reacts with reduced transition metals and results in highly reactive HO radical via Fenton-like reactions. Many rationally designed synthetic biomimetics are being developed against oxidative stress to treat various diseases, including Alzheimer’s, cancer, aging, neurodegenerative, inflammatory and heart diseases [12,13,14,15,16,17,18,19,20,21]. The identification and study of the structure of the above enzymes can greatly contribute to the design of new catalase mimetics for potential therapeutic use. The structural and functional investigations on low molecular-weight copper, manganese and iron containing model compounds that are focused mainly on the effect of the used metals and their ligand environment (donor sites) on redox potential and through it on reactivity. Based on the literature data, various Schiff base, aminopyridine and azamacrocyclic derivatives have proven to be the most popular and practical ligands [12,13,14,15,16,17,18,19,20,21]. Considering the reactivity of manganese catalases (KatB) toward H2O2, including a ping-pong mechanism interconventing between Mn2(II,II) and Mn2(III,III) states in a two-electron catalytic cycle, a similar interaction with H2O2 has not been reported for diiron(III) peroxo intermediates which may be proposed in the catalytic cycles of RBR. Progress in recent years has made a number of diiron(III)-peroxo complexes available that can be generated by stoichiometric amount of H2O2 [22,23,24,25,26,27]. Therefore, the opportunity has arisen to investigate both electrophilic and nucleophilic reactivity towards various substrates, such as phenols, N,N-dimethylanilines, and various aldehydes, respectively [25,26]. The differences in reactivity observed in the investigated systems, as well as the stability of the peroxo intermediates can in principle be explained by the disproportionation of H2O2 as a competing process. Clarification of the relationship between catalase-like activity and catalytic oxidation is an important factor in the design and development of high efficiency catalytic systems. In addition, the investigation of the metal-based disproportionation of H2O2 may help us to understand the nature of the catalytically active species present both, in the enzymatic and biomimetic systems. We have reported that the [FeII(L1)(solvent)3](ClO4)2 (1) is an efficient catalyst for the oxidation of thioanisoles and benzyl alcohols by the use of H2O2 as cooxidant, where a metastabile green species (5) with a FeIII(μ-O)(μ-1,2-O2)FeIII core was observed and characterized by UV-Vis, EPR, rRaman, and X-ray absorption spectroscopic measurements [24]. The reactivity of 5 was also published both in nucleophilic (deformylation of aldehydes) and electrophilic (oxidation of phenols) reactions [27].
As a possible functional model of RBR, we recently investigated the reactivity of the peroxo adduct [Fe2(μ-O2)(MeBzim-Py)4(CH3CN)]4+ (MeBzim-Py = (2-(2′-pyridyl)-N-methylbenzimidazole) with H2O2 and found direct kinetic and computational evidence for the formation of low-spin oxoiron(IV) via dissociation process [22,23]. As a continuation of this tudy, the previously reported and fully characterized FeIII2(μ-O)(μ-1,2-O2)(IndH)2(Solv)2]2+ (IndH = L1 = 1,3-bis(2-pyridyl-imino)-isoindolines) [24,25] and their substituted isoindolin-containing derivatives (L2–L4, 24) were chosen as model compounds, in which the dissociation of the diiron(III) core can be ruled out due to the μ-oxo-bridge, and investigated its reactivity towards H2O2 to gain insight into the mechanism of RBR enzymes (Scheme 2).

2. Results and Discussions

In view of the above results, further studies were planned to elucidate the mechanism of both the diiron(III)-peroxo formation and its decay mediated by H2O2. Complex 5 from its precursor complex 1 can be generated even at room temperature by the use of excess of H2O2 in CH3CN. Its formation and decay process was followed as an increase and decrease in absorbance at 690 nm (ε = 1500 M−1 cm−1 per Fe ion and assuming 100% conversion), which can be assigned as a charge transfer between FeIII and the O22− ligand (Figure 1a) [24]. Similar spectral feature can be observed for the complexes 68: 691 nm (ε = 1404 M−1 cm−1), 692 nm (ε = 1460 M−1 cm−1), and 693 nm (ε = 1308 M−1 cm−1), respectively. It is worth noting that the green species (5), after decomposition, can be regenerated 2–3 times by adding additional H2O2 without significant decrease on the yields (Table 1, Figure 1b and inset in Figure 2a). Similar behavior and yields were observed for 6, but much lower values were obtained for 7 and 8. These results are consistent with the stability and/or the reactivity of the intermediates at high concentrations of added water, providing further information for the possible role of the binding water during the formation of the reactive oxidant.
Preliminary kinetic studies were performed at 20 °C to establish the role of the observed diiron(III)-peroxo intermediates in the catalytic disproportionation reaction of H2O2. The formation and decomposition of 5 in the presence of H2O2 was monitored by following the rise and fall of its visible chromophore (Table 2 and Table 3), as well as the appearance of the dioxygen formed by gas volumetric methods (Figure 2b). Dioxygen formation started after a short lag phase during which the peroxo-adduct (5) accumulated. The lag phase was followed by a linear dioxygen formation period during which time 5 persisted in a pseudo-steady-state. This profile is similar to that found for the water-assisted decay of FeIII(H2O)(OOH) intermediate into the reactive FeV(O)(OH) species [28]. Upon depletion of H2O2, 5 decomposed rapidly and the dioxygen formation ceased. The yield of dioxygen, TON (turnover number = mol H2O2/mol catalyst) and the TOF (turnover frequency = mol H2O2/mol catalyst/h) values were determined to be Yield = 85%, TON = 10.4 and TOF = 288 h−1 as a result of volumetric method via the measurements of dioxygen evolution from the reaction mixture ([H2O2] = 0.243 M and [1] = 0.002 M) at 20 °C in CH3CN. Similar values were obtained for the complexes 2, 3 and 4 under the same conditions (Table 4). The results clearly indicate that diiron-peroxo intermediates play a key role in the formation of the reactive species responsible for the H2O2 oxidation.

2.1. Kinetic Studies on the Formation of Diiron-Peroxo Complexes (58)

Detailed kinetic studies on the reaction of 1 (2 mM) with H2O2 (24–98 mM) were carried out in CH3CN at 15 °C, and the formation of the green species 5 was followed by UV-vis spectroscopy as an increase in absorbance at 690 nm under pseudo-first-order conditions (excess of H2O2) (Table 2). At constant [1]0 = 2 mM, the formation of 5 exhibits a first-order dependence, and the first-order rate constants show a good linear dependence with [H2O2]0 (Figure 3a), affording the second-order rate constant k2 = 0.93 M−1 s−1 at 15 °C and 1.09 M−1 s−1 at 20 °C. These values are about 5–6 times smaller than those obtained for [FeII(MeBzim-Py)]2+ (5.5 M−1 s−1), FeII(MeBzim-Py)2+ (6.6 M−1 s−1) and FeII(TBA)2+ (6.54 M−1 s−1) complexes under the same conditions at 20 °C [23,26]. It is worth noting that adding large volumes of water (~0.5 M) resulted in nearly twice the rate of the diiron-peroxo formation. Furthermore, a kinetic isotope effect (KIE) of 2.99 and 3.07 were observed for the formation of 6 and 7, when the experiments were carried out in the presence of added H2O or D2O. This value may represent the result of much more multiple effects compared to the elemental step, but clearly indicate the key role of the water during the diiron-peroxo complex formation.
In the next step, we investigated the effect of the the ligand modification by introducing electron withdrawing (Cl) or electron releasing (CH3) substituents on the pyridine arms in the fourth and fifth positions with emphasis on the redox potential and Lewis acidity of the precursor complex. Since the redox potential of the substituted complexes can be measured under the same conditions, it can be used as excellent reactivity descriptor. On the cyclic voltammogram of the Fe(II)-isoindoline complexes (14) irreversible reduction and oxidation peaks can be recognized within the range of −440 to +1160 mV (vs. Fc/Fc+). The huge peak separations (ΔE = EpaEpc, Table 4) indicates the irreversibility of the redox processes and suggests chemical reactions coupled to the electrode reaction. The results, obtained at the scan rate of 100 mV/s, show that oxidation potential is mainly influenced by the electronic properties of the ligands, since the anodic peak potentials (Epa) of the complexes are increasing in the order of electron releasing ability of the substituent on the pyridine moiety (Figure 4). The Epa spans a 128 mV range from 655 mV (1) to 783 mV (4), which is not too large but informative, in terms of the effect of the substituents. However, the fact that the anodic peak of complex 1 can be found at lower potentials than it could be expected according to electronic effects, suggest that steric effects are also influencing the oxidation reactions. One can assume that the steric effect of the methylated and chlorinated ligands can cause an anodic shift in the oxidation peak potential of the complexes.
The four iron complexes, 14 have been compared to investigate the effect of the various aryl substituents. Figure 3a shows the results of the activity for the precursor complexes (14) during the formation of their appropriate diiron-peroxo intermediates (58). This demonstrates only small differences exist, based on the first-order rate constants obtained under the same conditions. It was found that complex 4 with 5-chloropyridyl side chains is the most reactive complex with kobs of 3.66 M−1 s−1, whilst complex 1 with unsubstituted side chains was proved to be the least active with kobs of 0.93 M−1 s−1 at 15 °C. Based on the CV data, we have got a clear evidence that the formation rate of the diiron-peroxo species (k2) increases almost linearly with the redox potential (Epa for FeIII/FeII) of the precursor complex (Figure 3b). In summary, the higher the redox potentials of FeIII/FeII redox couple the higher is the activity of the precursor complex towards to H2O2. Complexes with 4- and 5-chloro-pyridine side chains (3 and 4), whose metal sites are electron-deficient are faster in their reaction with H2O2 than electron-rich derivatives 1 and 2 with more Lewis basic side chains.

2.2. Catalase-Like Activity of Peroxo-Diiron Complexes

The ability of the in situ generated diiron-peroxo species (5) to mediate the disproportionation reaction of H2O2 into O2 and H2O was tested in CH3CN by UV-vis spectroscopy as a decrease in absorbance at 690 nm under pseudo-first-order conditions (excess of H2O2) (Table 3). The initial rate of decay of diiron-peroxo species was measured as a function of the complex and substrate concentrations in CH3CN. At constant [H2O2]0 = 300 mM, the disproportionation reaction shows first-order kinetic behavior on [5], and the first-order rate constant k1 = 4.968 × 10−3 s−1 was obtained at 20 °C from the slope of the plot of Vi versus [5]0 (Figure 5a). At constant [5]0 = 2 mM, initial rates (Vi) exhibit a linear dependence with [H2O2]0 (Figure 5b) affording first-order rate constant k1′ = 29.63 × 10−3 s−1. The second-order rate constant, k2 = 3.4 M−1 s−1 was obtained from either k1/[H2O2]0 or k1′/[5]0. This value is much smaller than that was found for the analogue [Mn(L1)]2+ complex (k2(kcat/KM) = 79 ± 4 M−1 s−1) in protic solution, where MnIV(O) intermediate was proposed as reactive species [29]. A kinetic isotope effect (KIE) of 1.73 and 1.66 were observed for the decay of 6 and 7, when the experiments were carried out in the presence of added H2O or D2O. These results indicate the key role of the water during the diiron-peroxo mediated disproportionation reaction.
In the next step, we investigated the effect of the ligand modification by introducing electron donating (-CH3) and electron withdrawing (-Cl) phenyl-ring substituents. The substituted ligand-containing diiron-peroxo intermediates 5, 6, 7 and 8 show an increasing rate in the listed order (Figure 6a). It was found that complex 8 with 5-chloropyridine side chains is the most efficient oxidant with the fastest rate k2 = 23.8 × 10−2 M−1 s−1, while complexes 5 and 6 with unsubstituted and 4-methylpyridine side chains are the less efficient oxidants with k2 = 1.52 × 10−2 M−1 s−1 and k2 = 3.05 × 10−2 M−1 s−1, respectively. These results give clear evidence that substituents affect the redox potential and the catalase activity of the complexes. Table 4 and Figure 6b show that the redox potentials (Epa) of the complexes and their activity increase with increasing electron-withdrawing ability of the substituent.
Based on the temperature dependence of the reactivity of diiron-peroxo complexes (Figure 7a), the experimentally determined the difference between ΔG values was 7.3 kJ mol−1. Significantly lower free activation energy value was obtained for the 8-mediated disproportionation of H2O2 in comparison to 5. The calculated Gibbs energy values correlate very well with the reaction rates (Figure 7b). The values of −TΔS were lower than ΔH in the investigated temperature range, indicating an enthalpy-controlled reactions. As a result of a compensation effect increasing activation enthalpies are offset by increasingly poitive entropies yielding ΔH = 81 kJ mol−1 at the intercept (Figure 8), which value is a little bit higher than that was obtained for the conversion alkylperoxo-iron(III) intermediates to oxoiron(IV) through O-O bond homolysis (ΔH = 61.3 kJ mol−1) [30].

3. Experimental

3.1. Materials and Methods

The ligands 1,3-bis(2′-pyridylimino)isoindolines (L1 = indH, L2 = 4Me-indH, L3 = 4Cl-indH, L4 = 5Cl-indH,) and their complexes (14) were prepared according to published procedures [31]. All manipulations were performed under a pure argon atmosphere using standard Schlenk-type inert-gas techniques. Solvents used for the reactions were purified by literature methods and stored under argon. The starting materials for the ligand are commercially available and they were purchased from Sigma Aldrich. The UV-visible spectra were recorded on an Agilent 8453 diode-array spectrophotometer using quartz cells. IR spectra were recorded using a Thermo Nicolet Avatar 330 FT-IR instrument (Thermo Nicolet Corporation, Madison, WI, USA), Samples were prepared in the form of KBr pellets. Microanalyses elemental analysis was done by the Microanalytical Service of the University of Pannonia.

3.2. Characterization of Ligands and Their Complexes

indH: FT-IR (ATR) ν = 3199 (w), 3061 (w), 1622 (s), 1606 (m), 1577 (s), 1550 (s), 1454 (s), 1427 (s), 1373 (m), 1306 (m), 1259 (s), 1217 (s), 1139 (m), 1097 (m), 1042 (m), 991 (w), 885 (w), 875 (w), 859 (w), 839 (w), 808 (m), 792 (s), 783 (s), 769 (s), 737 (s), 708 (m), 696 (s), 689 (s), 627 (w). UV/Vis (DMF): λmax (logε) = 317 (3.09), 332 (3.18), 348 (3.21), 368 (3.28), 386 (3.34), 410 (3.11) nm.
4Me-ind: FT-IR (ATR) ν = 3209 (w), 3049 (w), 1645 (m), 1627 (s), 1591 (s), 1541 (m), 1460 (s), 1408 (w), 1364 (m), 1305 (m), 1281 (m), 1242 (s), 1186 (m), 1132 (w), 1114 (w), 1101 (m), 1036 (m), 997 (w), 929 (m), 891 (m), 848 (w), 810 (s), 779 (m), 742 (w), 694 (s), 660 (w). UV/Vis (DMF): λmax (logε) = 316 (2.99), 332 (3.08), 348 (3.12), 366 (3.18), 387 (3.21), 411 (2.98) nm.
4Cl-ind: FT-IR (ATR) ν = 3213 (w), 1743 (w), 1637 (m), 1624 (m), 1565 (s), 1539 (m), 1448 (s), 1359 (m), 1309 (m), 1300 (m), 1281 (w), 1219 (m), 1184 (w), 1089 (m), 1040 (m), 989 (m), 895 (s), 877 (s), 804 (m), 779 (m), 735 (s), 701 (s) 687 (m), 656 (w). UV/Vis (DMF): λmax (logε) = 315 (3.16), 332 (3.27), 348 (3.29), 367 (3.31), 385 (3.35), 409 (3.12) nm.
5Cl-ind: FT-IR (ATR) ν = 3317 (w), 3228 (w), 1745 (m), 1633 (s), 1572 (s), 1470 (m), 1447 (s), 1359 (m), 1311 (m), 1256 (w), 1244 (w), 1228 (m), 1217 (m), 1184 (w), 1105 (s), 1034 (s), 1007 (m), 955 (w), 924 (w), 880 (w), 843 (s), 773 (s), 754 (m), 725 (w), 702 (s), 691 (s), 638 (m), 619 (w), 602 (m). UV/Vis (DMF): λmax (logε) = 338 (3.15), 354 (3.15), 376 (3.19), 395 (3.25), 421 (3.02) nm.
1: FT-IR (ATR) ν = 3329 (w), 1728 (w), 1654 (m), 1626 (s), 1614 (m), 1593 (w), 1554 (w), 1524 (s), 1485 (s), 1467 (m), 1433 (m), 1373 (w), 1304 (w), 1262 (w), 1207 (m), 1055 (s), 929 (w), 860 (w), 779 (s), 714 (m), 621 (m). C24H20Cl2FeN8O8 (675.22): calcd. C 42.69, H 2.99, Cl 10.15, N 16.60; found C 42.82, H 2.93, Cl 10.11, N 16.42. UV/Vis (CH3CN): λmax (logε) = 315 (3.36), 329 (3.45), 345 (3.48), 365 (3.56), 382 (3.61), 406 (3.37) nm.
2: FT-IR (ATR) ν = 3456 (w), 1734 (w), 1665 (m), 1654 (s), 1605 (s), 1551 (m), 1520 (s), 1493 (s), 1477 (m), 1367 (w), 1313 (w), 1279 (m), 1244 (w), 1215 (w), 1049 (s), 941 (w), 827 (w), 777 (w), 752 (w), 710 (m), 621 (m). C26H27Cl2FeN8O8 (706.29): calcd. C 44.21, H 3.85, Cl 10.04, N 15.87; found C 43.99, H 3.64, Cl 9.58, N 15.79. UV/Vis (CH3CN): λmax (logε) = 315 (3.48), 328 (3.58), 345 (3.62), 365 (3.71), 383 (3.76), 406 (3.50) nm.
3: FT-IR (ATR) ν = 3366 (w), 3217 (w), 1773 (w), 1724 (w), 1670(m), 1628 (m), 1564 (m), 1543 (m), 1516 (s), 1483 (m), 1460 (s), 1362 (w), 1306 (w), 1228 (w), 1205 (m), 1090 (s), 1051 (s), 910 (m), 867 (w), 823 (w), 744 (s), 710 (s), 621 (m). C24H21Cl4FeN8O8 (747.13): calcd. C 38.58, H 2.83, Cl 18.98, N 15.00; found C 38.25, H 2.95, Cl 18.63, N 15.52. UV/Vis (CH3CN): λmax (logε) = 328 (3.10), 344 (3.09), 363 (3.04), 383 (2.99), 406 (2.71) nm.
4: FT-IR (ATR) ν = 3323 (w), 3228 (w), 1773 (w), 1717 (w), 1651(w), 1628 (m), 1612 (m), 1546 (m), 1521 (s), 1477 (s), 1458 (s), 1391 (w), 1366 (m), 1335 (w), 1298 (w), 1261 (w), 1236 (m), 1205 (m), 1096 (s), 1076 (s), 922 (m), 851 (w), 835 (m), 781 (w), 752 (m), 711 (s), 650 (w), 621 (m). C24H21Cl4FeN8O8 (747.13): calcd. C 38.58, H 2.83, Cl 18.98, N 15.00; found C 38.81, H 3.04, Cl 18.63, N 15.39. UV/Vis (CH3CN): λmax (logε) = 335 (2.64), 353 (2.59), 373 (2.47), 395 (2.36), 418 (2.04) nm.

3.3. Cyclic Voltammetry

Cyclic voltammetry experiments were performed on a Radiometer Analytical PGZ-301 potentiostat with a conventional three-electrode configuration, consisting of a glassy carbon working electrode (ID = 3 mm), a platinum wire auxiliary electrode, and Ag/AgCl reference electrode. Potentials are referenced to the Fc/Fc+ redox couple. Ferrocene was added as an internal standard at the end of the experiments. Procedure for the cyclic voltammetry experiment: 0.05 mmol of 14 was dissolved in 10 mL dry acetonitrile, containing 349.1 mg (1 mmol) of tetrabutylammonium-perchlorate, which served as supporting electrolyte. The solution was bubbled with argon to remove dissolved gas residuals and to ensure inert atmosphere during measurements. The working electrode was wet polished on 0.5 μm alumina slurry or emery paper grade 500, after each measurement. Cyclic voltammograms were recorded with a scan rate of 100 mV s−1.

3.4. Generation of Diiron-Peroxo Complexes (58)

Precursor complexes (14) were dissolved in 1.5 mL acetonitrile and added 4 equivalent H2O2, in different temperature (0, 5, 10 15 °C) and the reactions were followed with UV-Vis spectroscopy at 680 nm, the cuvette length was 1 cm.

3.5. Diiron-Peroxo-Mediated Disproportionaton of H2O2

Reactions were carried out by mixing the substrate (H2O2) with the in situ generated complex 58 (1 mM) in 1.5 mL acetonitrile in different temperature (0, 5, 10 15 °C) and the reactions were followed with UV-Vis spectroscopy at 680 nm, the cuvette length was 1 cm. To ensure the reliability of the measurements, 2–3 parallel measurements were performed.
Catalytic reactions were carried out at 20 °C in a 30 cm3 reactor containing stirring bar under air. In a typical experiment the complex was dissolved dissolved in 20 cm3 CH3CN, and the flask was closed with a rubber septum. H2O2 was injected by syringe through the septum. The reactor was connected to a graduated burette filled with oil, and the evolved dioxygen was measured volumetrically at time intervals of 2 s.

4. Conclusions

In summary, the results of the reaction kinetics and CV measurements are shown in Table 4. Based on these data the following reaction mechanism, including the diiron-peroxo complex formation, and its reaction with H2O2, was proposed (Scheme 3). We obtained clear evidence for the role of the peroxo-intermediate in diiron catalase mimics. We also found that the higher the redox potentials of the FeIII/FeII redox couple, the higher the catalase-like activity, and the complexes with electron-deficient metal sites are significantly more reactive. It can be explained by their electrophilic character. The results obtained may contribute to the elucidation of the mechanism of both dimanganase and diiron-containing catalase enzymes.

Author Contributions

Conceptualization, J.K.; resources, D.L.-B., P.T., F.V.C., S.K. and B.G., writing—original draft preparation, J.K., writing—review and editing, J.K. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support of the GINOP-2.3.2-15-2016-00049 (J.K.) and TKP2020-IKA-07 (J.K. and D.L.-B.) are gratefully acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not available.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Beyer, W.F.; Fridovich, I. Catalases-with and without heme. Basic Life Sci. 1988, 49, 651–661. [Google Scholar] [PubMed]
  2. Nicholls, P.; Fita, I.; Loewen, P.C. Enzymology and structure of catalases. Adv. Inorg. Chem. 2001, 51, 51–106. [Google Scholar] [CrossRef]
  3. Kono, Y.; Fridovich, I. Isolation and characterization of the pseudocatalase of Lactobacillus plantarum. J. Biol. Chem. 1983, 258, 6015–6019. [Google Scholar] [CrossRef]
  4. Barynin, V.V.; Whittaker, M.M.; Antonyuk, S.V.; Lamzin, V.S.; Harrison, P.M.; Artymiuk, P.J.; Whittaker, J.W. Crystal Structure of Manganese Catalase from Lactobacillus plantarum. Structure 2001, 9, 725–738. [Google Scholar] [CrossRef]
  5. Antonyuk, S.V.; Melik-Adman, V.R.; Popov, A.N.; Lamzin, V.S.; Hempstead, P.D.; Harrison, P.M.; Artymyuk, P.J.; Barynin, V.V. Three-dimensional structure of the enzyme dimanganese catalase from Thermus thermophilus at 1 Å resolution. Crystallogr. Rep. 2000, 45, 105–113. [Google Scholar] [CrossRef]
  6. Barynin, V.V.; Grebenko, A.I. T-catalase is nonheme catalase of the extremely thermophilic bacterium Thermus thermophilus HB8. Dokl. Akad. Nauk SSSR 1986, 286, 461–464. [Google Scholar]
  7. Allgood, G.S.; Perry, J.J. Characterization of a manganese-containing catalase from the obligate thermophile Thermoleophilum album. J. Bacteriol. 1986, 168, 563–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Amo, T.; Atomi, H.; Imanaka, T. Unique Presence of a Manganese Catalase in a Hyperthermophilic Archaeon, Pyrobaculum calidifontis VA1. J. Bacteriol. 2002, 184, 3305–3312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Whittaker, J.W. Non-heme manganese catalase-the “other” catalase. Arch. Biochem. Biophys. 2012, 525, 111–120. [Google Scholar] [CrossRef] [Green Version]
  10. Cardenas, J.P.; Quatrini, R.; Holmes, D.S. Aerobic Lineage of the Oxidative Stress Response Protein Rubrerythrin Emerged in an Ancient Microaerobic, (Hyper)Thermophilic Environment. Front. Microbiol. 2016, 7, 1822–1831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Bihani, S.C.; Chakravarty, D.; Ballal, A. KatB, a cyanobacterial Mn-catalase with unique active site configuration: Implications for enzyme function. Free Radic. Biol. Med. 2016, 93, 118–129. [Google Scholar] [CrossRef]
  12. Signorella, S.; Palopoli, C.; Ledesma, G. Rationally designed mimics of antioxidant manganoenzymes: Role of structural features in the quest for catalysts with catalaseand superoxide dismutase activity. Coord. Chem. Rev. 2018, 365, 75–102. [Google Scholar] [CrossRef]
  13. Wu, A.J.; Penner-Hahn, J.E.; Pecoraro, V.L. Structural, Spectroscopic, and Reactivity Models for the Manganese Catalases. Chem. Rev. 2004, 104, 903–938. [Google Scholar] [CrossRef] [PubMed]
  14. Tovmasyan, A.; Maia, C.G.C.; Weitner, T.; Carballal, S.; Sampaio, R.S.; Lieb, D.; Ghazaryan, R.; Ivanovic-Burmazovic, I.; Radi, R.; Reboucas, J.S.; et al. A comprehensive evaluation of catalase-like activity of different classes of redox-active therapeutics. Free Radic. Biol. Med. 2015, 86, 308–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Batinic-Haberle, I.; Tovmasyan, A.; Spasojevic, I. An educational overview of the chemistry, biochemistry and therapeutic aspects of Mn porphyrins—From superoxide dismutation to HO-driven pathways. Redox Biol. 2015, 5, 43–65. [Google Scholar] [CrossRef] [Green Version]
  16. Kaizer, J.; Baráth, G.; Speier, G.; Réglier, M.; Giorgi, M. Synthesis, structure and catalase mimics of novel homoleptic manganese(II) complexes of 1,3-bis(2′-pyridylimino)isoindoline, Mn(4R-ind)2 (R = H, Me). Inorg. Chem. Commun. 2007, 10, 292–294. [Google Scholar] [CrossRef]
  17. Kaizer, J.; Kripli, B.; Speier, G.; Párkányi, L. Synthesis, structure, and catalase-like activity of a novel manganese(II) complex: Dichloro[1,3-bis(2’-benzimidazolylimino)isoindoline]manganese(II). Polyhedron 2009, 28, 933–936. [Google Scholar] [CrossRef]
  18. Kaizer, J.; Csay, T.; Kővári, P.; Speier, G.; Párkányi, L. Catalase mimics of a manganese(II) complex: The effect of axial ligands and pH. J. Mol. Catal. A Chem. 2008, 280, 203–209. [Google Scholar] [CrossRef]
  19. Kripli, B.; Garda, Z.; Sólyom, B.; Tircsó, G.; Kaizer, J. Formation, stability and catalase-like activity of mononuclear manganese(II) and oxomanganese(IV) complexes in protic and aprotic solvents. NJC 2020. [Google Scholar] [CrossRef]
  20. Stadtman, E.R.; Berlett, B.S.; Chock, P.B. Manganese-dependent disproportionation of hydrogen peroxide in bicarbonate buffer. Proc. Natl. Acad. Sci. USA 1990, 87, 384–388. [Google Scholar] [CrossRef] [Green Version]
  21. Kripli, B.; Sólyom, B.; Speier, G.; Kaizer, J. Stability and Catalase-Like Activity of a Mononuclear Non-Heme Oxoiron(IV) Complex in Aqueous Solution. Molecules 2019, 24, 3236. [Google Scholar] [CrossRef] [Green Version]
  22. Szávuly, M.I.; Surducan, M.; Nagy, E.; Surányi, M.; Speier, G.; Silaghi-Dumitrescu, R.; Kaizer, J. Functional models of nonheme enzymes: Kinetic and computational evidence for the formation of oxoiron(IV) species from peroxo-diiron(III) complexes, and their reactivity towards phenols and H2O2. Dalton Trans. 2016, 45, 14709–14718. [Google Scholar] [CrossRef] [PubMed]
  23. Pap, J.S.; Draksharapu, A.; Giorgi, M.; Browne, W.R.; Kaizer, J.; Speier, G. Stabilisation of mu-peroxido-bridged Fe(III) intermediates with non-symmetric bidentate N-donor ligands. Chem. Commun. 2014, 50, 1326–1329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pap, J.S.; Cranswick, M.A.; Balogh-Hergovich, É.; Baráth, G.; Giorgi, M.; Rohde, G.T.; Kaizer, J.; Speier, G.; Que, L., Jr. An Iron(II)[1,3-bis(2’-pyridylimino)isoindoline] Complex as a Catalyst for Substrate Oxidation with H2O2—Evidence for a Transient Peroxidodiiron(III) Species. Eur. J. Inorg. Chem. 2013, 3858–3866. [Google Scholar] [CrossRef] [Green Version]
  25. Kripli, B.; Csendes, F.V.; Török, P.; Speier, G.; Kaizer, J. Stoichiometric Aldehyde Deformylation Mediated by Nucleophilic Peroxo-diiron(III) Complex as a Functional Model of Aldehyde Deformylating Oxygenase. Chem. Eur. J. 2019, 25, 14290–14294. [Google Scholar] [CrossRef]
  26. Török, P.; Unjaroen, D.; Csendes, F.V.; Giorgi, M.; Browne, W.R.; Kaizer, J. A nonheme peroxo-diiron(III) complex exhibiting both nucleophilic and electrophilic oxidation of organic substrates. Dalton Trans. 2021, 50, 7181–7185. [Google Scholar] [CrossRef] [PubMed]
  27. Kripli, B.; Szávuly, M.; Csendes, F.V.; Kaizer, J. Functional models of nonheme diiron enzymes: Reactivity of the mu-oxo-mu-1,2-peroxo-diiron(III) intermediate in electrophilic and nucleophilic reactions. Dalton Trans. 2020, 49, 1742–1746. [Google Scholar] [CrossRef] [PubMed]
  28. Oloo, W.N.; Fielding, A.J.; Que, L., Jr. Rate determining water assisted O-O bond cleavage of a FeIII-OOH intermediate in a bioinspired nonheme iron-catalyzed oxidation. J. Am. Chem. Soc. 2013, 135, 6438–6441. [Google Scholar] [CrossRef]
  29. Meena, B.I.; Kaizer, J. Design and Fine-Tuning Redox Potentials of Manganese(II) Complexes with Ioindoline-Based Ligands: H2O2 Oxidation and Oxidative Bleaching Performance in Aqueous Solution. Catalysts 2020, 10, 404. [Google Scholar] [CrossRef] [Green Version]
  30. Jensen, M.P.; Payeras, A.M.I.; Fiedler, A.T.; Costas, M.; Kaizer, J.; Stubna, A.; Münck, E.; Que, L., Jr. Kinetic Analysis of the Conversion of Nonheme (Alkylperoxo)iron(III) Species to Iron(IV) Complexes. Inorg. Chem. 2007, 46, 2398–2408. [Google Scholar] [CrossRef] [Green Version]
  31. Csonka, R.; Speier, G.; Kaizer, J. Isoindoline-derived ligands and applications. RSC Adv. 2015, 5, 18401–18419. [Google Scholar] [CrossRef]
Scheme 1. Binding mode of H2O2 in nonheme dimanganese and diiron enzymes.
Scheme 1. Binding mode of H2O2 in nonheme dimanganese and diiron enzymes.
Molecules 26 04501 sch001
Scheme 2. Structural formulae for the ligands and their complexes used in the disproportionation reaction of H2O2.
Scheme 2. Structural formulae for the ligands and their complexes used in the disproportionation reaction of H2O2.
Molecules 26 04501 sch002
Figure 1. (a) Characteristic UV-Vis spectra of FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58) intermediates in CH3CN. (b) Dependence of the yield of peroxo complexes (58) on the number of cycles. [Fe]0 = 2 mM, [H2O2]0 = 24 mM (4x) in CH3CN at 20 °C.
Figure 1. (a) Characteristic UV-Vis spectra of FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58) intermediates in CH3CN. (b) Dependence of the yield of peroxo complexes (58) on the number of cycles. [Fe]0 = 2 mM, [H2O2]0 = 24 mM (4x) in CH3CN at 20 °C.
Molecules 26 04501 g001
Figure 2. (a) UV-Vis spectral changes of 5 during the 1-catalyzed disproportionation of H2O2. The inset shows the the time course of the formation and decay of 5 and the parallel formation of dioxygen for 5 cycles. [5]0 = 2 mM, [H2O2]0 = 24 mM (4×) in CH3CN at 20 °C (b) Time course for the H2O2 (24 mM) oxidation by 5 (2 mM), monitored at 690 nm in CH3CN at 20 °C and by volumetrically determined amounts of evolved dioxygen.
Figure 2. (a) UV-Vis spectral changes of 5 during the 1-catalyzed disproportionation of H2O2. The inset shows the the time course of the formation and decay of 5 and the parallel formation of dioxygen for 5 cycles. [5]0 = 2 mM, [H2O2]0 = 24 mM (4×) in CH3CN at 20 °C (b) Time course for the H2O2 (24 mM) oxidation by 5 (2 mM), monitored at 690 nm in CH3CN at 20 °C and by volumetrically determined amounts of evolved dioxygen.
Molecules 26 04501 g002
Figure 3. (a) Formation of diiron-peroxo (58) complexes. Plots of kobs against H2O2 concentration to determine a second-order rate constant. [14]0 = 2 mM, in CH3CN at 15 °C. (b) Dependence of the formation rate of peroxo-diiron complexes (58) on the oxidation potential (Epa) of the [FeII(L1–4)(solvent)3](ClO4)2 (14) complexes.
Figure 3. (a) Formation of diiron-peroxo (58) complexes. Plots of kobs against H2O2 concentration to determine a second-order rate constant. [14]0 = 2 mM, in CH3CN at 15 °C. (b) Dependence of the formation rate of peroxo-diiron complexes (58) on the oxidation potential (Epa) of the [FeII(L1–4)(solvent)3](ClO4)2 (14) complexes.
Molecules 26 04501 g003
Figure 4. Cyclic voltammograms of [FeII(L1–4)(solvent)3](ClO4)2 complexes (14) (5 mM in acetonitrile, supporting electrolyte: tetrabutylammonium-perchlorate (0.1 M), scan rate: 100 mV/s).
Figure 4. Cyclic voltammograms of [FeII(L1–4)(solvent)3](ClO4)2 complexes (14) (5 mM in acetonitrile, supporting electrolyte: tetrabutylammonium-perchlorate (0.1 M), scan rate: 100 mV/s).
Molecules 26 04501 g004
Figure 5. 5-mediated disproportionation of H2O2 in CH3CN at 25 °C. (a) Dependence of the reaction rate (Vi) for H2O2 oxidation on the diiron-peroxo complex 5 concentration. [H2O2]0 = 300 mM (b) Dependence of the reaction rate (Vi) for H2O2 oxidation on the H2O2 concentration. [5]0 = 1 mM.
Figure 5. 5-mediated disproportionation of H2O2 in CH3CN at 25 °C. (a) Dependence of the reaction rate (Vi) for H2O2 oxidation on the diiron-peroxo complex 5 concentration. [H2O2]0 = 300 mM (b) Dependence of the reaction rate (Vi) for H2O2 oxidation on the H2O2 concentration. [5]0 = 1 mM.
Molecules 26 04501 g005
Figure 6. Peroxo-diiron complexes-mediated disproportionation of H2O2 in CH3CN at 15 °C. (a) Dependence of the reaction rate (Vi) for H2O2 oxidation on the diiron-peroxo complex 58 concentration. [H2O2]0 = 24 mM (b) Dependence of the decay rate of peroxo-diiron complexes (58) on the oxidation potential (Epa) of the [FeII(L1–4)(solvent)3](ClO4)2 (14) complexes.
Figure 6. Peroxo-diiron complexes-mediated disproportionation of H2O2 in CH3CN at 15 °C. (a) Dependence of the reaction rate (Vi) for H2O2 oxidation on the diiron-peroxo complex 58 concentration. [H2O2]0 = 24 mM (b) Dependence of the decay rate of peroxo-diiron complexes (58) on the oxidation potential (Epa) of the [FeII(L1–4)(solvent)3](ClO4)2 (14) complexes.
Molecules 26 04501 g006
Figure 7. (a) Arrhenius plots of logk2 versus 1/T for the FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58)-mediated disproportionation of H2O2 in CH3CN. [58]0 = 2 mM, [H2O2]0 = 24 mM. (b) Plot of ΔG versus lgk2 for the reaction of [FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58) with H2O2.
Figure 7. (a) Arrhenius plots of logk2 versus 1/T for the FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58)-mediated disproportionation of H2O2 in CH3CN. [58]0 = 2 mM, [H2O2]0 = 24 mM. (b) Plot of ΔG versus lgk2 for the reaction of [FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58) with H2O2.
Molecules 26 04501 g007
Figure 8. Isokinetic plot for the for the FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58)-mediated disproportionation of H2O2 in CH3CN.
Figure 8. Isokinetic plot for the for the FeIII2(μ-O)(μ-1,2-O2)(L1–4)2(Solv)2]2+ (58)-mediated disproportionation of H2O2 in CH3CN.
Molecules 26 04501 g008
Scheme 3. Proposed mechanism for the formation of FeIII2(μ-O)(μ-1,2-O2)(L1-L4)2(Solv)2]2+ (58) complexes and their catalase-like reactions.
Scheme 3. Proposed mechanism for the formation of FeIII2(μ-O)(μ-1,2-O2)(L1-L4)2(Solv)2]2+ (58) complexes and their catalase-like reactions.
Molecules 26 04501 sch003
Table 1. Dependence of the yield of peroxo complexes (58) and the evolved oxygen (parantheses) on the number of cycles. [Fe]0 = 2 mM, [H2O2]0 = 24 mM (4×) in CH3CN at 20 °C.
Table 1. Dependence of the yield of peroxo complexes (58) and the evolved oxygen (parantheses) on the number of cycles. [Fe]0 = 2 mM, [H2O2]0 = 24 mM (4×) in CH3CN at 20 °C.
CyclesYield (5) and (O2)/%Yield (6)/%Yield (7)/%Yield (8)/%
I.100 (95)100100100
II.91 (92)874843
III.73 (78)703920
IV.61 (64)533113
V.41 (27)362612
Table 2. Kinetic data for the formation of intermediates 5, 6, 7 and 8 in CH3CN.
Table 2. Kinetic data for the formation of intermediates 5, 6, 7 and 8 in CH3CN.
Nr.Complex T/°C [1–4]0/mM [H2O2]0/M kobs10−2 s−1k2/M−1 s−1
111520.02432.120.87 ± 0.03
211520.03673.150.86 ± 0.04
311520.04894.070.83 ± 0.03
411520.07346.760.92 ± 0.04
511520.09799.040.92 ± 0.03
621520.02434.451.83 ± 0.09
721520.03675.831.58 ± 0.05
821520.04898.441.72 ± 0.06
921520.073410.51.43 ± 0.04
1021520.097914.451.48 ± 0.04
1131520.02439.223.79 ± 0.15
1231520.036712.643.44 ± 0.13
1331520.048916.953.46 ± 0.14
1431520.073424.613.35 ± 0.14
1531520.097931.073.17 ± 0.11
1641520.02439.253.81 ± 0.15
1741520.036713.173.59 ± 0.14
1841520.048918.483.78 ± 0.16
1941520.073427.073.69 ± 0.15
2041520.097935.833.66 ± 0.15
Table 3. Kinetic data for the disproportionation of H2O2 mediated by 5, 6, 7 and 8 in CH3CN.
Table 3. Kinetic data for the disproportionation of H2O2 mediated by 5, 6, 7 and 8 in CH3CN.
Nr.ComplexT/°C[1–4]0/mM[H2O2]0/Mkobs/10−3 s−1k2/10−2 M−1 s−1
11150.7380.02430.3581.47 ± 0.06
21151.480.02430.3871.59 ± 0.07
311520.02430.3701.52 ± 0.06
41520.02430.1160.477 ± 0.02
511020.02430.1600.658 ± 0.03
612020.02430.5232.15 ± 0.11
712520.13.333.33 ± 0.11
812520.25.572.79 ± 0.08
912520.310.13.36 ± 0.09
1012520.617.32.89 ± 0.07
1112520.927.13.01 ± 0.06
121250.30.310.33.42 ± 0.12
131250.60.310.13.38 ± 0.11
1412510.311.13.68 ± 0.14
151251.40.39.473.16 ± 0.13
1611020.31.990.663 ± 0.02
1711520.33.631.21 ± 0.05
1812020.36.122.04 ± 0.08
192150.7380.02430.8233.39 ± 0.11
202151.070.02430.6922.85 ± 0.09
212151.480.02430.7723.18 ± 0.10
222151.770.02430.7443.06 ± 0.09
2321520.02430.7303.00 ± 0.11
242520.02430.3321.37 ± 0.04
2521020.02430.4131.70 ± 0.05
2622020.02431.094.49 ± 0.27
273150.7380.02433.9216.1 ± 0.72
283151.480.02434.7119.4 ± 0.62
2931520.02434.8319.9 ± 0.68
303520.02432.429.96 ± 0.35
3131020.02433.4114.01 ± 0.63
3232020.02438.9036.6 ± 2.01
334150.7380.02435.4922.6 ± 0.79
344151.480.02435.4422.4 ± 0.83
3541520.02435.7523.8 ± 0.85
364520.02432.8911.9 ± 0.48
3741020.02434.2017.3 ± 0.67
3842020.024310.242.01 ± 1.77
Table 4. Rate constants, electrochemical data and activation parameters for the formation of FeIII2(μ-O)(μ-1,2-O2)(L1-L4)2(Solv)2]2+ (5–8) complexes and their catalase-like reactions.
Table 4. Rate constants, electrochemical data and activation parameters for the formation of FeIII2(μ-O)(μ-1,2-O2)(L1-L4)2(Solv)2]2+ (5–8) complexes and their catalase-like reactions.
1 (5)2 (6)3 (7)4 (8)
Epa(FeII/FeIII)/mV vs. Fc/Fc+655683765783
Epc(FeII/FeIII)/(mV vs. Fc/Fc+−90−223−155−143
ΔE/mV745905920925
yield/%851009285
TON ([S]/[Fe])10.412.211.310.4
TOF ([S]/[Fe]/h)288418354234
k2(cat)/10−2 M−1 s−1 (15 °C)1.523.0519.923.8
EA/kJ mol−172.355.957.555.5
ΔS/J mol−1 K−1−37.7−88.1−65.9−71.4
ΔH≠/kJ mol−17053.55553.1
ΔG/kJ mol−18178.97473.7
kdecay/10−2s−1 (15 °C)0.130.151.221.41
kform/M−1 s−1 (15 °C)0.931.433.163.66
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lakk-Bogáth, D.; Török, P.; Csendes, F.V.; Keszei, S.; Gantner, B.; Kaizer, J. Disproportionation of H2O2 Mediated by Diiron-Peroxo Complexes as Catalase Mimics. Molecules 2021, 26, 4501. https://doi.org/10.3390/molecules26154501

AMA Style

Lakk-Bogáth D, Török P, Csendes FV, Keszei S, Gantner B, Kaizer J. Disproportionation of H2O2 Mediated by Diiron-Peroxo Complexes as Catalase Mimics. Molecules. 2021; 26(15):4501. https://doi.org/10.3390/molecules26154501

Chicago/Turabian Style

Lakk-Bogáth, Dóra, Patrik Török, Flóra Viktória Csendes, Soma Keszei, Beatrix Gantner, and József Kaizer. 2021. "Disproportionation of H2O2 Mediated by Diiron-Peroxo Complexes as Catalase Mimics" Molecules 26, no. 15: 4501. https://doi.org/10.3390/molecules26154501

Article Metrics

Back to TopTop