Next Article in Journal
Structurally and Compositionally Tunable Absorption Properties of AgCl@AgAu Nanocatalysts for Plasmonic Photocatalytic Degradation of Environmental Pollutants
Previous Article in Journal
The Effect of AgInS2, SnS, CuS2, Bi2S3 Quantum Dots on the Surface Properties and Photocatalytic Activity of QDs-Sensitized TiO2 Composite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Fine-Tuning Redox Potentials of Manganese(II) Complexes with Isoindoline-Based Ligands: H2O2 Oxidation and Oxidative Bleaching Performance in Aqueous Solution

Department of Chemistry, University of Pannonia, 8201 Veszprém, Hungary
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(4), 404; https://doi.org/10.3390/catal10040404
Submission received: 25 March 2020 / Revised: 6 April 2020 / Accepted: 6 April 2020 / Published: 7 April 2020

Abstract

:
A series of divalent manganese complexes [MII(HL1–6)Cl2] with the 1,3-bis(2’-Ar-imino)isoindolines (HLn, n = 1–6, Ar = pyridyl, 4-methylpyridyl, imidazolyl, thiazolyl, benzimidazolyl and N-methylbenzimidazolyl, respectively) including the previously reported ligands (HL1–2, 4–6) and complexes ([MII(HL1,5)Cl2]) have been prepared and characterized by electrochemical and spectroscopic methods. In these complexes, it was possible to control the redox potential of the metal center by varying the aryl substituent on the bis-iminoisoindoline moiety, and investigate its effect in a catalase-like reaction, and oxidative bleaching process in buffered aqueous solution. The kinetics of the dismutation of H2O2 into H2O and O2, and the oxidative degradation of morin by H2O2 were investigated in buffered water, where the reactivity of the catalysts in both systems was markedly influenced by the redox and Lewis acidic properties of the metal centers and the concentration of the bicarbonate ions. Both the catalase-like and bleaching activity of the catalysts showed a linear correlation with the MnIII/MnII redox potentials. The E1/2 spans a 561 mV range from 388 mV (Ar = benzymidazolyl) to 948 mV (Ar = 4-methylpyridyl) vs. the SCE. The amount of bicarbonate is a critical issue for the in situ formation of peroxycarbonate as a versatile oxidant, and its participation in the formation of high valent MnIV = O species.

Graphical Abstract

1. Introduction

1,3-bis(2’-Ar-imino)isoindolines are pincer type ligands with a tridentate coordination mode and aromatic planarity around the metal centers [1]. They can behave either as monoanionic or protonated ligands during the complexation reaction (Scheme 1). Manganese ion as redox center plays an important role in many oxidoreductases such as manganese-containing ribonucleotide reductase (RNS) from Corynebacterium [2], heme-type peroxidase (MnP) from the Aspergillus species [3,4,5] and non-heme-type binuclear catalases (MnCat) from Lactobacillus plantarum [6,7], Thermus thermophilus [8,9] and Thermoleophilum album [10]. These enzymes are responsible for many types of oxidative processes in biological systems. For example, MnCat enzymes, as important alternatives to heme-type catalases, catalyze the redox disproportionation of hydrogen peroxide into dioxygen and water, protecting the living organisms from the reactive oxygen species (ROS) induced “oxidative stress” reactions [11,12], while the ribonucleotide reductases provide the DNA precursors, deoxyribonucleotides required for DNA replication and repair in all living organisms [2]. Finally, the MnP can oxidize Mn(II) to Mn(III) and in turn oxidizes the phenolic moieties of lignin and some organic pollutants. MnP has also been shown to promote the peroxidation of unsaturated lipids in the absence of H2O2. After the discovery of lignolytic enzymes such as MnP, the pulp industry had great expectations with regard to developing cell-free enzymatic delignification and/or a bleaching system in its industrial processes [13,14].
Many synthetic biomimetics of catalase and peroxidase are being developed as potential therapeutic agents against cancer, Alzheimer’s aging and inflammatory and heart diseases [15,16,17,18], and as efficient catalytic systems for the degradation of various pollutants and bleaching which can be applied for example in food, dairy and textile industries [19,20,21,22,23,24]. Recently, we synthesized and characterized a series of mononuclear transition metal complexes with [MII(Ln)2] (M = Fe(II) [25], Mn(II) [26,27], Ni(II) [27], Co(II) [28] and Cu(II) [29], n = 1, 2 and 4–6), [FeIII(Ln)Cl2] [30], [CuII(HLn)X2] (X = ClO4, Cl, n = 1, 2 and 4–6)] [29] and [MnII(HL1 and 5)Cl2] [31,32] compositions, and used as biomimetics of superoxide dismutase (SOD) [27], catalase [26,31,32], phenoxazinone synthase, catechol oxidase [33] and dioxygenase [30] enzymes. As a continuity of our research, we synthesized the analogs of [MnII(HL2–4, and 6)Cl2] complexes as catalysts against H2O2, and morin by the use of H2O2 as an oxidant in context with the ligand modification by varying the aryl substituent on the bis-iminoisoindoline moiety and redox chemistry. These systems may serve as functional models for the MnCat enzyme, and furthermore, the title complexes could serve as potential candidates for oxidative delignification and/or bleaching performance.

2. Results and Discussion

2.1. Synthesis and Characterization of [MnII(HLn)Cl2] Complexes

The synthesis of the new 1,3-bis(2’-imidazolyl)isoindoline ligand was carried out according to the literature by the reaction of 1,3-diiminoisoindoline with two equivalents of 2-amino-imidazole sulfate in 1-butanol in the presence of sodium carbonate [1]. It was characterized by electronic, 1H-NMR, 13C-NMR, IR and UV–vis measurements. Reaction of equimolar amounts of 1,3-bis(2’-Ar-imino)isoindoline and MnCl2.4H2O in methanol resulted in the formation of mononuclear pentacoordinate complexes with [MnII(HLn)Cl2] composition (Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6)). In the infrared region, the free and the nondeprotonated isoindoline ligands exhibit intense bands in the 1600–1660 cm−1 region that can be assigned as coupled nonspecific νC=N vibrations which can be explained by the two different endocyclic and exocyclic position of the amino groups, and one additional weak absorption around 3300 cm−1, characteristic of a nondeprotonated (νNH) isoindoline ligand [1]. Absorption maxima in the UV–vis spectra of [MnII(HL1–6)Cl2] that are found between 350 and 500 nm can be assigned to the π–π* transitions of the coordinated neutral ligands, except for the lowest energy bands in the range of 450–600 nm, which can be attributed to charge transfer transitions from the manganese(II) ion to a ligand π* orbital (LUMO, MLCT; Figure 1). Deprotonation of the NH group of the isoindoline moiety could be reduced the difference in energy between occupied and unoccupied π-molecular orbitals of the ligand resulting in a red shift of π–π* transitions. In the absence of the significant (>40 nm) bathochromic shift in the three lowest energy bands, the anionic binding of the ligands can be excluded in all cases [1].
Redox properties of the model complexes [MnII(HL1–6)Cl2] were also examined in cyclic voltammetry (CV) experiments in DMF (Figure 2). Electrochemical data for the complexes are listed in Table 1 in mV at 100 mV/s scan rate. The E1/2 spans a 561 mV range from 388 mV (Ar = benzymidazolyl) to 948 mV (Ar = 4-methylpyridyl) vs. the SCE. A narrower range (400 to 600 mV versus SCE) with lower potentials was observed for the recently studied MnII(L)2 complexes. Compound [MnII(HL3)Cl2] (3) undergoes an irreversible transition in the observed potential range in contrast to the quasi-reversible transition of 12 and 46. It can be concluded that the annulation of the imidazolyl group results in a shift (~400 mV) in the MnIII/MnII potentials and stabilizes both the reduced and oxidized forms of the transition (see 5 and 6 in Figure 2). Surprisingly, the applied ligand affects not only the formal potential, but the peak separations (ΔEp = Epa−Epc), too. Significantly larger values (120–140 mV) were observed for the [MnII(HL1–2)Cl2] complexes with 6-membered pyridyl side chains than those found for [MnII(HL4–6)Cl2] with five-membered N-donor thiazolyl and benzimidazolyl containing side chains.
More importantly, a linear correlation was found between the energy of the π–π* (charge transfer (CT)) absorption (and ligand-to-metal charge transfer (LMCT)) band (ν1 and ν2 = λmax−1, respectively) and the oxidation potential, Epa of the manganese center of the [MnII(HL1–6)Cl2] complexes (Figure 3A,B), indicating that the observed shift in the p–p* and LMCT absorption bands can be assigned indirectly to the electronic effect of the ligands (Figure 3). This feature can be attributed to the oxidation of the metal center (MnII/MnIII) being sensitive to the isoindoline ligand used in this study.

2.2. Catalase-Like Reactivity of [MnII(HL1–7)Cl2] Complexes in Aqueous Solution

It was found earlier that the manganese-catalyzed disproportionation of H2O2 at physiological pH occurs only in the bicarbonate (HCO3) containing buffer solution [34]. Recent work in our group has found that complex [MnII(HL1)Cl2] in bicarbonate buffer solution catalyzes the disproportionation of H2O2 into water and O2, and its reactivity increases with the increasing pH and goes through a maximum (pH ~9.6). Under this condition, the reaction rate is directly proportional to the concentration of Mn(II) and shows Michaelis–Menten-type saturation kinetics on [H2O2]0 (Vmax = 8.1 × 10−3 Ms−1, KM = 489 mM, kcat = 38 ± 2 s−1 and k2(kcat/KM) = 79 ± 4 M−1s−1) at pH 9.5 by the use of carbonate buffer [32]. Thus, we attempted to investigate the effect of the HCO3 and prepare new isoindoline-based manganese catalysts by introducing different side chains at the imine functions in order to elucidate the role of electronic factors, and obtain a new catalyst with an increase of its stability and reactivity in the catalase-like reaction.
Catalase activity was carried out volumetrically via the measurements of dioxygen evolution at 20 °C in bicarbonate buffers at pH 9.6. Experiments by Stadtman [34] show that the bicarbonate concentration increases with increasing pH and the effect of pH on the manganese-dependent disproportionation of H2O2 can be assigned to the bicarbonate content alone. Similarly to the previously published results, the initial rate of the reaction (V0) is a linear function of the bicarbonate concentration, suggesting that one equivalent of HCO3 is coordinated to the manganese center during the formation of the catalytically active complex. The first-order dependence with respect to the concentration of bicarbonate is shown in Figure 4. It can be seen that a 10-fold increase in bicarbonate concentration results in a two-fold increase in the rate of H2O2 disproportionation. The increase in the reaction rate may be explained by the formation of the versatile oxidizing agent, HCO4, from the reaction of H2O2 with HCO3. Earlier studies led to the proposition that the dismutation of hydrogen peroxide undergoes redox cycling between Mn(II), Mn(III) and Mn(IV) species [19,35]. The results of previous kinetic studies, as well as the results described here, may suggest both the formation of high valent MnIVO species in Route 1 (Scheme 2) and carbonate radical in Fenton-type chemistry in Route 2 (Scheme 2, [34]) as key catalytic intermediates which may lead to the dismutation of H2O2. We believe that bicarbonate ions act upon the redox potential of an Mn(II) complex.
In the next step, we investigated the effect of the ligand modification by varying the aryl substituent on the bis-iminoisoindoline moiety with emphasis on the redox potential and Lewis acidity of the catalyst that may serve as guidelines for the synthesis of more active and more selective catalysts. The six manganese complexes, [MnII(HL1–6)Cl2] and the MnCl2 salt have been compared to investigate the effect of the various aryl substituents (Figure 5 and Figure 6). Figure 5 shows the results of the catalase activity for the [MnII(HL1–6)Cl2] complexes and demonstrates significant differences based on the number of H2O2 molecules (TON = turnover number) disproportionate by one molecule of the complex after four minutes (240 s). The turnover frequency (TOF = mol H2O2/mol catalyst/h) values, which present the ratios of initial rates (−d[H2O2]/dt) and concentrations of catalysts, are given in Table 2. It was found that complex [MnII(HL2)Cl2] with nonannulated 4-methylpyridyl side chains is the most efficient catalyst with the fastest rate observed at 0.682 × 10−3 Ms−1 and approximately 6.5 (TOF) molecules of H2O2 broken down per second at the fastest rate of activity (V0), while complex [MnII(HL5)Cl2] with annulated benzimidazolyl side chains is a less efficient catalase mimic when compared to complex 2 with the fastest rate of 0.187 × 10−3 Ms−1 and a TOF of 1.77 s−1 for H2O2. Based on these results the lower activity may be due to the annulated aromatic side chains within the ligand system which may prevent access of H2O2 to the manganese center (steric effect) and/or the unfavorable redox and Lewis acidic properties of the catalyst.
Since the redox potential of the catalysts can be measured under the same conditions, it can be used as an excellent reactivity descriptor. By the use of these data, we have clear evidence that the activity (V0) increases almost linearly with the redox potential (Epa for MnIII/MnII) of the catalyst (Figure 6). This finding also suggests that the redox potential of the catalysts acts as the driving force of the reaction, and the dismutation would most likely involve the Mn(II)/Mn(IV)O redox couple (Scheme 2), similarly to what was proposed for the peroxynitrite reductase activity of the manganese porphyrin system [36,37,38,39,40]. The redox potential values of the MnIII/MnII redox couple describe the propensity of the manganese(II) complexes to react with the nucleophilic HO2 and/or HCO4; where the more electron-deficient metal center has a much higher affinity for HO2 and/or HCO4 binding. In summary, the higher the redox potentials of the MnIII/MnII redox couple, the higher is the catalase-like activity, logV0. [MnII(HL1–2)Cl2] (12) complexes with pyridyl side chains, whose metal sites are electron-deficient and faster in catalyzing H2O2 disproportionation than electron-rich derivatives [MnII(HL5–6)Cl2] (56) with more Lewis basic benzimidazolyl side chains.

2.3. Oxidative Degradation of Morin: [MnII(HL1–6)Cl2] Complexes as Bleaching Catalysts

Several manganese complexes with terpyridine, Schiff base, 1,4,7-triazacyclononane and N-methylpropanoate (N-propanoate)-N,N-bis-(2-pyridylmethyl)amine [41] ligands have been tested earlier as bleaching catalysts, where generally high-valent oxo species, LMnIV(O) or LMnV(O), have been proposed as key oxidants [42]. In order to evaluate the bleaching potential of [MnII(HL1–6)Cl2] complexes, we investigated the oxidation of morin (2’,3,4’,5,7-pentahydroflavone), which can be considered as a good model compound for bleaching stain. Experiments were carried out at 25 °C with 1.6 μM catalyst, and the oxidation of morin was followed as the decrease in absorbance at 410 nm (Scheme 3, Figure 7A).
Under this condition as described previously for terpyridine manganese complexes [22], the best activity was observed for [MnII(HL2)Cl2] (2) where the bleaching of morin was completed within five minutes with approximately 20 catalytic cycles per minute (Figure 7B). From our detailed kinetic measurements, the rate of morin decomposition is described by the relationship –d[morin]/dt = V = kox[1–6][H2O2][HCO3][morin], where kox = 7.79 × 106 M−3s−1 for 1 and 0.675 × 106 M−3s−1 for 5 (Figure 8, Table 3 and Table 4). The catalytic activity of the 4-methylpyridyl containing manganese complex (2) was at least 10–12 times higher than that of [MnII(HL5)Cl2] (5) with benzimidazolyl side chains. At low substrate concentration, the reaction is first-order on all reactants; H2O2 (Figure 8A), HCO3¯ (Figure 8B), [MnII(HL1)Cl2] (Figure 8C) and morin (Figure 8D). It is worth noting that the bicarbonate concentration similarly to our previous results plays an important role in the bleaching process, probably during the formation of the catalytically active MnIV(O) species (Scheme 2), which can easily oxidize the morin in an intermolecular or intramolecular manner. In separate experiments under air without H2O2, we found clear evidence for the complexation (~ 4–500 nm) and for the slow base-catalyzed oxidation of morin (Figure 9) [43,44,45,46]. Based on these results, the intramolecular oxidation of the coordinated morin cannot be excluded.
The activity of the isoindolin complexes increases significantly in the order of [Mn(HL5)Cl2] (5) < [Mn(HL6)Cl2] (6) < [Mn(HL4)Cl2] (4) < [Mn(HL3)Cl2] (3) < [Mn(HL1)Cl2] (1) < [Mn(HL2)Cl2] (2). Hence, by changing the more Lewis basic five-membered benzimidazolyl rings to six-membered pyridyl pendant arms, the catalytic activity can be remarkably enhanced. Furthermore, the calculated kox for 16 correlate with the observed oxidation potentials, Epa (MnIII/MnII), giving further evidence that the activity of the catalyst can be controlled by the modification of electron donor properties of the ligand (Figure 10). A similar correlation has been described previously for our catalase-like system (Figure 6B), suggesting that the same oxidant (MnIV(O)) is responsible for the decomposition of H2O2 and morin.

3. Materials and Methods

3.1. Materials and Methods

The ligands 1,3-bis(2′-pyridylimino)isoindoline (HL1), 1,3-bis(4′-methyl-2′-pyridylimino)isoindoline (HL2), 1,3-bis(2′-thiazolylimino)isoindoline (HL4), 1,3-bis(2′-benzimidazolylimino)isoindoline (HL5) and 1,3-bis(2’-N-methyl-benzimidazolylimino)isoindoline (HL6), and the complexes [MnII(HL1 and 6)Cl2] were synthesized according to published procedures [1]. 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 Kft. (Budapest, Hungary).
The UV–visible spectra were recorded on an Agilent 8453 diode-array spectrophotometer (Agilent Technologies, Hewlett-Packard-Strasse 8, Waldbronn, Germany) 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. The NMR spectrum was recorded on a Bruker Avance 400 spectrometer (Bruker Biospin AG, Fällanden, Switzerland). Elemental analysis was done by the Microanalytical Service of the University of Pannonia. Cyclic voltammograms (CV) were taken on a Volta Lab 10 potentiostat with Volta Master 4 software for data processing. The electrodes were as follows: glassy carbon (working), Pt wire (auxiliary) and Ag/AgCl with 3M KCl (reference). The potentials were referenced vs. the ferrocenium/ferrocene (Fc+/Fc) redox couple.

3.2. Syntheses and Characterization

3.2.1. 1,3-Bis(2’-imidazolyl)isoindoline (HL3)

A mixture of 10.81 mmol (1.57 g) of 1,3-diiminoisoindoline and 22.70 mmol (3.00 g) of 2-amino-imidazole sulfate in 25 mL of 1-butanol with sodium carbonate. The solution was heated with stirring at reflux for 20 h. The reaction mixture was cooled, filtered and the solid part obtained was washed with distilled water and cold methanol. The crude product was recrystallized from methanol to yield 0.919 g (30%) of brownish-red crystals. UV–vis [dmf], [λmax, nm logε/dm3 mol−1 cm−1], 356 (3.84), 391 (3.83), 415 (3.84), 443 (3.67), FT–IR bands (KBr pellet cm−1): 3289 w, 3215 w, 3156 w, 3107 w, 2872 w, 1657 s, 1613 s, 1567 s, 1499 m, 1452 m, 1382 w, 1324 w, 1274 s, 1160 m, 1033 m, 753 m, 689 m, 641 m, 574 w, 517 w. Anal Calcd for C40H37F6FeN6O10S2: C, 48.25; H, 3.75; N, 8.45. Found: C, 48.22; H, 3.72; N, 8.47. 1H-NMR (DMSO-d6), δ (ppm): 5.75 (s, 1 H); 7.10 (m, 4H); 7.70 (m, 2H); 7.90 (m, 2H); 12.50 (s, 2H). 13C-NMR (DMSO-d6), δ (ppm): 121.9 (2C); 123.5; 130.3; 131.9; 132.3; 134; 134.5; 136.3; 149.2 (2C); 150.2 (2C); 167.5.

3.2.2. [MnII(HL3)Cl2] (3)

A solution of 0.14 g (0.72 mmol) of MnCl2 4H2O in 2.5 cm3 CH3OH was added to a suspension of 0.20 g (0.72 mmol) LH3 in 2.5 cm3 CH3CN and the brown suspension was refluxed for 6 h. The solvent was removed by evaporation and the crude product was washed with cold CH3OH and diethyl ether, and then dried under vacuum (0.16 g, 53%). UV–vis [dmf], [λmax, nm logε/dm3 mol−1 cm−1], 382 (4.04), 403sh (3.85), 430sh (3.49), 460 (3.25), FT–IR bands (KBr pellet cm−1): 3382 w, 3253 w, 3100 w, 2920 w, 2847 w, 1657 s, 1614 s, 1469 m, 1361 w, 1311 w, 1254 w, 1092 m, 1048 m, 780 m, 709 m, 694 m, 530 w. Anal Calcd for C14H11Cl2MnN7: C, 41.71; H, 2.75; N, 24.32. Found: C, 41.66; H, 2.72; N, 24.35.

3.2.3. [MnII(HL2)Cl2] (2)

Yield: 75%. UV–vis [dmf] [λmax, nm logε/dm3 mol−1 cm−1]: 227 (4.26), 296 (4.22), 330 (4.19), 347 (4.21), 367 (4.28), 386 (4.37), 409 (4.14), 453 (3.32), FT–IR bands (KBr pellet cm−1): 3444 w, 3239 w, 3039 w, 2953 w, 2847 w, 1654 s, 1634 s, 1597 m, 1517 m, 1491 s, 1356 w, 1209 m, 1066 m, 939 m, 829 w, 719 m, 453 m. Anal Calcd for C20H17Cl2MnN5: C, 53.00; H, 3.78; N, 15.45. Found: C, 53.02; H, 3.80; N, 15.48.

3.2.4. [MnII(HL4)Cl2] (4)

Yield: 70%. UV–vis [dmf] [λmax, nm logε/dm3 mol−1 cm−1]: 287 (4.22), 373 (4.29), 396 (4.39), 419 (4.44), 448 (4.22), FT–IR bands (KBr pellet cm−1): 3423 w, 3199 w, 3105 w, 3084 w, 1656 s, 1622 s, 1504 s, 1364 m, 1291 m, 1213 m, 1123 m, 1099 m, 1052 m, 874 m, 772 m, 702 m, 624 w, 526 w. Anal Calcd for C14H9Cl2MnN5S2: C, 38.46; H, 2.07; N, 16.02. Found: C, 38.45; H, 2.05; N, 16.03.

3.2.5. [MnII(HL6)Cl2] (6)

Yield: 75%. UV–vis [dmf] [λmax, nm logε/dm3 mol−1 cm−1]: 371 (4.02), 382 (4.00), 420 (3.91), 448 (3.96), 482sh (3.75), 535 (2.93), FT–IR bands (KBr pellet cm−1): 3427 w, 3043 w, 2925 w, 1629 s, 1613 s, 1552 s, 1499 m, 1475 m, 1290 m, 1180 m, 1090 m, 1066 m, 735 s, 706 m, 543 w. Anal Calcd for C24H19Cl2MnN7: C, 54.26; H, 3.60; N, 18.45. Found: C, 54.24; H, 3.57; N, 18.43.

3.3. Test Reactions of the Catalase-Like Activity

All reactions were carried out at 20 °C in a 30 cm3 reactor containing a stirring bar under air. The stoichiometry of the reaction was measured by the simultaneous determination of the amount of O2 and H2O2 concentrations. The evolved dioxygen was measured volumetrically. In a typical experiment, the appropriate aqueous solutions (19 cm3 0.025 M Na2B4O7.10H2O/0.1 M HCl pH 9.5; or 0.05 M NaHCO3/0.1 M KOH pH 9.5 buffer and I = 0.15 M KNO3) was added to the complex (final concentration is 0.211 mM) dissolved in 1 cm3 DMF, and the flask was closed with a rubber septum. Hydrogen peroxide (final concentration is 0.447 M) was injected through the septum with a syringe. The reactor was connected to a graduated burette filled with oil and dioxygen evolution was measured volumetrically at time intervals of 15 s. Observed initial rates were expressed as mol/dm3 s1 by taking the volume of the solution (20 cm3) into account and calculated from the maximum slope of the curve describing the evolution of O2 versus time.

3.4. Bleaching of Morin

Oxidative degradation of morin in the presence of [MnII(HL1–6)Cl2] complexes was measured in 3 mL optical quartz cells. In a typical experiment, the cuvette was filled with a 1.5 mL buffer solution containing 0.16 mM morin, 10 mM H2O2 and 1.6 μM catalyst, and the bleaching reaction was followed by measuring the decrease in the absorption maximum of morin at 410 nm with 10 s intervals at 25 °C.

4. Conclusions

Efforts have been made to work out a highly efficient and highly selective manganese-based catalytic system for the disproportionation reaction of H2O2 as synthetic catalase mimics and for the oxidation of morin as oxidative bleaching performances. We synthesized six manganese [MII(HL1–6)Cl2] complexes with 1,3-bis(2’-Ar-imino)isoindolines (HLn, n = 1–6, Ar = pyridyl, 4-methylpyridyl, imidazolyl, thiazolyl, benzimidazolyl and N-methylbenzimidazolyl) ligands, and characterized by various electrochemical and spectroscopic methods. In the next step, we investigated the effect of the ligand modification by varying the aryl substituent on the bis-iminoisoindoline moiety with emphasis on the redox potential and Lewis acidity of the catalyst that may serve as guidelines for the synthesis of more active and more selective catalysts toward H2O2, and morin by the use of H2O2 as oxidant. In conclusion, we observed that the higher the redox potentials of the MnIII/MnII redox couple the higher is the catalase-like and bleaching activity. In summary, whose metal sites are electron-deficient are faster in catalyzing H2O2 disproportionation and morin oxidation than electron-rich derivatives with more Lewis basic side chains. It is also worth noting that the bicarbonate concentration plays an important role in both the catalase-like reaction and bleaching process, probably during the formation of the proposed catalytically active MnIV(O) species. These studies may lead to the development of more active oxidation catalysts.

Author Contributions

Resources, B.I.M.; writing—original draft preparation, supervision, J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Financial support of the Hungarian National Research Fund (OTKA K108489), and GINOP-2.3.2-15-2016-00049 are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Csonka, R.; Speier, G.; Kaizer, J. Isoindoline-derived ligands and applications. RSC Adv. 2015, 5, 18401–18419. [Google Scholar] [CrossRef]
  2. Oehlmann, W.; Auling, G. Ribonucleotid reductase (RNR) of Corynebacterium glutamicum ATCC 13032-genetic characterization of a second class IV enzyme. Microbiology 1999, 145, 1595–1604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Glenn, J.K.; Gold, M.H. Purification and characterization of an extracellular Mn(II)-dependent peroxidase from the lignin-degrading basidiomycete, Phanerochaete chrysosporium. Arch. Biochem. Biophys. 1985, 242, 329–341. [Google Scholar] [CrossRef]
  4. Hofrichter, M. Review: Lignin conversion by manganese peroxidase (MnP). Enzym. Microb. Technol. 2002, 30, 454–466. [Google Scholar] [CrossRef]
  5. Banci, L.; Bertini, I.; Dal Pozzo, L.; Del Conte, R.; Tien, M. Monitoring the role of oxalate in manganese peroxidase. Biochemistry 1998, 37, 9009–9015. [Google Scholar] [CrossRef] [PubMed]
  6. Kono, Y.; Fridovich, I. Isolation and characterization of the pseudocatalase of Lactobacillus plantarum. J. Biol. Chem. 1983, 258, 6015–6019. [Google Scholar] [PubMed]
  7. 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]
  8. 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]
  9. 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]
  10. 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] [Green Version]
  11. Beyer, W.F.; Fridovich, I. Catalases-with and without heme. Basic Life Sci. 1988, 49, 651–661. [Google Scholar] [PubMed]
  12. Nicholls, P.; Fita, I.; Loewen, P.C. Enzymology and structure of catalases. Adv. Inorg. Chem. 2001, 51, 51–106. [Google Scholar]
  13. Call, H.P.; Mijcke, I. History, overview and application of mediated lignolytic systems, especially laccase-mediator-systems (LignozymR-process). J. Biotechnol. 1997, 53, 163–202. [Google Scholar] [CrossRef]
  14. Bourbonnais, R.; Paice, M.G. Oxidation of non-phenolic substrates, an expanded role for laccase in lignin biodegradation. FEBS Lett. 1990, 267, 99–102. [Google Scholar] [CrossRef] [Green Version]
  15. 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]
  16. 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]
  17. Boelrijk, A.E.M.; Dismukes, G.C. Mechanism of hydrogen peroxide dismutation by a dimanganese catalase mimic:  Dominant role of an intramolecular base on substrate binding affinity and rate acceleration. Inorg. Chem. 2000, 39, 3020. [Google Scholar] [CrossRef]
  18. 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] [Green Version]
  19. Ember, E.; Gazzaz, H.A.; Rothbart, S.; Puchta, R.; van Eldik, R. MnII–A fascinating oxidation catalyst: Mechanistic insight into the catalyzed oxidative degradation of organic dyes by H2O2. Appl. Catal. B Environ. 2010, 95, 179–191. [Google Scholar] [CrossRef]
  20. Hage, R.; Lienke, A. Bleach and oxidation catalysis by manganese-1,4,7-triazacyclononane complexes and hydrogen peroxide. J. Mol. Catal. A Chem. 2006, 251, 150–158. [Google Scholar] [CrossRef]
  21. Sorokin, A.B.; Kudrik, E.V. Phthalocyanine metal complexes: Versatile catalysts for selective oxidation and bleaching. Catal. Today 2011, 159, 37–46. [Google Scholar] [CrossRef]
  22. Wieprecht, T.; Xia, J.; Heinz, U.; Dannacher, J.; Schlingloff, G. Novel terpyridine-manganese(II) complexes and their potential to activate hydrogen peroxide. J. Mol. Catal. A Chem. 2003, 203, 113–128. [Google Scholar] [CrossRef]
  23. Sen, P.; Yildiz, S.Z. The investigation of oxidative bleaching performance of peripherally Schiff base substituted tri-nuclear cobalt-phthalocyanine complexes. Inorg. Chim. Acta 2017, 462, 30–39. [Google Scholar] [CrossRef]
  24. Sen, P.; Yildirim, E.; Yildiz, S.Z. New alkaline media-soluble functional zinc(II) phthalocyanines bearing poly(hydroxylmethyl)iminomethane Schiff base complexes in catalytic bleaching. Synth. Met. 2016, 215, 41–49. [Google Scholar] [CrossRef]
  25. Kripli, B.; Baráth, G.; Balogh-Hergovich, É.; Giorgi, M.; Simaan, A.J.; Párkányi, L.; Pap, J.S.; Kaizer, J.; Speier, G. Correlation between the SOD-like activity of hexacoordinate iron(II) complexes and their Fe3+/Fe2+ redox potentials. Inorg. Chem. Commun. 2011, 14, 205–209. [Google Scholar] [CrossRef]
  26. 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]
  27. Pap, J.S.; Kripli, B.; Váradi, T.; Giorgi, M.; Kaizer, J.; Speier, G. Comparison of the SOD-like activity of hexacoordinate Mn(II), Fe(II) and Ni(II) complexes having isoindoline-based ligands. J. Inorg. Biochem. 2011, 105, 911–918. [Google Scholar] [CrossRef]
  28. Pap, J.S.; Kripli, B.; Giorgi, M.; Kaizer, J.; Speier, G. Redox properties of cobalt(II) complexes with isoindoline-based ligands. Transit. Met. Chem. 2011, 36, 481–487. [Google Scholar] [CrossRef]
  29. Pap, J.S.; Kripli, B.; Bányai, V.; Giorgi, M.; Korecz, L.; Gajda, T.; Árus, D.; Kaizer, J.; Speier, G. Tetra-, penta- and hexacoordinate copper(II) complexes with N3 donor isoindoline-based ligands: Characterization and SOD-like activity. Inorg. Chim. Acta 2011, 376, 158–169. [Google Scholar] [CrossRef]
  30. Váradi, T.; Pap, J.S.; Giorgi, M.; Párkányi, L.; Csay, T.; Speier, G.; Kaizer, J. Iron(III) complexes with meridional ligands as functional models of intradiol-cleaving catechol dioxygenases. Inorg. Chem. 2013, 52, 1559–1569. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. Kaizer, J.; Baráth, G.; Csonka, R.; Speier, G.; Korecz, L.; Rockenbauer, A.; Párkányi, L. Catechol oxidase and phenoxazinone synthase activity of a manganese(II) isoindoline complex. J. Inorg. Biochem. 2008, 102, 773–780. [Google Scholar] [CrossRef]
  34. 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]
  35. 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. New J. Chem. 2020, 44, 5545–5555. [Google Scholar] [CrossRef]
  36. Batinic-Haberle, I.; Tovmasyan, A.; Roberts, E.R.; Vujaskovic, Z.; Leong, K.W.; Spasojevic, I. SOD therapeutics: Latest insights into their structure-activity relationships and impact on the cellular redox-based signaling pathways. Antioxid. Redox Signal. 2014, 20, 2372–2415. [Google Scholar] [CrossRef] [PubMed]
  37. 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] [PubMed] [Green Version]
  38. Ferrer-Sueta, G.; Batinic-Haberle, I.; Spasojevic, I.; Fridovich, I.; Radi, R. Catalytic scavenging of peroxynitrite by isomeric Mn(III) N-methylpyridylporphyrins in the presence of reductants. Chem. Res. Toxicol. 1999, 12, 442–449. [Google Scholar] [CrossRef]
  39. Ferrer-Sueta, G.; Quijano, C.; Alvarez, B.; Radi, R. Reactions of manganese porphyrins and manganese-superoxide dismutase with peroxynitrite. Methods Enzymol. 2002, 349, 23–37. [Google Scholar] [PubMed]
  40. Ferrer-Sueta, G.; Vitturi, D.; Batinic-Haberle, I.; Fridovich, I.; Goldstein, S.; Czapski, G.; Radi, R. Reactions of manganese porphyrins with peroxynitrite and carbonate radical anion. J. Biol. Chem. 2003, 278, 27432–27438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Pap, J.S.; Kripli, B.; Bors, I.; Bogáth, D.; Giorgi, M.; Kaizer, J.; Speier, G. Transition metal complexes bearing flexible N3 or N3O donor ligands: Reactivity toward superoxide radical anion and hydrogen peroxide. J. Inorg. Biochem. 2012, 117, 60–70. [Google Scholar] [CrossRef] [PubMed]
  42. Hage, R.; Lienke, A. Applications of transition-metal catalysts to textile and wood-pulp bleaching. Angew. Chem. Int. Ed. 2005, 45, 206–222. [Google Scholar] [CrossRef] [PubMed]
  43. Kaizer, J.; Baráth, G.; Pap, J.; Speier, G.; Giorgi, M.; Réglier, M. Manganese and iron flavonolates as flavonol 2,4-dioxygenase mimics. Chem. Commun. 2007, 48, 5235–5237. [Google Scholar] [CrossRef] [PubMed]
  44. Barhács, L.; Kaizer, J.; Speier, G. Kinetics and mechanism of the oxygenation of potassium flavonolate. Evidence for an electron transfer mechanism. J. Org. Chem. 2000, 65, 3449–3452. [Google Scholar] [CrossRef]
  45. Pap, J.S.; Kaizer, J.; Speier, G. Model systems for the CO-releasing flavonol 2,4-dioxygenase enzyme. Coord. Chem. Rev. 2010, 254, 781–793. [Google Scholar] [CrossRef]
  46. Kaizer, J.; Balogh-Hergovich, É.; Czaun, M.; Csay, T.; Speier, G. Redox and nonredox metal assisted model systems with relevance to flavonol and 3-hydroxyquinolin-4(1H)-one 2,4-dioxygenase. Coord. Chem. Rev. 2006, 250, 2222–2233. [Google Scholar] [CrossRef]
Scheme 1. Structure of the 1,3-bis(2’-Ar-imino)isoindoline ligands involved in this study.
Scheme 1. Structure of the 1,3-bis(2’-Ar-imino)isoindoline ligands involved in this study.
Catalysts 10 00404 sch001
Figure 1. Electronic spectra of [MnII(HLn)Cl2] (HL = 1,3-bis(2’-Ar-imino)isoindoline) complexes in DMF solution with nonannulated (Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4)), aromatic side chains (A) and annulated Ar = benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6) aromatic side chains (B).
Figure 1. Electronic spectra of [MnII(HLn)Cl2] (HL = 1,3-bis(2’-Ar-imino)isoindoline) complexes in DMF solution with nonannulated (Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4)), aromatic side chains (A) and annulated Ar = benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6) aromatic side chains (B).
Catalysts 10 00404 g001
Figure 2. Cyclic voltammetry of the [MnII(HLn)Cl2] (HL = 1,3-bis(2’-Ar-imino)isoindoline; Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6)) complexes in DMF at 25 °C.
Figure 2. Cyclic voltammetry of the [MnII(HLn)Cl2] (HL = 1,3-bis(2’-Ar-imino)isoindoline; Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6)) complexes in DMF at 25 °C.
Catalysts 10 00404 g002
Figure 3. Correlation between the oxidation potential of the [MnII(HL1–6)Cl2] complexes and the energy of the π–π* absorption band (A). Correlation of the oxidation potential of the [MnII(HL1–6)Cl2] complexes and the energy of the LMCT absorption band (B) in DMF.
Figure 3. Correlation between the oxidation potential of the [MnII(HL1–6)Cl2] complexes and the energy of the π–π* absorption band (A). Correlation of the oxidation potential of the [MnII(HL1–6)Cl2] complexes and the energy of the LMCT absorption band (B) in DMF.
Catalysts 10 00404 g003
Figure 4. Kinetic analysis for the hydrogen peroxide degradation catalyzed by MnCl2 and [MnII(HL1)Cl2] (1) in bicarbonate buffer (pH = 9.6). (A) Time traces for the reaction of MnCl2 and 1 with H2O2 in the presence of bicarbonate. (B) Dependence of the reaction rate on the bicarbonate concentration for the [MnII(HL1)Cl2]-catalyzed disproportionation of H2O2. Conditions: [1 or MnCl2]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Figure 4. Kinetic analysis for the hydrogen peroxide degradation catalyzed by MnCl2 and [MnII(HL1)Cl2] (1) in bicarbonate buffer (pH = 9.6). (A) Time traces for the reaction of MnCl2 and 1 with H2O2 in the presence of bicarbonate. (B) Dependence of the reaction rate on the bicarbonate concentration for the [MnII(HL1)Cl2]-catalyzed disproportionation of H2O2. Conditions: [1 or MnCl2]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Catalysts 10 00404 g004
Scheme 2. Proposed mechanisms for dismutation of H2O2 by [MnII(HL1–6)Cl2] complexes.
Scheme 2. Proposed mechanisms for dismutation of H2O2 by [MnII(HL1–6)Cl2] complexes.
Catalysts 10 00404 sch002
Figure 5. Comparison of the catalase-like activity of [MnII(HLn)Cl2] (16, HLn = 1,3-bis(2’-Ar-imino)isoindoline, Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6) complexes in bicarbonate buffer (pH = 9.6). Conditions: [Mn]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Figure 5. Comparison of the catalase-like activity of [MnII(HLn)Cl2] (16, HLn = 1,3-bis(2’-Ar-imino)isoindoline, Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6) complexes in bicarbonate buffer (pH = 9.6). Conditions: [Mn]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Catalysts 10 00404 g005
Figure 6. Kinetic analysis for the hydrogen peroxide degradation catalyzed by [MnII(HL1–6)Cl2] (16) complexes in bicarbonate buffer (pH = 9.6). (A) Time traces for the reaction of [MnII(HL1–6)Cl2] (16) complexes with H2O2 in the presence of bicarbonate. (B) Dependence of the reaction rate on the oxidation potential (Epa) of the [MnII(HL1–6)Cl2] complexes for the [MnII(HL1)Cl2]-catalyzed disproportionation of H2O2. Conditions: [1–6]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Figure 6. Kinetic analysis for the hydrogen peroxide degradation catalyzed by [MnII(HL1–6)Cl2] (16) complexes in bicarbonate buffer (pH = 9.6). (A) Time traces for the reaction of [MnII(HL1–6)Cl2] (16) complexes with H2O2 in the presence of bicarbonate. (B) Dependence of the reaction rate on the oxidation potential (Epa) of the [MnII(HL1–6)Cl2] complexes for the [MnII(HL1)Cl2]-catalyzed disproportionation of H2O2. Conditions: [1–6]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Catalysts 10 00404 g006
Scheme 3. Catalytic oxidation of morin by [MnII(HL1–6)Cl2] complexes.
Scheme 3. Catalytic oxidation of morin by [MnII(HL1–6)Cl2] complexes.
Catalysts 10 00404 sch003
Figure 7. (A) Time-dependent UV–vis spectra of a 0.16 mM solution of morin at pH 10 in the presence of 10 mM H2O2 and 1.6 μM [MnII(HL1)Cl2] (1) at 25 °C. (B) Typical kinetic traces of morin oxidation catalyzed by [MnII(HL1–6)Cl2] (16) complexes. Conditions: [morin]0 = 0.16 mM, [H2O2]0 = 10 mM, [1–6]0 = 1.6 μM, pH = 10, at 25 °C.
Figure 7. (A) Time-dependent UV–vis spectra of a 0.16 mM solution of morin at pH 10 in the presence of 10 mM H2O2 and 1.6 μM [MnII(HL1)Cl2] (1) at 25 °C. (B) Typical kinetic traces of morin oxidation catalyzed by [MnII(HL1–6)Cl2] (16) complexes. Conditions: [morin]0 = 0.16 mM, [H2O2]0 = 10 mM, [1–6]0 = 1.6 μM, pH = 10, at 25 °C.
Catalysts 10 00404 g007
Figure 8. Catalytic oxidation of morin with [Mn(HL1)Cl2] (1) in bicarbonate buffer at pH 10 and 25 °C. (A) Dependence of the first-order rate constant (kobs) for morin oxidation on the H2O2 concentration (Table 3). (B) Dependence of the first-order rate constant (kobs) for morin oxidation on the HCO3¯ concentration (Table 3). (C) Dependence of the first-order rate constant (kobs) for morin oxidation on the [Mn(HL1)Cl2] (1) concentration (Table 3). (D) Dependence of the reaction rate (V0) for morin oxidation on the morin concentration (Table 3).
Figure 8. Catalytic oxidation of morin with [Mn(HL1)Cl2] (1) in bicarbonate buffer at pH 10 and 25 °C. (A) Dependence of the first-order rate constant (kobs) for morin oxidation on the H2O2 concentration (Table 3). (B) Dependence of the first-order rate constant (kobs) for morin oxidation on the HCO3¯ concentration (Table 3). (C) Dependence of the first-order rate constant (kobs) for morin oxidation on the [Mn(HL1)Cl2] (1) concentration (Table 3). (D) Dependence of the reaction rate (V0) for morin oxidation on the morin concentration (Table 3).
Catalysts 10 00404 g008
Figure 9. Time-dependent UV–vis spectra of a 0.16 mM solution of morin at pH 10 in the presence of 1.6 μM [MnII(HL1)Cl2] (1) under air at 25 °C without H2O2. (B) Kinetic traces of complexation and oxidation of morin.
Figure 9. Time-dependent UV–vis spectra of a 0.16 mM solution of morin at pH 10 in the presence of 1.6 μM [MnII(HL1)Cl2] (1) under air at 25 °C without H2O2. (B) Kinetic traces of complexation and oxidation of morin.
Catalysts 10 00404 g009
Figure 10. Dependence of the first-order rate constant (kobs) for morin oxidation on the oxidation potential (Epa) of the [MnII(HL1–6)Cl2] complexes in bicarbonate buffer (pH 10). Conditions: [1–6]0 = 1.6 × 10−6 M, [morin]0 = 0.16 × 10−3 M, [H2O2]0 = 0.010 M at 25 °C.
Figure 10. Dependence of the first-order rate constant (kobs) for morin oxidation on the oxidation potential (Epa) of the [MnII(HL1–6)Cl2] complexes in bicarbonate buffer (pH 10). Conditions: [1–6]0 = 1.6 × 10−6 M, [morin]0 = 0.16 × 10−3 M, [H2O2]0 = 0.010 M at 25 °C.
Catalysts 10 00404 g010
Table 1. Electrochemical and spectroscopic data for [MnII(HL1–6)Cl2] complexes.
Table 1. Electrochemical and spectroscopic data for [MnII(HL1–6)Cl2] complexes.
CatalystEpa
(mV)
Epc
(mV)
E01/2
(mV vs SCE)
ΔEp = Epa-Epc
(mV)
λmax (π–π*)1
(nm)
ν1max−1) 1
(104cm−1)
ν2max−1) 2
(104cm−1)
[Mn(HL1)Cl2] (1)9878659261223862.5912.174
[Mn(HL2)Cl2] (2)10168809481363662.732.188
[Mn(HL3)Cl2] (3)816--1313912.562.128
[Mn(HL4)Cl2] (4)625573600524192.392.028
[Mn(HL5)Cl2] (5)421354388674552.201.859
[Mn(HL6)Cl2] (6)455395425604482.231.894
1 Charge transfer (CT), 2 ligand-to-metal charge transfer (LMCT) bands.
Table 2. Comparison of the catalase-like activity of [MnII(HLn)Cl2] (16) complexes in bicarbonate buffer (pH = 9.6; HLn = 1,3-bis(2’-Ar-imino)isoindoline, Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6)).
Table 2. Comparison of the catalase-like activity of [MnII(HLn)Cl2] (16) complexes in bicarbonate buffer (pH = 9.6; HLn = 1,3-bis(2’-Ar-imino)isoindoline, Ar = pyridyl (n = 1), 4-methylpyridyl (n = 2), imidazolyl (n = 3), thiazolyl (n = 4), benzimidazolyl (n = 5) and N-methylbenzimidazolyl (n = 6)).
Catalyst 1E0pa
(mV)
E0pc
(mV)
E01/2 vs SCE
(mV)
Yield
(%)
TONV0
(10−3Ms−1)
TOF
(h−1)
[Mn(HL1)Cl2] (1)98786592632.66920.56919416
[Mn(HL2)Cl2] (2)101688094836.67750.68223272
[Mn(HL3)Cl2] (3)81668575027.15740.42213820
[Mn(HL4)Cl2] (4)62557360021.84630.32812012
[Mn(HL5)Cl2] (5)42135438816.93200.1876382
[Mn(HL6)Cl2] (6)45539542515.133570.2046962
MnCl2/HCO3---8.881880.0811382
1 Conditions: [Mn]0 = 2.11 × 10−4 M, [H2O2]0 = 4.47 × 10−1 M at 20 °C.
Table 3. Summary of kinetic data for the catalytic oxidation of morin with [Mn(HL1)Cl2] (1) in bicarbonate buffer at pH 10 and 25 °C.
Table 3. Summary of kinetic data for the catalytic oxidation of morin with [Mn(HL1)Cl2] (1) in bicarbonate buffer at pH 10 and 25 °C.
Catalyst (1)
(10−6M)
[H2O2]
(10−3M)
[HCO3]
(10−3M)
[Morin]
(10−3M)
kobs
(10−3s−1)
kox
(106M−3s−1)
0.6210500.161.63 ± 0.065.26 ± 0.2
1.610500.164.19 ± 0.165.24 ± 0.2
2.510500.166.73 ± 0.375.38 ± 0.3
0.6210500.161.63 ± 0.065.33 ± 0.2
0.627.5500.161.22 ± 0.055.25 ± 0.2
0.625.0500.160.78 ± 0.025.12 ± 0.1
0.622.5500.160.41 ± 0.015.29 ± 0.2
0.6210500.161.63 ± 0.065.26 ± 0.2
0.62101000.163.02 ± 0.065.24 ± 0.1
0.62102000.167.01 ± 0.135.24 ± 0.1
0.62103000.1610.5 ± 0.65.37 ± 0.3
0.6210500.161.63 ± 0.065.26 ± 0.2
0.6210500.121.65 ± 0.035.32 ± 0.1
0.6210500.081.70 ± 0.065.48 ± 0.2
0.6210500.041.73 ± 0.065.5 ± 0.2
Table 4. Summary of kinetic data for the catalytic oxidation of morin with [Mn(HL1–6)Cl2] (16) in bicarbonate buffer at pH 10 and 25 °C.
Table 4. Summary of kinetic data for the catalytic oxidation of morin with [Mn(HL1–6)Cl2] (16) in bicarbonate buffer at pH 10 and 25 °C.
Catalyst 1E0pa
(mV)
E0pc
(mV)
E01/2vs SCE
(mV)
kobs
(10−3 s−1)
kox
(106 M−3 s−1)
[Mn(HL1)Cl2] (1)9878659264.194 ± 0.1265.241 ± 0.161
[Mn(HL2)Cl2] (2)10168809486.230 ± 0.1567.790 ± 0.192
[Mn(HL3)Cl2] (3)8166857502.171 ± 0.0862.713 ± 0.108
[Mn(HL4)Cl2] (4)6255736001.101 ± 0.0221.376 ± 0.028
[Mn(HL5)Cl2] (5)4213543880.541 ± 0.0150.676 ± 0.018
[Mn(HL6)Cl2] (6)4553954250.780 ± 0.0210.975 ± 0.026
- 0.0076 ± 0.0002
1 Conditions: [morin]0 = 0.16 mM, [H2O2]0 = 10 mM, [1–6]0 = 1.6 μM, pH = 10, at 25 °C.

Share and Cite

MDPI and ACS Style

Meena, B.I.; Kaizer, J. Design and Fine-Tuning Redox Potentials of Manganese(II) Complexes with Isoindoline-Based Ligands: H2O2 Oxidation and Oxidative Bleaching Performance in Aqueous Solution. Catalysts 2020, 10, 404. https://doi.org/10.3390/catal10040404

AMA Style

Meena BI, Kaizer J. Design and Fine-Tuning Redox Potentials of Manganese(II) Complexes with Isoindoline-Based Ligands: H2O2 Oxidation and Oxidative Bleaching Performance in Aqueous Solution. Catalysts. 2020; 10(4):404. https://doi.org/10.3390/catal10040404

Chicago/Turabian Style

Meena, Bashdar I., and József Kaizer. 2020. "Design and Fine-Tuning Redox Potentials of Manganese(II) Complexes with Isoindoline-Based Ligands: H2O2 Oxidation and Oxidative Bleaching Performance in Aqueous Solution" Catalysts 10, no. 4: 404. https://doi.org/10.3390/catal10040404

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop