Next Article in Journal
Isolation, Structural Characterization, and Biological Activity of the Two Acidic Polysaccharides from the Fruits of the Elaeagnus angustifolia Linnaeus
Next Article in Special Issue
Direct Hydroxylation of Benzene with Hydrogen Peroxide Using Fe Complexes Encapsulated into Mesoporous Y-Type Zeolite
Previous Article in Journal
An Eco-Friendly Quaternary Ammonium Salt as a Corrosion Inhibitor for Carbon Steel in 5 M HCl Solution: Theoretical and Experimental Investigation
Previous Article in Special Issue
The Steric Effect in Preparations of Vanadium(II)/(III) Dinitrogen Complexes of Triamidoamine Ligands Bearing Bulky Substituents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Conversion of Superoxide to Hydroperoxide on Cobalt(III) Depends on the Structural and Electronic Properties of Azole-Based Chelating Ligands

1
Department of Material and Life Chemistry, Faculty of Engineering, Kanagawa University, 3-27-1 Rokkakubashi, Kanagawa-ku, Yokohama 221-8686, Japan
2
Department of Life Science, University of Hyogo, Ako-gun, Hyogo 678-1297, Japan
*
Author to whom correspondence should be addressed.
Current address: Department of Applied Chemistry, Sanyo-Onoda City University, Yamaguchi 756-0884, Japan.
Passed away on 23 July 2017.
Molecules 2022, 27(19), 6416; https://doi.org/10.3390/molecules27196416
Submission received: 30 August 2022 / Revised: 24 September 2022 / Accepted: 24 September 2022 / Published: 28 September 2022

Abstract

:
Conversion from superoxide (O2) to hydroperoxide (OOH) on the metal center of oxygenases and oxidases is recognized to be a key step to generating an active species for substrate oxidation. In this study, reactivity of cobalt(III)-superoxido complexes supported by facially-capping tridentate tris(3,5-dimethyl-4-X-pyrazolyl)hydroborate ([HB(pzMe2,X)3]; TpMe2,X) and bidentate bis(1-methyl-imidazolyl)methylborate ([B(ImN-Me)2Me(Y)]; LY) ligands toward H-atom donating reagent (2-hydroxy-2-azaadamantane; AZADOL) has been explored. The oxygenation of the cobalt(II) precursors give the corresponding cobalt(III)-superoxido complexes, and the following reaction with AZADOL yield the hydroperoxido species as has been characterized by spectroscopy (UV-vis, resonance Raman, EPR). The reaction of the cobalt(III)-superoxido species and a reducing reagent ([CoII(C5H5)2]; cobaltocene) with proton (trifluoroacetic acid; TFA) also yields the corresponding cobalt(III)-hydroperoxido species. Kinetic analyses of the formation rates of the cobalt(III)-hydroperoxido complexes reveal that second-order rate constants depend on the structural and electronic properties of the cobalt-supporting chelating ligands. An electron-withdrawing ligand opposite to the superoxide accelerates the hydrogen atom transfer (HAT) reaction from AZADOL due to an increase in the electrophilicity of the superoxide ligand. Shielding the cobalt center by the alkyl group on the boron center of bis(imidazolyl)borate ligands hinders the approaching of AZADOL to the superoxide, although the steric effect is insignificant.

1. Introduction

Metal-superoxido (M-O2) and hydroperoxido species (M-OOH) are recognized to be key intermediates in biological oxidation reactions mediated by oxygenases and oxidases, which contain transition metal centers because the conversion of metal-bound superoxide ion to hydroperoxide species occurs during the enzymatic reductive O2 activation process. In the catalytic cycle of heme-iron oxygenases cytochrome P-450, an iron(III)-superoxido species reacts with H+ and electron to yield an iron(III)-hydroperoxido complex, which is a precursor for an iron(IV)-oxo species with porphyrin π-cation radical ligand “Compound I”, that is the active species for alkane hydroxylation [1]. During the course of the enzymatic reactions by some non-heme iron and copper oxygenases, it is proposed that the metal-superoxide species abstracts a hydrogen atom from the C-H bond of the substrate to give a metal-hydroperoxide species [2,3]. In the catalytic process of galactose oxidase, copper(II)-hydroperoxido species generates from the copper(II)-superoxido precursor via hydrogen atom abstraction (HAT) reaction from phenol moiety of tyrosine [4]. For the copper monooxygenases, another mechanism of the conversion from copper(II)-superoxido to -hydroperoxido through the reaction with an appropriate reductant, such as ascorbate and proton, similar to that found in cytochrome P-450, is proposed [5]. Therefore, interest in the reactivity of synthetic metal-superoxido complexes toward hydrogen donating reagents giving M-OOH species has been growing [6,7].
We have reported that the formation of non-heme superoxido complexes of iron(III) and cobalt(III), [M(O2)(TpMe2,X)(LY)] where TpMe2,X and LY denote hydrotris(3, 5-dimethyl-4-X-1-pyrazolyl)borate and bis(1-methyl-2-imidazolyl)methyl-Y-borate, respectively, through oxidative addition of O2 to the corresponding penta-coordinated divalent metal precursors. O2 affinity of the cobalt(II) centers is controlled by electronic and structural properties of substituents X and Y incorporated on the fourth position of the pyrazole rings of TpMe2,X and the boron center of LY [8]. We have also revealed that the iron(III)-superoxido complex reacts with a hydrogen-donating reagent 2-hydroxy-2-azaadamantane (AZADOL) at low temperature to give the corresponding iron(III)-hydroperoxido complex. The same iron(III)-hydroperoxido species formation also occurs in the reaction of the superoxido complex with proton and a reducing reagent such as an [Fe(Cp)2] derivative [9]. Thus, we have explored the reactivity of a series of [Co(O2)(TpMe2,X)(LY)] toward the transformation to the corresponding hydroperoxido species (Scheme 1). Especially the ligand effect on the reactivity of the cobalt(III)-superoxido species is interesting. Due to the lower thermal instability of the iron(III)-superoxido species (irreversible decomposition occurred), comparison of the various ligands compounds is difficult. In contrast, the cobalt(III)-superoxido complexes are robust against irreversible oxidative transformations. Therefore, the study of cobalt complexes would facilitate examination of the effects of the ligands.

2. Results

2.1. Formation and Characterization of Cobalt(III)-Hydroperoxo Species

We have previously reported that the iron(III)-superoxido complex, [FeIII(O2)(TpMe2,H)(LPh)], reacts with AZADOL to yield the corresponding hydroperoxido species [FeIII(OOH)(TpMe2,H)(LPh)] [9]. Therefore, at the beginning of this study, the reactivity of the analogous cobalt(III)-superoxido complex with the same ligand set, i.e., [CoIII(O2)(TpMe2,H)(LPh)] (1S), was examined.
The superoxido species 1S was generated by the reaction of the corresponding cobalt(II) precursor, [CoII(TpMe2,H)(LPh)] (1R), with O2 in THF at −60 °C. To the THF solution involving 1S, an excess amount (25 equivalents) of AZADOL was added. A UV-vis spectrum of the resulting solution revealed the increasing absorption around 300 nm with decreasing absorption bands around 390 and 530 nm, as shown in Figure 1. This spectral change might indicate that the superoxide complex 1S was converted to the hydroperoxido species 1H. Electronic spectral patterns were simulated by using the optimized structures obtained by DFT calculation as described below (Figure S1). The superoxido species exhibits intense bands in the UV region and a moderate band in the visible to near IR region. In the hydroperoxido complex, intense bands in the UV region are also observed, but no band can be found around the visible region. These trends are consistent with the observed spectral patterns shown in Figure 1, although the predicted spectra showed blue-shift trend compared to the experimental ones.
A spectral change similar to Figure 1 was observed in the mixing of 1S and [CoII(Cp)2] (cobaltocene) with trifluoroacetic acid (TFA). However, 1S did not react with [CoII(Cp)2] in the absence of trifluoroacetic acid. Moreover, 1S did not react with 1-benzyl-1,4-dihydro-nicotinamide (BNAH), of which the BDE of C–H is comparable to the BDE of O–H of AZADOL. These behaviors are the same as those of the reported iron derivative [FeIII(O2)(TpMe2,H)(LPh)] [9].
The resonance Raman analysis of the generated species carried out by 413 nm excitation supported the assignment of 1H as the hydroperoxido complex. Raman bands appeared at 562 cm−1 and 823 cm−1 on the compound of 16O2-derived 1S with AZADOL. These bands were shifted to 542 cm−1 and 780 cm−1 on the compound derived from 18O2-labeled 1S with AZADOL (Figure 2). The ν(Co–O) and ν(O–O) of 1H were similar to the reported ones of cobalt(III)-OOH species, as summarized in Table 1 [10,11,12,13].
The EPR spectrum of the reaction mixture of 1S and AZADOL exhibited signals attributed to an N-O• radical of 2-oxyl-2-azaadamantane with the remaining 1S (Figure 3). This result also supported that the reaction of 1S with AZADOL yields low-spin cobalt(III)-hydroperoxido complex through the proton-coupled electron transfer.
The kinetics of the reaction of 1S with the excess amount of AZADOL was examined by monitoring the increase of the absorption at 300 nm. The reaction followed pseudo-first-order kinetics. The kinetic isotope effect (KIE; kH/kD = 9.4) was observed in the reaction of 1S with the O-D group-containing AZADOL (AZADOL-d1), as shown in Figure 4. This result indicated the O-H bond cleavage of AZADOL was involved as the rate-determining step in the formation of the cobalt(III)-hydroperoxido complex 1H.
We have also explored the DFT calculations based on the reported X-ray structure of the cobalt(III)-superoxido complex 2S [8,9]. The spin state of the superoxido and hydroperoxido species is set to S = 1/2 (low-spin cobalt(III) + an unpaired electron of O2) and S = 0 (low-spin cobalt(III)), respectively. The resulting optimized molecular structure of the superoxido complex was consistent with the reported X-ray structure (Table 2). The overall molecular structure of the optimized hydroperoxido species was similar to that of the superoxido one. The O-O bond length of the hydroperoxido species elongated to 1.425 Å that falls within the range of the peroxide (Figure 5).

2.2. Ligand Effect on the Formation of the Cobalt(III)-Hydroperoxido Species

Previously, we have revealed that the difference in the O2 affinity of [CoII(TpMe2,X)(LY)] (1R6R) is caused by the difference in the substituent groups introduced into the ligands TpMe2,R and LY. The boron-attached substituent Y in LY changes the size of the space covering the cobalt(II) center, and the order of the O2 affinity in a series of [CoII(TpMe2,H)(LY)] is consistent with the space around the sixth vacant (=O2 binding) site estimated by the solid angle analysis of the X-ray crystal structures. The substituent X attached to the fourth position of the pyrazole ring in TpMe2,X affects the electron density on the cobalt center. The electron-donating methyl substituent increases, whereas the electron-withdrawing bromine substituent decreases the O2 affinity (pressure-based equilibrium constant KO2). The overall order of O2 affinity of [CoII(TpMe2,X)(LY)] is [X:Y] = [Me:Ph] (5R) > [H:Ph] (1R) > [H:OiPr] (2R) > [Br:Ph] (6R) > [H:Me] (3R) > [H:nBu] (4R) [8]. In this study, we investigated the substituent effect on the conversion of the cobalt(III) superoxido to hydroperoxo species.
The kinetics of the formation of cobalt(III) hydroperoxo species having different ligands were examined. At −60 °C to the superoxido species 1S6S, which were generated by the bubbling of O2 to THF solutions of the cobalt(II) precursors 1R6R, THF solution of AZADOL with various concentrations was added. In all cases, the observed change of the UV-vis spectrum was almost the same as that observed in the above-mentioned reaction of 1S with AZADOL. Therefore, pseudo-first-order reaction rate analysis was performed by monitoring the increase in absorption intensity at 300 nm. As a result, a linear dependence between the pseudo-first-order rate constant kobs and the concentration of AZADOL was observed, from which the second-order rate constant k2 was estimated, as summarized in Table 3.
In a series of TpMe2,H complexes, the order of k2 was 1H > 2H > 3H > 4H (Figure 6). This order was consistent with that of the O2 affinity of the corresponding cobalt(II) precursor attributed to the structural effect derived from LY. As we have reported, the order of KO2 and the Co(II)/Co(III) oxidation potential of the cobalt(II) complexes 1R4R is not consistent [8]. This result suggests that the steric effect of the boron-attached substituent affects the accessibility of AZADOL to the superoxide ligand.
Interestingly, the order of k2 depending on the pyrazoles’ fourth substituent X was 6H > 5H > 1H (Figure 7). In the series of the cobalt(II) complexes with LPh (1R, 5R and 6R), the orders of the O2 affinity (5R > 1R > 6R) and the Co(II)/Co(III) oxidation potential (6R > 1R > 5R) are inverse, and these orders are not consistent the order of k2 of the hydroperoxido species. The largest k2 of 6H may be due to the electron-withdrawing effect of the Br group on TpMe2,Br, which reduces the electron density of the superoxide ligand in 6S and increases its electrophilicity, accelerating the HAT from AZADOL. The electron-donating Me group containing TpMe2,Me complex 5H exhibited a somewhat larger k2 value compared to the TpMe2,H complex 1H. This result seems to be correlated with the affinity of the superoxide toward the protic hydrogen atom in AZADOL. The accelerating effect of both electron-withdrawing and electron-donating ligands on the conversion from the cobalt(III)-superoxido to hydroperoxido species might suggest that this process proceeds through a proton-coupled electron transfer reaction.

3. Discussion

The cobalt(III)-superoxido complexes supported by facially-capping tridentate tris(pyrazolyl)borate and bidentate bis(imidazolyl)borate ligands were converted to the corresponding cobalt(III)-hydroperoxido species by the reaction with the mild H-atom donating reagent AZADOL. Similar reactivity of cobalt(III)-superoxido species toward N-hydroxy group involving compound TEMPO-H have been reported [11,14]. The conversion of superoxido to hydroperoxido species through the reaction with TEMPO-H in the synthetic non-heme metal other than cobalt (copper, manganese, and chromium) and heme iron complexes has also been reported so far [15,16,17,18,19,20]. Therefore, the hydrogen atom transfer (HAT) from the N-hydroxy compounds to the metal-bound superoxide is a common reaction. In contrast, reactivity toward HAT from C-H substrates varied depending on the central metals and the supporting ligands [19,21,22]. As estimated by the DFT calculation, Mulliken spin densities of two oxygen atoms of O2 in 2S are almost equal. That suggests the weak radical reactivity of our superoxido species due to the delocalization of an unpaired electron (Figure S2).
In our system, the conversion from the superoxido complex to the hydroperoxido one also proceeded on the reaction with the reductant (i.e., [CoII(Cp)2]) followed by proton (i.e., trifluoroacetic acid). No reaction occurred in the mixing of the superoxido complex and reductant without any additive. In addition, the spectral pattern did not change in the mixing of the superoxido complex with trifluoroacetic acid. Those may indicate the generation of the hydroperoxido species is not through a stepwise electron transfer to proton transfer (ET/PT) mechanism but a concerted proton-coupled electron transfer (PCET) or a proton transfer to electron transfer (PT/ET).
As revealed by the kinetic analyses, the second order formation rate constants k2 for the cobalt(III)-hydroperoxido complexes 1H6H depend on the structural and electronic properties of the cobalt-supporting chelating ligands. Noteworthy, the electron-withdrawing bromine-involved TpMe2,Br complex 6H exhibited the largest k2 value. This result can be explained by increasing the electrophilicity of the superoxide due to the electron-withdrawing effect of the axial pyrazolyl ligand in TpMe2,Br, which is located at the trans position of the superoxide. Such electronic effect of the trans ligand has been proposed based on the order of the HAT activity of chromium(III)-superoxido complexes [19]. On the other hand, the electron-donating TpMe2,Me complex 5H showed a larger k2 value than that of the TpMe2,H complex 1H. The nucleophilicity of the superoxide affects the affinity of the proton. The degree of the electrophilic or nucleophilic nature of the superoxide ligand was tuned by the electronic property of the opposite ligand. The non-linear relationship between the electronic properties of TpMe2,X and the reactivity of the cobalt(III)-superoxido complexes can be attributed to the affinity of the electrons and protons, respectively, for the superoxide ligand. Although the O2 affinity (pressure-based equilibrium constant KO2) of the cobalt(II) complex with TpMe2,Me (i.e., 5R) is about twenty times larger than that of the TpMe2,Br complex 6R [8], the difference between the k2 of 5H and 6H was not so large. The electronic nature of TpMe2,X affects the cobalt center directly, while such effect on the superoxide ligand is weakened. Regarding the steric effect derived from the bidentate ligands LY, the order of k2 agreed with that of the O2 affinity (KO2) of the corresponding cobalt(II) complexes. However, these steric effects on the reaction of the superoxido complexes with AZADOL were not so large because the distal oxygen atom of the superoxide ligand was far from the cobalt center, as evidenced by the reported molecular structure of 2S [9]. The optimized structure of the hydroperoxido complex 2H shown in Figure 5 also suggests that the steric hindrance derived from Y of LY is relatively small.
Noteworthy, the DFT calculation of our superoxido and hydroperoxido complexes revealed that the interaction between the cobalt-O-O(-H) moiety and the bis(imidazolyl)borate ligands LY contributed to the SOMO of the superoxido species and the HOMO of the hydroperoxido species, respectively (Figures S4 and S5). Therefore, the electronic property of equatorial ligands might also affect the stability and reactivity of the superoxido and hydroperoxido species.

4. Materials and Methods

4.1. General

UV-vis spectra were measured on an Agilent 8453 UV-vis spectrometer with a UNISOK CoolSpeK cell holder. ESR measurement was performed on a JEOL AJES-FA100 ESR spectrometer with a liquid nitrogen Dewar at Tokyo Institute of Technology. Resonance Raman scattering was excited at 413 nm from a Kr+ laser (10 mW, Spectra Physics, BeamLok 2060, Oxfordshire, UK). The resonance Raman scattering was dispersed by a single polychromator (Ritsu Oyo Kogaku, MC-100DG, Saitama, Japan) and was detected by a liquid-nitrogen-cooled CCD detector (HORIBA JOBIN YVON, Symphony 1024 × 128 Cryogenic Front Illuminated CCD Detector, Edison, NJ, USA). A 135° back-scattering geometry was used. Theoretical calculations were performed using the DFT method implemented in the Gaussian 16 package of programs. The structures were fully optimized using the (U)B3LYP method. All calculations were performed using the scalar relativistic contracted versions of the def2-TZVP (Co) and def2-SVP (C, H, B, N, and O) basis sets. The structural optimization was performed by reading the coordinate data from the crystal structure, and the resulting molecular orbitals were visualized using Gauss View 6 programs. All calculations employed Tight SCF convergence criteria and were performed using the polarizable continuum model (PCM) to compute the structures in solutions. The excited states were calculated using the TD-DFT method within the Tamm–Dancoff approximation as implemented in Gaussian 16. These calculations employ the hybrid (U)B3LYP functional along with the basis sets described above. At least 100 excited states were computed in each calculation.
All commercial reagents and solvents were used without further purification unless otherwise noted. Preparations of oxygen-sensitive compounds were performed in a UNICO UN-800F glove box with an MF70 gas purification system or by the Schlenk technique under Ar atmosphere. Deuterated AZADOL was prepared by mixing D2O and AZADOL in THF for 10 min. After the addition of Na2SO4 into the mixture to remove excess D2O, the deuterated AZADOL solution was used directly for UV-vis kinetic measurement. The cobalt(II) complexes [CoII(TpMe2,X)(LY)] (1R6R) were prepared according to the literature.

4.2. Reaction of the Cobalt(III)-Superoxido Complex [CoIII(O2)(TpMe2,H)(LPh)] (1S) with AZADOL

In a glove box, 3.0 mg (4.83 mmol) of the cobalt(II) precursor [CoII(TpMe2,H)(LPh)] (1R) was dissolved in 15 mL. From the prepared 0.32 mM solution, 3 mL of this was placed into a quartz cell with the septum rubber sealing cap. The resulting cell was cooled at −60 °C, and then O2 gas was passed through this cold solution of 1R for 30 min. At this point, the UV-vis spectrum of this solution showed the saturation of the formed superoxido complex 1S. To this solution, 0.2 mL of THF solution of AZADOL (124 mM) was added.

4.3. Reaction of 1S with [CoII(Cp)2] and Trifluoroacetic Acid

A 0.32 mM THF solution (2.5 mL) of 1R was cooled at −60 °C, and oxygen was passed through until this solution was saturated with the superoxide complex 1S formed. To this solution, we added 0.2 mL of a THF solution of [CoII(Cp)2] (4.0 mM). Next, 0.8 mL of a 3.9 mM THF solution of trifluoroacetic acid was added.

4.4. Kinetic Analyses for the Formation of the Cobalt(III)-Superoxido Complexes 1H–6H

The reaction of the superoxido complexes 1S6S and AZADOL was traced by increasing the intensity of the absorbance at 300 nm in UV-vis measurement. THF solutions (0.31 mM, 2.5 mL) of the cobalt(II) precursors were cooled at −60 °C, and oxygen was passed through until these solutions were saturated with the superoxide complexes formed. To these solutions, we added 0.75 mL of THF solutions of AZADOL with various concentrations. The time course of the intensity of the 300 nm was analyzed as a pseudo-first-order reaction to obtain rate constants kobs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196416/s1, Figure S1: Simulated absorption spectra; Figure S2: Mulliken spin density and spin density distribution of the superoxido complex; Figure S3: Charge distribution of the superoxido and hydroperoxido complexes obtained by the natural population analysis; Figure S3: selected molecular orbitals of the superoxido complex; Figure S4: selected molecular orbitals of the hydroperoxido complex. Figure S5: Selected molecular orbitals of the hydroperoxido complex. References [23,24,25,26,27,28,29] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, T.N., J.N. and S.H.; investigation, T.N., T.O. (Takehiro Ohta); J.N. and M.O.; resources, T.O. (Takashi Ogura); data curation, T.N., J.N. and M.O.; writing—original draft preparation, T.N.; writing—review and editing, M.O. and S.H.; supervision, S.H. All authors except T.O. (Takashi Ogura) have read and agreed to the published version of the manuscript.

Funding

This research was funded by CREST, JST (JPMJCR16P1) and Kanagawa University (ordinary budget for 411).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are contained within this article and are supported by the data in the Supplementary Materials.

Acknowledgments

We acknowledge Toshiro Takao in Tokyo Institute of Technology for EPR measurement.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples are not available from the authors.

References

  1. Poulos, T.L. Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 3919–3962. [Google Scholar] [CrossRef]
  2. Bollinger, J.M., Jr.; Krebs, C. Enzymatic C−H activation by metal−superoxo intermediates. Curr. Opin. Chem. Biol. 2007, 11, 151–158. [Google Scholar] [CrossRef]
  3. Klinman, J.P. The Copper-Enzyme Family of Dopamine-Monooxygenase and Peptidylglycine—Hydroxylating Monooxygenase: Resolving the Chemical Pathway for Substrate Hydroxylation. J. Biol. Chem. 2006, 281, 3013–3016. [Google Scholar] [CrossRef]
  4. Whittaker, J.W. Free Radical Catalysis by Galactose Oxidase. Chem. Rev. 2003, 103, 2347–2363. [Google Scholar] [CrossRef]
  5. Hedegard, E.D.; Ryde, U. Molecular mechanism of lytic polysaccharide monooxygenases. Chem. Sci. 2018, 9, 3866–3880. [Google Scholar] [CrossRef]
  6. Fukuzumi, S.; Lee, Y.-M.; Nam, W. Structure and reactivity of the first-row d-block metal-superoxo complexes. Dalton. Trans. 2019, 48, 9469–9489. [Google Scholar] [CrossRef]
  7. Noh, H.; Cho, J. Synthesis, characterization and reactivity of non-heme 1st row transition metal-superoxo intermediates. Coord. Chem. Rev. 2019, 382, 126–144. [Google Scholar] [CrossRef]
  8. Nishiura, T.; Chiba, Y.; Nakazawa, J.; Hikichi, S. Tuning the O2 Binding Affinity of Cobalt(II) Centers by Changing the Structural and Electronic Properties of the Distal Substituents on Azole-Based Chelating Ligands. Inorg. Chem. 2018, 57, 14218–14229. [Google Scholar] [CrossRef]
  9. Oddon, F.; Chiba, Y.; Nakazawa, J.; Ohta, T.; Ogura, T.; Hikichi, S. Characterization of Mononuclear Non-heme Iron(III)-Superoxo Complex with a Five-Azole Ligand Set. Angew. Chem. Int. 2015, 54, 7336–7339. [Google Scholar] [CrossRef]
  10. Rajani, C.; Kincaid, J.R.; Petering, D.H. Resonance Raman Studies of HOO-Co(III)Bleomycin and Co(III)Bleomycin: Identification of Two Important Vibrational Modes, (Co-OOH) and (O-OH). J. Am. Chem. Soc. 2004, 126, 3829–3836. [Google Scholar] [CrossRef]
  11. Wang, C.-C.; Chang, H.-C.; Lai, Y.-C.; Fang, H.; Li, C.-C.; Hsu, H.-K.; Li, Z.-Y.; Lin, T.-S.; Kuo, T.-S.; Neese, F.; et al. A Structurally Characterized Nonheme Cobalt−Hydroperoxo Complex Derived from Its Superoxo Intermediate via Hydrogen Atom Abstraction. J. Am. Chem. Soc. 2016, 138, 14186–14189. [Google Scholar] [CrossRef]
  12. Shin, B.; Sutherlin, K.D.; Ohta, T.; Ogura, T.; Solomon, E.I.; Cho, J. Reactivity of a Cobalt(III)−Hydroperoxo Complex in Electrophilic Reactions. Inorg. Chem. 2016, 55, 12391–12399. [Google Scholar] [CrossRef]
  13. Anandababu, K.; Muthuramalingam, S.; Velusamy, M.; Mayilmurugan, R. Single-step benzene hydroxylation by cobalt(II) catalysts via a cobalt(III)-hydroperoxo intermediate. Catal. Sci. Technol. 2020, 10, 2540–2548. [Google Scholar] [CrossRef]
  14. Gordon, J.B.; Vilbert, A.C.; Siegler, M.A.; Lancaster, K.M.; Moënne-Loccoz, P.; Goldberg, D.P. A Nonheme Thiolate-Ligated Cobalt Superoxo Complex: Synthesis and Spectroscopic Characterization, Computational Studies, and Hydrogen Atom Abstraction Reactivity. J. Am. Chem. Soc. 2019, 141, 3641–3653. [Google Scholar] [CrossRef]
  15. Tano, T.; Okubo, Y.; Kunishita, A.; Kubo, M.; Sugimoto, H.; Fujieda, N.; Ogura, T.; Itoh, S. Redox Properties of a Mononuclear Copper(II)-Superoxide Complex. Inorg. Chem. 2013, 52, 10431–10437. [Google Scholar] [CrossRef]
  16. Kindermann, N.; Günes, C.-J.; Dechert, S.; Meyer, F. Hydrogen Atom Abstraction Thermodynamics of a μ-1,2-Superoxo Dicopper(II) Complex. J. Am. Chem. Soc. 2017, 139, 9831–9834. [Google Scholar] [CrossRef]
  17. Bailey, W.D.; Dhar, D.; Cramblitt, A.C.; Tolman, W.B. Mechanistic Dichotomy in Proton-Coupled Electron-Transfer Reactions of Phenols with a Copper Superoxide Complex. J. Am. Chem. Soc. 2019, 141, 5470–5480. [Google Scholar] [CrossRef]
  18. Lin, Y.-H.; Cramer, H.H.; van Gastel, M.; Tsai, Y.-H.; Chu, C.-Y.; Kuo, T.-S.; Lee, I.-R.; Ye, S.; Bill, E.; Lee, W.-Z. Mononuclear Manganese(III) Superoxo Complexes: Synthesis, Characterization, and Reactivity. Inorg. Chem. 2019, 58, 9756–9765. [Google Scholar] [CrossRef]
  19. Goo, Y.R.; Maity, A.C.; Cho, K.-B.; Lee, Y.-M.; Seo, M.S.; Park, Y.J.; Cho, J.; Nam, W. Tuning the Reactivity of Chromium(III)-Superoxo Species by Coordinating Axial Ligands. Inorg. Chem. 2015, 54, 10513–10520. [Google Scholar] [CrossRef]
  20. Kim, H.; Rogler, P.J.; Sharma, S.K.; Schaefer, A.W.; Solomon, E.I.; Karlin, K.D. Heme-FeIII Superoxide, Peroxide and Hydroperoxide Thermodynamic Relationships: FeIII-O2 Complex H-Atom Abstraction Reactivity. J. Am. Chem. Soc. 2020, 142, 3104–3116. [Google Scholar] [CrossRef]
  21. Chiang, C.-W.; Kleespies, S.T.; Stout, H.D.; Meier, K.K.; Li, P.-Y.; Bominaar, E.L.; Que, L., Jr.; Münck, E.; Lee, W.-Z. Characterization of a Paramagnetic Mononuclear Nonheme Iron-Superoxo Complex. J. Am. Chem. Soc. 2014, 136, 10846–10849. [Google Scholar] [CrossRef] [PubMed]
  22. Blakely, M.N.; Dedushko, M.A.; Poon, P.C.Y.; Villar-Acevedo, G.; Kovacs, J.A. Formation of a Reactive, Alkyl Thiolate-Ligated FeIII-Superoxo Intermediate Derived from Dioxygen. J. Am. Chem. Soc. 2019, 141, 1867–1870. [Google Scholar] [CrossRef] [PubMed]
  23. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  24. Weigend, F.; Ahlrichs., A. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef] [PubMed]
  25. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6.1; Semichem Inc.: Shawnee Mission, KS, USA, 2016. [Google Scholar]
  26. Cossi, M.; Scalmani, G.; Rega, N.; Barone, V.J. New developments in the polarizable continuum model for quantum mechanical and classical calculations on molecules in solution. Chem. Phys. 2002, 117, 43. [Google Scholar] [CrossRef]
  27. Casida, M.E.; Jamorski, C.; Casida, K.C.; Salahub, D.R.J. Molecular excitation energies to high-lying bound states from time-dependent density-functional response theory: Characterization and correction of the time-dependent local density approximation ionization threshold. Chem. Phys. 1998, 108, 4439. [Google Scholar] [CrossRef]
  28. Statmann, R.E.; Scuseria, G.E.J. An efficient implementation of time-dependent density-functional theory for the calculation of excitation energies of large molecules. Chem. Phys. 1998, 109, 8218. [Google Scholar]
  29. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
Scheme 1. Formation of cobalt(III)-hydroperoxo species investigated in this study.
Scheme 1. Formation of cobalt(III)-hydroperoxo species investigated in this study.
Molecules 27 06416 sch001
Figure 1. UV-vis spectrum of 1S and the resultant 1H formed by the reaction of 1S with AZADOL.
Figure 1. UV-vis spectrum of 1S and the resultant 1H formed by the reaction of 1S with AZADOL.
Molecules 27 06416 g001
Figure 2. Resonance Raman spectra of 1H derived from 16O2 (red) and 18O2 (blue). The Co-O stretching bands attributed to the remaining superoxido complex 1S were observed.
Figure 2. Resonance Raman spectra of 1H derived from 16O2 (red) and 18O2 (blue). The Co-O stretching bands attributed to the remaining superoxido complex 1S were observed.
Molecules 27 06416 g002
Figure 3. EPR spectra of 1S (a) and (b) the reaction mixture of 1S and AZADOL in THF at −196 °C.
Figure 3. EPR spectra of 1S (a) and (b) the reaction mixture of 1S and AZADOL in THF at −196 °C.
Molecules 27 06416 g003
Figure 4. Pseudo first-order analyses of the reaction with AZADOL and AZADOL-d1.
Figure 4. Pseudo first-order analyses of the reaction with AZADOL and AZADOL-d1.
Molecules 27 06416 g004
Figure 5. Optimized molecular structures of the superoxido (left) and hydroperoxido (right) complexes with TpMe2,H and LOiPr.
Figure 5. Optimized molecular structures of the superoxido (left) and hydroperoxido (right) complexes with TpMe2,H and LOiPr.
Molecules 27 06416 g005
Figure 6. Plots of the concentration of AZADOL versus kobs for 1H4H.
Figure 6. Plots of the concentration of AZADOL versus kobs for 1H4H.
Molecules 27 06416 g006
Figure 7. The plot of the concentration of AZADOL versus kobs of 1H, 5H, and 6H.
Figure 7. The plot of the concentration of AZADOL versus kobs of 1H, 5H, and 6H.
Molecules 27 06416 g007
Table 1. ν(Co-O) and ν(O-O) data of 1S, 1H, and reported cobalt(III)-hydroperoxido species.
Table 1. ν(Co-O) and ν(O-O) data of 1S, 1H, and reported cobalt(III)-hydroperoxido species.
Complex(Co-O)/cm−1ν(O-O)/cm−1Reference
16O218O216O218O2
1S54452311501090[8]
1H562542823780This work
(bleomycin)CoIII(OOH)545518828784[10]
[CoIII(BDPP)(OOH)] 1--795748[11]
[CoIII(OOH)(Me3-TPADP)(MeCN)]2+ 2571551851803[12]
[CoIII(OOH)(L2)] 3564544832781[13]
1 H2BDDP = 2,6-bis((2-(S)-diphenylhydroxylmethyl-1-pyrrolidinyl)methyl)pyridine. 2 Me3-TPADP = 3,6,9-trimethyl-3,6,9-triaza-1(2,6)-pyridinacyclodecaphane. 3 L2 = N,N-bis(2-pyridylmethyl)-N′,N′-dimethylpropane-1,3-diamine.
Table 2. Selected bond lengths (Å) of the optimized structures of the superoxido and hydroperoxido species with the experimental data of 2S.
Table 2. Selected bond lengths (Å) of the optimized structures of the superoxido and hydroperoxido species with the experimental data of 2S.
SuperoxidoHydroperoxido
Experimental/Å 1Optimized/ÅOptimized/Å
O–O1.301 (5)1.2761.425
Co–O1.901 (3)1.9231.894
Co–NPz,ax2.052 (4)2.0562.050
Co–NPz,eq1.978 (3)2.0062.019
Co–NPz,eq2.005 (4)2.0032.002
Co–NIm,eq1.962 (4)1.9591.958
Co–NIm,eq1.952 (3)1.9571.954
1 From ref. [9].
Table 3. The second-order formation rate constants (k2) for the hydroperoxido species 1H6H and related parameters of the cobalt(II) precursors 1R6R.
Table 3. The second-order formation rate constants (k2) for the hydroperoxido species 1H6H and related parameters of the cobalt(II) precursors 1R6R.
ComplexX of TpMe2,XY of LYk2 for H
/M−1 s−1
KO2 of R
/atm−1 1
Co(II)/(III) Oxidation Potential of R/V 1
1HPh4.4 × 10218.880.18
2HOiPr4.0 × 1027.400.21
3HMe3.0 × 1022.620.07
4HBu2.5 × 1020.770.12
5MePh5.3 × 10261.410.11
6BrPh8.2 × 1022.990.36
1 From ref. [8].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nishiura, T.; Ohta, T.; Ogura, T.; Nakazawa, J.; Okamura, M.; Hikichi, S. The Conversion of Superoxide to Hydroperoxide on Cobalt(III) Depends on the Structural and Electronic Properties of Azole-Based Chelating Ligands. Molecules 2022, 27, 6416. https://doi.org/10.3390/molecules27196416

AMA Style

Nishiura T, Ohta T, Ogura T, Nakazawa J, Okamura M, Hikichi S. The Conversion of Superoxide to Hydroperoxide on Cobalt(III) Depends on the Structural and Electronic Properties of Azole-Based Chelating Ligands. Molecules. 2022; 27(19):6416. https://doi.org/10.3390/molecules27196416

Chicago/Turabian Style

Nishiura, Toshiki, Takehiro Ohta, Takashi Ogura, Jun Nakazawa, Masaya Okamura, and Shiro Hikichi. 2022. "The Conversion of Superoxide to Hydroperoxide on Cobalt(III) Depends on the Structural and Electronic Properties of Azole-Based Chelating Ligands" Molecules 27, no. 19: 6416. https://doi.org/10.3390/molecules27196416

Article Metrics

Back to TopTop