Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Jan 31.
Published in final edited form as: Inorg Chem. 2023 Sep 20;62(39):15842–15855. doi: 10.1021/acs.inorgchem.3c01234

Hydrogen Peroxide Disproportionation Activity Is Sensitive to Pyridine Substitutions on Manganese Catalysts Derived from 12-Membered Tetra-Aza Macrocyclic Ligands

David M Freire 1, Hannah M Johnston 1, Katherine J Smith 1, Kristof Pota 1, Magy A Mekhail 1, Sugam Kharel 1, Kayla N Green 1
PMCID: PMC10829483  NIHMSID: NIHMS1958338  PMID: 37729496

Abstract

The abundance of manganese in nature and versatility to access different oxidation states have made manganese complexes attractive as catalysts for oxidation reactions in both biology and industry. Macrocyclic ligands offer the advantage of substantially controlling the reactivity of the manganese center through electronic tuning and steric constraint. Inspired by the manganese catalase enzyme, a biological catalyst for the disproportionation of H2O2 into water and O2, the work herein employs 12-membered tetra-aza macrocyclic ligands to study how the inclusion of and substitution to the pyridine ring on the macrocyclic ligand scaffold impacts the reactivity of the manganese complex as a H2O2 disproportionation catalyst. Synthesis and isolation of the manganese complexes was validated by characterization using UV–vis spectroscopy, SC-XRD, and cyclic voltammetry. Potentiometric titrations were used to study the ligand basicity as well as the thermodynamic equilibrium with Mn(II). Manganese complexes were also produced in situ and characterized using electrochemistry for comparison to the isolated species. Results from these studies and H2O2 reactivity showed a remarkable difference among the ligands studied, revealing instead a distinction in the reactivity regarding the number of pyridine rings within the scaffold. Moreover, electron-donating groups on the 4-position of the pyridine ring enhanced the reactivity of the manganese center for H2O2 disproportionation, demonstrating a handle for control of oxidation reactions using the pyridinophane macrocycle.

Graphical Abstract

graphic file with name nihms-1958338-f0001.jpg

INTRODUCTION

Oxidation catalysis with mild oxidants, such as H2O2, has been desired due to the possibility of performing important oxidation steps in industry at a lower cost and less impact to the environment. The oxidation reactivity of H2O2 can be hindered by a mutually exclusive reaction known as disproportionation, where H2O2 is both oxidized and reduced to produce molecular oxygen and water. Such reactivity is common in nature, which has developed the superior catalytic tools to appropriately mitigate the reactivity of H2O2 and yield unrivaled turnover frequencies and reaction rates. For comparison, peroxidases use H2O2 to catalyze oxidation of organic matter, while catalases disproportionate H2O2. Consistent with other metalloenzymes, both peroxidase and catalase contain a metal center in their structure, and their reactivity is modulated by a dynamic environment composed of amino acid residues capable of fine-tuned regulation of the coordination sphere using electronics. As a result, the enzymes can achieve a range of chemical potentials that allow the performance of their target reactions with high catalytic efficiency and selectivity, making the isolation and study of the active site chemistry an attractive challenge.1

The manganese-based catalase (MnCAT) is one of the most active enzymes in nature, performing disproportionation of H2O2 with a rate of 1.8–8.4 × 106 M−1 min−1 and turnover frequencies in the range of 106 s−1. 2,3 These remarkable values exemplify the importance of MnCAT for efficient disproportionation of H2O2 in nature, which has been a challenge to transfer to the bench. This is, in part, due to the unique redox activity of the binuclear active site (Figure 1) that can cycle between Mn(II)/Mn(II) and Mn(III)/Mn(III) redox states during catalytic turnovers. Moreover, the oxidized and reduced forms of the enzymes differ from one another by the bridging ligands connecting the manganese centers. The reduced form of the enzyme features an active site containing two Mn(II) ions connected by the carboxylate group of a glutamate residue in a synsyn fashion.4 The motif is completed by two bridging μ-OH and μ-OH2 moieties. In contrast, the oxidized form features two μ-O bridged ligands along with the carboxylate bridging unit.4

Figure 1.

Figure 1.

Schematic representation of the active site of MnCAT from Lactobacillus plantarum, highlighting the (left) reduced form and (right) oxidized form, which differed by the bridging ligands based on ref 4.

Considerable efforts have been made to develop synthetic mimics of MnCAT for applications in oxidation catalysis and as pharmaceutical agents to regulate excess H2O2 and other reactive oxygen species (ROS).5 Nevertheless, mimetics are far from reaching analogous activities compared to native enzymes, largely because of the inability to perform the same reactivity using low Mn(II)/Mn(III) oxidation states. Instead, mimetics must reach high oxidation states such as Mn(IV) and Mn(V), which hinder the kinetic efficiency due to the energy required to achieve these highly reactive oxidation states.6 New approaches and ligand scaffolds are continuously being developed to improve on the oxidation-type reactions through the high-valent pathways. Catalase-like reactivity is a relatively simple reaction that can be used as a probe to learn the behavior of manganese-based small molecules in oxidation catalysis using H2O2 as the main oxidant to classify manganese complexes of diverse structure based on their performance.

The ligand scaffold is important to consider in attempts to stabilize high oxidation states of manganese. Macrocycles stand out due to their high thermodynamic stability of complex formation with catalytically active transition metals, compared to open chain ligands. Moreover, the diversity within the library of macrocyclic ligand frameworks offers the opportunity to adjust electronic effects, cavity size, and basicity to subsequently tune the properties of the metal complex to yield a preferred structure for a target reactivity. For example, inclusion of pyridine rings in pyclen-based pyridinophanes (PyN3) has shown important results in stabilizing high oxidation states of iron in the literature, as opposed to cyclen (N4) and cyclam scaffolds. Based on these results, pyridinophanes have been used in oxidation catalysis with manganese.7 Previous studies have shown that manganese complexes of PyN3 were more active than N4 for the disproportionation of H2O2. However, the catalytic efficiencies or any mechanistic insights of the manganese complexes of PyN3 and N4 toward H2O2 disproportionation have not been described. Moreover, spectroscopic evidence from the preliminary studies suggested the presence of bis-μ-oxo dimers for the ligands N4 and PyN3, but identification of such a species with the addition of the second pyridine ring in the Py2N2 scaffold of the catalyst remains unclear.8

Inspired by the ability of 12-membered pyridinophane macrocycles to modulate the reactivity of manganese and knowledge gathered from DFT calculations on MnCAT mechanisms,911 we hypothesized that manganese complexes of 12-membered tetra-aza macrocyclic ligands would be more efficient for catalytic disproportionation of H2O2 if they possess (1) electron-donating groups (EDGs) that facilitate higher oxidation states on the metal center and (2) pyridine moieties that can assist delocalization of the charge to modulate electron density through different oxidation states. With this goal, we prepared a group of ligands that ranged from cyclen (N4), pyclen (PyN3) and its pyridine substituted congeners (OHPyN3 and OMePyN3) as well as Py2N2 (Figure 2), and their manganese complexes to systematically understand what properties were modulated by the differences between each ligand set and how these results related to metrics of H2O2 disproportionation.

Figure 2.

Figure 2.

Structures of the ligands organized by the variables considered in this study.

RESULTS AND DISCUSSION

Pyridine Modifications on the Ligand Scaffold Tune Their Basicity.

The ligand series studied (Figure 2) was selected to systematically modify 12-membered tetra-aza pyridinophanes by changing the number of pyridine rings on the macrocycle or attaching EDGs to the pyridine. Apart from cyclen (N4), which was obtained commercially, the ligands were prepared following previously published methods.1215 The PyN3 series were isolated as HCl salts, and the Py2N2 ligand as a free base. All ligands are highly water-soluble, which aided greatly in the studies described herein.

Inclusion of pyridine rings within the macrocycle is known to increase the rigidity of the macrocyclic ligand and subsequently a decrease in basicity, which can be quantified by the pH of the protonation events of each ligand.16,17 Therefore, stepwise protonation constants of the ligand series were obtained by pH-potentiometric titrations (25 °C and 0.15 M NaCl). The pKa (logKiH) values for each measurable protonation event within the N4, PyN3, and Py2N2 series (Table 1) were consistent with the decrease of basicity from secondary amines to pyridine rings.18 Based on the sum of the pKa values determined for the N atom donors of each ligand (Σ pKa,N-donors), basicity decreased from N4 (21.35) > PyN3 (21.20) ≫ Py2N2 (15.63). This trend shows that substitution of two secondary amines for a pyridine nitrogen has a much bigger impact on the overall basicity of the ligand than just a single substitution, which only decreased the basicity slightly. On the other hand, EDGs on the 4-position of the pyridine ring decreased the total basicity from OHPyN3 (21.97) > PyN3 (21.20) ≫ OMePyN3 (20.07). Substitutions of EDG on the 4-position of the pyridine ring of the PyN3 ligand fine-tune the overall basicity of the ligand over a range of roughly two log units (Σ pKa,N-donors = 20.07–21.97), while substitution of secondary amines with pyridine units results in much more broad range changes (Σ pKa,N-donors = 15.63–21.35).

Table 1.

Stepwise Protonation Constants (pKa) of 12-Membered Pyridinophane Macrocycles (T = 298 K, I = 0.15 M NaCl)

N4a Py2N2 PyN3b OHPyN3b OMePyN3c
pKa 1 10.81(1) 8.27(4) 11.37(1) 11.38(1) 10.32(2)
pKa 2 9.66(2) 7.36(4) 8.22(5) 8.85(3) 8.00(3)
pKa 3 0.88(2) 1.61(5) 5.59(4) 1.75(4)
pKa 4 1.74(4)
Σ pKa 21.35 15.63 21.20 27.56 20.07
Σ pKa,N-donors 21.35 15.63 21.20 21.97 20.07
a

From ref 8.

b

From ref 18.

c

From ref 19.

Ligand Basicity Modulate Manganese Complexation and Hydrolysis.

Speciation studies of each ligand with Mn(II) were essential, given that complex [(Mn(L)] formation often occurs at pH values that overlap with formation of intractable Mn(OH)x species, which complicates studies of these complexes in solution.20 Therefore, thermodynamic stability of each ligand with Mn(II) ions was obtained by pH-potentiometric titrations (25 °C and 0.15 M NaCl) to determine the optimal pH for Mn(II) complex formation and to identify the species present in solution at different pH values. This approach significantly aids in the goal of preparing each manganese complex in situ. pH-potentiometric titrations were performed with Mn(II) ions under a N2 atmosphere since Mn(III) species are known to both easily hydroxylate and disproportionate into Mn(II) and Mn(IV) species, concomitant with formation of MnOx species.21

Thermodynamic stability (log KML) of the Mn(II) complexes of the ligand series (Table 2) shows that the PyN3 scaffold forms the most stable complex regardless of the pyridine substitution on the 4-position, following the basicity trend observed for the protonation steps of the ligand alone: OHPyN3 (10.96) > PyN3 (10.11) > OMePyN3 (9.70). The effect of the number of pyridine rings within the ligand scaffold on the log KML did not follow the trend observed with ligand basicity. Rather, it decreased from PyN3 (10.11) > N4 (8.34) ≈ Py2N2 (8.26). Since the ligands studied here have the same number of atoms in the cavity size, the difference in stability observed with changing the number of pyridine rings suggests that this observation is a consequence of variability in steric energy required for conformational reorganization of the macrocyclic cavity to complex Mn(II).22,23

Table 2.

Stability Constants, Protonation Constants, and pMn Values for the Mn(II) Complexes (I = 0.15 M NaCl, T = 25°C)

ligand logKML logKMHL logKM(OH)L pMn
N4 8.34(2)a 10.22(6)a 5.01
Py2N2 8.26(9) 9.7(2) 5.96
PyN3 10.11(4)a 11.0(5)a 5.25
OHPyN3 10.96(4) 6.40(7) 10.18(6) 5.39
OMePyN3 9.70(9) 9.5(2) 5.63
a

From ref 8.

To determine the effect of the ligand basicity on the speciation of Mn(II) ions in solution at different pH values, the pMn (a conditional stability constant) was calculated for each Mn(II) complex. pMn was calculated by pMn = −log[Mn(II)]free from speciation curves at a concentration of [L] = [M] = 0.01 mM at pH = 7.4.16,17,24 Higher pMn values reflect the ability of a given ligand to complex with Mn(II) ions at more acidic pH values. The calculated pMn values for complexes with different number of pyridine rings (Table 2) correlated with the basicity of the ligands: the more basic the ligand, the higher the competition of the metal versus protons for the donor sites, pMn for Py2N2 (5.96) > PyN3 (5.25) > N4 (5.01). Substitution of the 4-position of the pyridine ring in the PyN3 systems resulted in a slight increase of the pMn values from PyN3 (5.25) > OHPyN3 (5.39) > OMePyN3 (5.63). This is consistent with the trend observed for log KML, as well. The lowest basicity within the PyN3 ligands was obtained for OMePyN3, which indicates that there is less competition of Mn(II) ions for donor sites, resulting in the highest pMn among the PyN3 series. Overall, the effect of the ligand basicity on Mn(II) complexation can be visualized in Figure 3, where speciation of free Mn(II) begins to decrease at around pH = 5.5 for Py2N2. In contrast, this transition occurs around pH = 6.5 for OMePyN3 and OHPyN3. PyN3 and N4 exhibit complexation of Mn(II) at higher pH values (PyN3 pH = 7, N4 pH = 8), revealing the lowest pMn in this study.

Figure 3.

Figure 3.

Speciation of free Mn(II) ions with 12-membered tetra-aza macrocycles (I = 0.15 M NaCl, T = 25 °C, [L] = [M] = 0.01 mM).

Manganese Complexes Were Isolated as Mononuclear and Dinuclear Species with Distinctive Spectroscopic and Crystallographic Features.

Complexation of manganese was pursued with methods driven by understanding the optimal aqueous formation conditions. The pH for complex formation was a key choice to consider when working with manganese. Speciation diagrams derived from potentiometric titrations can help to circumvent trial and error. As shown in Figures S1S5, maximum [M(L)] complex formation for N4, PyN3, and OHPyN3 is achieved at pH = 8, while the population of the corresponding hydroxyl complex [M(OH)L] at pH = 8 was higher for OMePyN3 and Py2N2, compared to N4, PyN3, and OHPyN3. As predicted from this data, monomeric species were obtained by mixing an aqueous solution of OMePyN3 or Py2N2 at pH = 7 with equimolar amounts of the precursor Mn(II) salt without further adjustment of pH, while the pH was set to 8 for OHPyN3. The solution showed slow darkening due to oxidation of Mn(II) to Mn(III) under air atmosphere. Water was evaporated under reduced pressure and products were isolated in MeCN as pink powders. The color of these compounds is characteristic of a mononuclear [MnL]3+ manganese complex with OHPyN3, OMePyN3, or Py2N2 using previously reported procedures with some variations based on the results of potentiometric titrations.8 Dark forest green powder characteristic of dinuclear [Mn2(μ-O)2L2]3+ complexes was obtained with OHPyN3 using a reported procedure without modifications, but in the case of OMePyN3 and Py2N2, only monomers were obtained.8

Monomeric complexes [Mn(OHPyN3)Cl2][ClO4], [Mn(OMePyN3)Cl2][Cl], and [Mn(Py2N2)Cl2][Cl] were dissolved in MeCN, and each solution exhibited a pink color characteristic of cis-Mn(III) mononuclear complexes in the Literature2527. Ligand to metal charge transfer (LMCT) bands were observed from 350 to 400 nm along with a d–d transition band centered at 525 nm (Figure 4, top). The dimer complex [Mn2(μ-O)2(OHPyN3)2][ClO4]3 was dissolved in water, exhibiting its characteristic dark forest green color. Spectroscopic features matched the bands reported for other bis-μ-oxo complexes of manganese, LMCT absorptions in the region from 350 to 450 nm, the characteristic sharp band at 550 nm corresponding to a Mn(III) d–d transition, broad absorbance centered at 650 nm (LMCT oxo to Mn(IV)), and an intervalence band at low energy (~750 nm) (Figure 4, bottom).28,29

Figure 4.

Figure 4.

Electronic absorption spectrum at 25 °C of (top) monomeric Mn(III) complexes of OHPyN3, OMePyN3, and Py2N2 in MeCN, and (bottom) dimeric Mn2(III,IV) complex of OHPyN3 in water.

Single crystal X-ray diffraction (SC-XRD) analysis was carried out on the crystalline materials obtained from metalation of the ligand series. Mononuclear Mn(III) species were modeled from manganese complexation with ligands OMePyN3 and Py2N2, [Mn(OMePyN3)F2][BF4], and [Mn(Py2N2)Cl2][BF4] (Figure 5). Throughout the series, Mn(III) is bound in a cis conformation, with two nitrogen atoms in the axial plane (N1 and N3) and two nitrogen atoms in the equatorial plane (N2 and N4). The distorted octahedral coordination is completed by two Cl ions with Py2N2 and two F ions for OMePyN3. The presence of fluoride ligands was not originally anticipated; however, it is consistent with reports of metal complexes with highly nucleophilic metal centers or by the presence of HF that have produced F after slow hydrolysis of salts of BF4 over time.3032

Figure 5.

Figure 5.

ORTEP representations of solid-state structures of (A) [Mn2(μ-O)2(OHPyN3)2][ClO4]3, (B) [Mn(OMePyN3)F2][BF4], and (C) [Mn(Py2N2)Cl2][BF4]. Counterions and solvents have been removed here for clarity. Atoms are drawn at 50% thermal probability ellipsoids.

Mononuclear Complexes.

Bond lengths and angles were compared with the previously reported [Mn(PyN3)Cl2][ClO4] complex (Table 3).8 Bond lengths Mn–N1 and Mn–N3 depict elongation of ~0.2 Å compared to the N2 and N4 bonds, typically observed for d4 Mn(III) complexes due to Jahn–Teller distortion. The N1–Mn–N3 axial angle was similar for [Mn(PyN3)Cl2][ClO4] and [Mn(OMePyN3)F2][BF4] (149.22° and 149.08°, respectively), but [Mn(Py2N2)Cl2][BF4] shows a smaller angle (145.53°), consistent with the increase of the rigidity by the addition of the second pyridine ring. The equatorial angle N2–Mn–N4 shows modest changes among the complexes with similar trends to the axial angle.

Table 3.

Selected Bond Lengths (Å) and Angles (°) from SC-XRD Analysis of Monomeric Mn(lIl) Complexes

bond length (Å)/angle (°) PyN3a OMePyN3b Py2N2c Py2MeN2d
Mn–N1 2.247(2) 2.298(1) 2.255(2) 2.328(2)
Mn–N2 2.076(2) 2.078(1) 2.041(2) 2.213(2)
Mn–N3 2.241(3) 2.272(1) 2.255(2) 2.327(2)
Mn–N4 2.059(2) 2.032(1) 2.041(2) 2.198(2)
N1–Mn–N3 149.22(9) 149.08(5) 145.53(12) 143.44(6)
N2–Mn–N4 87.33(10) 86.62(5) 78.01(8) 77.61(6)
a

[Mn(PyN3)Cl2][ClO4], Freire et al.8

b

[Mn(OMePyN3)F2][BF4], this work.

c

[Mn(Py2N2)Cl2][BF4], this work.

d

[Mn(Py2MeN2)(MeCN)2][PF6]2, Lee et al.36

The bond distance between Mn(III) and the N atom of the pyridine ring can be used to assess the strength of their bonding. Functionalization of the pyridine ring has been shown to modulate the bond length between the pyridine and metal centers such as Fe(III), Ni(II), and Cu(II).3335 [Mn(PyN3)Cl2][ClO4] has a longer Mn–N4 bond (2.059 Å) than [Mn(OMePyN3)F2][BF4] (2.032 Å), which is consistent with the trend observed in the literature for other metal centers with EDGs substituted on the pyridine ring.3335 Interestingly, [Mn(Py2N2)Cl2][BF4] has a bond distance Mn–N4 intermediate (2.041 Å) between [Mn(PyN3)Cl2][ClO4] and [Mn(OMePyN3)F2][BF4], which could be attributed to the change of the ligand scaffold.

Bond distances for Mn–N1 and Mn–N3 can be used as a readout for the effect of the basicity of the N-donors of the amines as well. For example, [Mn(Py2N2)Cl2][BF4] provides shorter bond distances (Mn–N1 = Mn–N3 = 2.041 Å) than the previously reported complex [Mn(Py2MeN2)(MeCN)2][PF6]2 (Mn–N1 = Mn–N3 = 2.327 Å),36 consistent with the higher basicity of secondary amines (Py2N2) in comparison with tertiary amines (Py2MeN2); nevertheless, the difference in oxidation state of manganese, being Mn(II) for Py2MeN2 and Mn(III) for Py2N2, does not allow a clear comparison of ligand basicity due to the differences in acidity of Mn(II) vs Mn(III). Substitutions of the PyN3 scaffold provide a much better comparison to study the impact of changes to the macrocyclic ligand on the solid-state structure. The structure modeled for [Mn(PyN3)Cl2][ClO4] shows shorter bond distances for Mn–N1 and Mn–N3 (Mn–N1 = 2.247 Å, Mn–N3 = 2.241 Å) compared to [Mn(OMePyN3)F2][BF4] (Mn–N1 = 2.298 Å, Mn–N3 = 2.272 Å), which is in agreement with the higher basicity of the PyN3 scaffold obtained with pH-potentiometric titrations.

Dinuclear Complexes.

X-ray quality crystals of a dinuclear Mn2(III,IV) complex with formulation of [Mn2(μ-O)2(OHPyN3)2][ClO4]3 was observed for materials obtained from metalation of OHPyN3 with Mn(ClO4)2 salts. All attempts to obtain monomeric crystalline congeners were unsuccessful and resulted only in isolation of dimeric species. The geometry around each manganese ion is distorted octahedral, with two cis μ-oxo bridges and four N atoms from the ligand. An inversion center on the O atom results in crystallographically equivalent manganese atoms within this system, consistent with the Mn–N bond lengths that are intermediate between a d3 Mn(IV) and d4 Mn(III). This effect has been observed in several other Mn2(III,IV) dimers of this type, including the previously reported structure [Mn2(μ-O)2(N4)2][Cl]3, attributing the effect to static disorder in the crystals.37

In contrast, the previously reported complex [Mn2(μ-O)2(PyN3)2][ClO4]3 shows evidence for two crystallographically distinct metal centers through inspection of the Mn–N bond lengths between the two manganese centers (Table 4). Mn1 is assigned as the d3 Mn(IV) ion with axial and equatorial bond distances relatively similar to one another (Mn1–N1 = 2.086 Å, Mn1–N2 = 2.091 Å, and Mn1–N3 = 2.090 Å). Mn2 is consistent with a d4 Mn(III) ion with Jahn–Teller distortion observed in the lengthening of the Mn–N axial bonds (Mn2–N1 = 2.246 Å and Mn2–N3 = 2.266 Å), in contrast with the Mn–N equatorial bonds (Mn2–N2 = 2.110 Å and Mn2–N4 = 2.058 Å).

Table 4.

Selected Bond Lengths (Å) from SC-XRD Analysis of Dimeric Mn2(III,IV) Complexes

PyN3a

bond length (Å) Mn1, d3 Mn(IV) Mn2, d4 Mn(III) OHPyN3b N4c



Mn–N1 2.086(3) 2.246(3) 2.193(15) 2.166(2)
Mn–N2 2.091(3) 2.110(3) 2.104(13) 2.085(2)
Mn–N3 2.090(3) 2.266(3) 2.185(15) 2.175(2)
Mn–N4 1.991(3) 2.058(3) 2.00(12) 2.094(2)
Mn–O1 1.809(2) 1.861(2) 1.834(12) 1.823(1)
Mn–O2 1.800(2) 1.849(2) 1.830(11) 1.812(2)
Mn1–Mn2 2.712(7) 2.709(4) 2.694(1)
a

[Mn2(μ-O)2(PyN3)2][ClO4]3, Freire et al.8

b

[Mn2(μ-O)2(OHPyN3)2][ClO4]3, this work.

c

[Mn2(μ-O)2(N4)2][Cl]3, Goodson et al.37

Electrochemistry Is an Analytical Probe To Differentiate Monomeric vs Dimeric Species.

Cyclic voltammetry (CV) was used to evaluate the electronic impact of the ligand scaffold on the manganese center as a monomeric and dimeric bis-μ-oxo complex, when possible. CV data obtained in MeCN were referenced to the redox couple of Fc+/0, while experiments performed in H2O were referenced to the redox couple of [Fe(CN)6]3−/4−. Monomeric species were analyzed in MeCN, showing a quasi-reversible and diffusion controlled redox couple (Figures 6A and S12S15). Decreasing the number of pyridine rings from [Mn(Py2N2)Cl2][BF4] to [Mn(PyN3)Cl2][ClO4] makes the complex harder to reduce, evidenced by the shift of E1/2 to less positive potentials (Table 5). This result is consistent with previous data obtained from pH-potentiometric titrations showing that Py2N2 is a weaker manganese binding ligand compared to PyN3. However, the redox potential can be fine-tuned through substitutions of the pyridine ring as evidenced by OHPyN3 and OMePyN3 showing a negative shift of E1/2 in comparison with PyN3, exhibiting the stabilizing effect of EDGs on the PyN3 scaffold (Table 5). No monomeric complex of manganese was obtained with N4, which prevented the analysis of electrochemical data of this ligand among this series.

Figure 6.

Figure 6.

Cyclic voltammograms of (A) monomeric manganese complexes in MeCN (ν = 100 mV s−1, 0.1 M TBAP), (B) dimeric bis-μ-oxo manganese complexes in H2O (ν = 100 mV s−1, 0.1 M NaClO4), (C) dimeric bis-μ-oxo manganese complexes in MeCN (ν = 100 mV s−1, 0.1 M TBAP), and (D) [Mn2(μ-O)2(N4)2][Cl]3 in MeCN with addition of H2O (ν = 100 mV s−1, 0.1 M TBAP). All scans were obtained by scanning in the cathodic direction from the open circuit potential.

Table 5.

Electrochemical Data of Monomeric and Dimeric Isolated Manganese Complexes

complexes redox wave solventa Epa (mV) Epc (mV) E1/2 (mV)
[Mn(PyN3)Cl2][ClO4] Mn(II/III) MeCN 371 217 294
[Mn(OHPyN3)Cl2][ClO4] 236 61 148
[Mn(OMePyN3)F2][BF4] 329 148 239
[Mn(Py2N2)Cl2][BF4] 484 388 436
[Mn2(μ-O)2(N4)2][Cl]3 Mn(III,IV/IV,IV) H2O 485 380 433
MeCN 510 387 448
Mn(III,IV/III,III) H2O −5 −103 −54
MeCN −279 −399 −339
[Mn2(μ-O)2(PyN3)2][ClO4]3 Mn(III,IV/IV,IV) H2O
MeCN
Mn(III,IV/III,III) H2O 193 52 123
MeCN 67 −16 26
[Mn2(μ-O)2(OHPyN3)2][ClO4]3 Mn(III,IV/IV,IV) H2O 526 401 464
MeCN
Mn(III,IV/III,III) H2O 155 −41 57
MeCN
a

Manganese complexes in MeCN were referenced vs Fc+/0, and 0.1 M TBAP was used as electrolyte. Manganese complexes in H2O were referenced vs [Fe(CN)6]3−/4−, and 0.1 M NaClO4 was used as electrolyte.

First, the dimeric bis-μ-oxo species were analyzed in water. CV results show the traditional two electrochemical waves observed in previous reports of bis-μ-oxo dimers of N4 (cyclen) and cyclam corresponding to the Mn(III,IV/III,III) reduction couple at more positive potentials, and a Mn(III,IV/IV,IV) oxidation couple at more negative potentials.38 Likewise, both of these two characteristic waves are observed for solutions of [Mn2(μ-O)2(OHPyN3)2][ClO4]3 and [Mn2(μ-O)2(N4)2][Cl]3 (Figure 6B). However, analysis of [Mn2(μ-O)2(PyN3)2][ClO4]3 only results in one reduction wave Mn(III,IV/III,III). All CVs show a quasi-reversible and diffusion controlled behavior (Figures 6B and S16S20). As the number of pyridine units decreases from PyN3 to N4, the E1/2 shifts to more negative potentials, consistent with data obtained from the monomeric complexes (Table 5). The dimer of OHPyN3 as EDG also shows the same shift observed in monomeric species to more negative potentials in comparison with the unsubstituted congener PyN3 (Table 5).

To compare electrochemical data from monomeric and dimeric species, electrochemical analysis of dimeric complexes was also performed in MeCN, obtaining quasi-reversible waves that were diffusion controlled for [Mn2(μ-O)2(PyN3)2][ClO4]3 and [Mn2(μ-O)2(N4)2][Cl]3 (Figures 6C and S21S23). [Mn2(μ-O)2(OHPyN3)2][ClO4]3 was not soluble in MeCN, which prevented the study of this complex. Overall, electrochemical waves were comparatively sharper in MeCN while maintaining the general features observed in water. Since the electrochemical waves were more broad in water, we decided to carry out an experiment for [Mn2(μ-O)2(N4)2][Cl]3 where water was added to MeCN in portions to understand the impact of water on the position of the electrochemical waves in MeCN (Figure 6D). As more water is added, the reduction wave of MnIII,IV/III,III shifts to more positive potentials, while the oxidation wave Mn(III,IV/IV,IV) shifts to more negative potentials. A prominent oxidation wave is also observed at positive potentials concomitant with the increase of water aliquots, which is consistent with water oxidation. The results revealed that water impacts both electrochemical events, making both oxidation and reduction waves easier to access.

CV experiments were then performed on manganese complexes formed in situ to assign them as mono- or dinuclear complexes (Figure 7). The setup was selected to keep the conditions as consistent as possible with the H2O2 catalytic experiments described in the next section. Complexes of manganese with the library of ligands: N4, PyN3, Py2N2, OHPyN3, and OMePyN3, were prepared in situ using standardized solutions of MnCl2 and ligands at pH = 7.5 in a H2O:MeCN ratio of 1:5 to achieve a total concentration of the complex of 5 mM. Using MeCN in the solvent mixture, the electrochemical behavior of the isolated species could be compared with those prepared in situ. Complexes prepared in situ with MnCl2 rapidly (~2 min) oxidize, which is evident by the strong color change from transparent to dark green under aerobic conditions. All complexes were allowed to mix aerobically for 10 min, and then the atmospheric air was replaced by a blanket of nitrogen prior to electrochemical measurements.

Figure 7.

Figure 7.

Cyclic voltammograms of (top) [Mn(PyN3)Cl2][ClO4] and [Mn2(μ-O)2(PyN3)2][ClO4]3 in MeCN, highlighting the difference between the Epc waves (ν = 100 mV s−1, 0.1 M TBAP); (bottom) manganese complexes prepared in situ in a solvent mixture of H2O:MeCN 1:5 (ν = 100 mV s−1, 0.1 M TBAP).

The redox couple of manganese complexes in situ was assigned using electrochemical data of isolated complexes (Table 6) as the reduction wave of bis-μ-oxo complexes, Mn(III,IV/III,III). The cathodic wave of Mn/N4 is shifted to more negative potentials (Epc = −343 mV) in comparison with the cathodic wave of [Mn2(μ-O)2(N4)2][Cl]3 (Epc = −399 mV), which is consistent with the shift of the reduction wave to positive potentials observed for [Mn2(μ-O)2(N4)2][Cl]3 after addition of water (Figure 6D). Mn/PyN3 shows electrochemical activity on a region of the chemical potential (E1/2 = 26 mV) that matches the reduction wave of [Mn2(μ-O)2(PyN3)2][ClO4]3 (E1/2 = 26 mV).

Table 6.

Electrochemical Data of Manganese Complexes Prepared In Situa

complexes Epa (mV) Epc (mV) E1/2 (mV)
Mn/PyN3 90 −39 26
Mn/OMePyN3 72 −98 −13
Mn/OHPyN3 −73 −205 −139
Mn/N4 −171 −343 −257
a

Manganese complexes were dissolved in H2O:MeCN (1:5) and referenced vs Fc+/0 using 0.1 M TBAP as electrolyte.

Addition of a pyridine ring in the macrocycle going from Mn/N4 to Mn/PyN3 shifts the E1/2 to more positive potentials (Table 5), following the same trend observed for [Mn2(μ-O)2(N4)2][Cl]3 and [Mn2(μ-O)2(PyN3)2][ClO4]3 (Table 6). Consistent with isolated complexes, EDGs of manganese complexes prepared in situ (–OMe and –OH substituents) also shifted the E1/2 to more negative potentials. It was also observed a prominent oxidation wave at ~120 mV that grew as the potential was swapped to more positive values; we believed that this might have prevented the detection of Mn/Py2N2 which was expected to show at more positive potentials than Mn/PyN3, based on the measurements done for [Mn(PyN3)Cl2][ClO4] and [Mn(Py2N2)Cl2][BF4] (Table 6).

Comparison of redox waves of isolated crystals vs complexes prepared in situ highlights that electrochemical measurements can serve as an analytical probe to differentiate monomeric vs dimeric species of manganese when the chemical potential at which they interact with is sufficiently distinct (201 mV difference in the present study for manganese complexes of PyN3).

Catalytic H2O2 Disproportionation Studies.

Kinetic studies were carried out to quantify the effect of the ligand structure on the performance of manganese to catalyze disproportionation of H2O2 into O2 and H2O. In our previous report, the ability to disproportionate H2O2 was quantified with manganese complexes of N4 and PyN3 in terms of turn over number (TON) and turn over frequency (TOF).8 Nevertheless, a comparison of our results with the literature is challenging due to the fact that both TON (mol of product/mol of catalyst) and TOF (TON/time) depend on the concentration of substrate and/or catalyst, temperature, pressure, and time frame selected, which are not parameters homogeneous in the literature.39 Instead, comparative studies of MnCAT biomimetics typically show the second-order kinetic constant (M−1 s−1), denoted as k (initial rates method, ri) or kcat/KM (Michaelis–Menten model).3,5,6,40 In this study, we use the kinetic constant k from initial rates method to describe catalytic activity and employ TON to quantify the longevity and overall robustness of the catalyst.

The ability of manganese in the presence of N4, PyN3, OHPyN3, OMePyN3, and Py2N2 to catalyze the disproportionation of H2O2 into O2 and H2O was tested using a one-pot approach to generate manganese complexes in situ, at a pH value optimal among all the ligands for complex formation based on data from potentiometric titrations (Table 2 and Figures S1S5). The standardized reagent solutions were added to a starting volume of ultrapure water in the following order: ligand, buffer (Tris, pH = 8.0), and then MnCl2. This one-pot approach avoided complications related to (1) the rapid formation of inactive manganese oxide/hydroxide species from the isolated complexes, and (2) the need to adjust pH after addition of MnCl2, which would affect the relative concentration of the manganese complex in solution. In our previous study, we demonstrated that neither MnCl2 (buffered or unbuffered) or the ligand produce any catalase activity under the conditions of our experiment (Figure S28).8 Isolated complexes presented similar reactivity to species in situ, thus were preferred to streamline the catalytic experiments in this latest study. This work also showed that different mixing times of MnCl2 and ligands (instantaneous, 6 h, or overnight) followed by addition of H2O2 did not provide a difference in catalase reactivity based on the TON recorded.

Results regarding the number of pyridine rings showed significant changes for the catalytic activity measured with k among the ligands: N4 (0.6 M s−1) ≪ PyN3 (1.8 M s−1) ≫ Py2N2 (0.5 M s−1) (Table 7 and Figure 8A). Within this group, PyN3 possess the right scaffold for efficiently catalyze H2O2 disproportionation in comparison with N4 and Py2N2.

Table 7.

Kinetic and Reactivity Results of the Disproportionation of H2O2 Using Buffered Aqueous Solutions Containing MnCl2 and the Respective Macrocyclic Ligand ([MnCl2]0 = 1.50 mM, [ligand]0 = 1.53 mM, [Tris] = 50 mM, pH = 8, and T = 298 K)

N4 PyN3 OHPyN3 OMePyN3 Py2N2
k (M−1 s−1) 0.60(3) 1.8(1) 2.2(1) 2.3(1) 0.5(1)
TONa 11.26(4) 24.67(3) 36.89(1) 37.31(2) 28.85(3)
a

TON values were calculated at 20 min after the initial injection of H2O2, and results are representative of an average of five individual samples.

Figure 8.

Figure 8.

(A,B) Second-order kinetic constant, k, and (C,D) TON values of the catalytic disproportionation of H2O2 using buffered aqueous solutions containing MnCl2 and the respective macrocyclic ligand ([MnCl2]0 = 1.50 mM, [ligand]0 = 1.53 mM, [Tris] = 50 mM, pH = 8, and T = 298 K).

TON values (Figure 8B) indicate that the Py2N2 ligand forms the most robust manganese catalyst regarding the series with different number of pyridine rings. Addition of H2O2 causes a steady release of H+, evidenced by measurements of pH in unbuffered solutions throughout the reaction time. This change eventually leads to dissociation of the manganese complex as evidenced by mass spectrometry measurements (Figures S36S40). Py2N2 is able to consistently sequester Mn(II) ions in a wider range of pH (pMn = 5.96), which could explain the higher TON observed with this complex.

Results regarding substitution on the 4-position of the pyridine ring revealed that the unsubstituted PyN3 ligand shows a less reactive catalyst when bound to manganese (k = 1.8 M s−1, TON = 24.7) in comparison to OHPyN3 (k = 2.2 M s−1, TON = 36.9) and OMePyN3 (k = 2.3 M s−1, TON = 37.3) (Table 7 and Figure 8C,D). The EDG substitutions on PyN3 (–OH and –OMe) stabilize higher oxidation states of manganese, based on CV studies performed on species in situ (E1/2), increasing the donor strength of the N atom in the pyridine ring closer to the strength of the secondary amines as in N4 (Figure 7), while retaining the conformational reorganization energy of the PyN3 scaffold. EDGs on the pyridine ring show that the electronic properties of the macrocycle offer a handle to fine-tune the reactivity (k and TON) without considerable changes to the coordination environment, a property that has also been observed with similar macrocycles bound to Fe(III) and Cu(II).33,35 Therefore, inclusion of EDGs was also hypothesized to increase the catalytic activity of a given manganese complex, by stabilizing key high-valent state intermediates, which have been postulated to play an important role in the mechanism of reactions of this type.41,42 While MnCAT achieves catalytic reactivity by modulating the protein environment surrounding the manganese binuclear center,43 small molecules depend on the thermodynamic stability of complex formation and electronic properties to modulate the reactivity of the metal center. The results presented in this section show that addition of pyridine rings and their modification with EDGs must play a role in the stabilization of high valent manganese intermediates that may be present in the catalytic cycle, which have been reported in mechanistic studies with heme and nonheme catalase biomimetics.3,44

An induction time at the onset of the reaction was observed after injection of H2O2, which motivated the construction of plots of O2 production resulting from addition of H2O2 to solutions containing manganese complexes of N4, PyN3 and Py2N2 (Figure 9). This induction time for both N4 and PyN3 was roughly 30 s after injection of H2O2, while the time for Py2N2 was 2 min. These observations show that the manganese complex initially injected must go through steps to produce the active form of the catalyst to perform H2O2 disproportionation.

Figure 9.

Figure 9.

Time course of partial pressure of O2 (PO2) after injecting H2O2 to buffered aqueous solutions containing MnCl2 and ligands (A) N4, (B) PyN3, and (C) Py2N2 ([MnCl2]0 = 1.50 mM, [ligand]0 = 1.53 mM, [Tris] = 50 mM, pH = 8, T = 298 K).

The effect of the temperature on the rate constant was investigated between 5 and 25 °C to gather thermodynamic data regarding the activation parameters. The activation energy Ea and pre-exponential factor (A) were calculated using the Arrhenius model, as shown in the following equation.

kt=AeEa/RT

The plot of ln kobs vs T1 for each manganese complex was used to obtain the Ea from the slope of the linear regression and A from the intercept (Table 8 and Figures S34S38).

Table 8.

Rate Constants (kobs) at Different Temperatures, Activation Energy (Ea), and Pre-Exponential Factor (A)

temperature (°C) kobs (s−1) Ea (kcal mol−1) A (s−1)
N4 5 0.00084(2) 1.1(2) 0.005(7)
10 0.00086(2)
15 0.00091(8)
25 0.00094(4)
PyN3 5 0.00163(7) 4.5(5) 5.4(2)
10 0.00185(8)
15 0.00202(5)
25 0.00283(2)
Py2N2 5 0.00019(2) 13.4(4) 6.5 × 106(8)
10 0.00024(1)
15 0.00068(3)
25 0.00086(6)
OHPyN3 5 0.00192(5) 4.2(5) 3.9(3)
10 0.00213(7)
15 0.00230(1)
25 0.00321(1)
OMePyN3 5 0.00187(1) 4.6(3) 7.3(2)
10 0.00217(1)
15 0.00263(1)
25 0.00324(2)

The Ea calculated from kinetic data using the Arrhenius equation (Table 8) showed an increment as the number of pyridine rings increases from N4 (Ea = 1.1 kcal mol−1) < PyN3 (Ea = 4.5 kcal mol−1) < Py2N2 (Ea = 13.4 kcal mol−1), showing how the reorganization energy of the macrocyclic ligand might play a role in achieving the desired reactivity.

CONCLUSIONS

This report details the isolation and characterization of a novel library of manganese complexes comprised of 12-membered tetra-aza macrocyclic ligands that bind in a monomeric or bis-μ-oxo dimers to the metal center. The structure of the ligand was selected to evaluate the impact of the number of pyridine rings as well as EDGs bound to the pyridine moiety. Ligand basicity showed a decrease with the number of pyridine rings, but substitutions did not show a straightforward trend. –OH substitution showed an increase in the basicity of the ligand, but tautomerism keto–enol disclosed an important factor in the complexation of Mn(II) ions, presenting a shift of the mesomeric effect of the –OH group. On the other hand, –OMe substitution did not reflect an electron-donating character on the basicity of the ligand, showing a decrease of basicity in comparison with the parent PyN3 unsubstituted ligand.

The basicity of the ligand played a fundamental role in equilibrium studies performed with Mn(II) ions in aqueous solutions and allowed us to determine the best strategy to synthesize and isolate the corresponding complexes. Depending on the structure of the ligand, some complexes were isolated as monomeric structures or as bis-μ-oxo dimers as shown by SC-XRD. The features of isolated species in solution were evaluated using UV–vis spectroscopy and cyclic voltammetry. Electrochemistry was fundamental to characterize the species observed in solution when manganese complexes were prepared in situ since the potential for monomer vs dimer species is separated by a potential window of 201 mV.

Species in situ were used for catalytic disproportionation of H2O2, showing similar results to isolated species. TON values increased consistently with the number of pyridine rings, but the kinetic constant k favored the PyN3 system. Py2N2 showed a decreased of k and a longer induction period. On the other hand, substitutions on the 4-position of the pyridine ring favored both TON and k to a bigger extent, highlighting the importance of these substitution on the reactivity of manganese.

We hope that this study stimulates the field into venturing in the reactivity of dimers of manganese that expose structural properties required for the desired reactivity. Learning how to tune the reactivity of H2O2 with manganese catalysts as an oxidant for catalysis or as a substrate for disproportionation would greatly aid in the field of oxidation catalysis. More importantly, studies with similar catalysts under air atmosphere would facilitate their implementation in industrial applications.

EXPERIMENTAL SECTION

General Methods.

Caution! Perchlorate salts are explosive and should be handled with care; such compounds should never be heated as solids. All chemical reagents were purchased from either Millipore Sigma or Alfa Aesar and used without further purification unless otherwise indicated. Combustion elemental analysis was performed by Atlantic Microlab in Norcross, GA. Electronic absorption spectra were collected between 190 and 1100 nm using a Cary 60 spectrophotometer (Agilent) and a 3 mL quartz cuvette with a path length of 1.0 cm. Molar extinction coefficients were calculated using the Beer–Lambert law (A = εbc). Electrospray ionization mass spectrometry (ESI-MS) was performed on an Advion ExpressionL CMS instrument.

Synthesis of Ligands.

OHPyN3 was prepared following the synthetic procedure of Lincoln et al., with no further modifications.14 PyN3, OMePyN3, and Py2N2 were prepared by adapting procedures reported previously.13,15,4552 Scheme S1 shows the overall synthetic procedure. Modified methods are described further herein. Chelidamic acid, 3b, and 10 were obtained from commercial sources.

Synthesis of 1.

Small modifications to the previously reported methods were used.4547 Thionyl chloride (27.2 mL, 44.4 g, 373 mmol) was added dropwise to a suspension of chelidamic acid monohydrate (25.0 g, 124 mmol) in absolute EtOH (250 mL) at 0 °C. The mixture was allowed to warm at room temperature under stirring (15 min) and then heated at 80 °C for 4h. The solution was concentrated under reduced pressure as a eutectic mixture with toluene (3×, 100 mL), using solid Na2CO3 in the receiving flask. An oil was obtained and mixed with H2O:Et2O (1:1, 300 mL) to precipitate a brown solid, which was thoroughly washed with Et2O. A flash column was performed (SiO2, CH2Cl2:MeOH = 98:2) to obtain 1 (22.2 g, 93.0 mmol, 75%) as a yellow solid. 1H NMR spectrum was consistent with the literature values.

Synthesis of 5a and 5b.

These compounds were synthesized by following the reported literature methods with slight modifications.13,49,50 11 (5.71 g, 10.1 mmol) and CsCO3 (13.2 g, 40.4 mmol) were mixed together in DMF (110 mL). A solution of either 4a or 4b (4.52 g, 10.1 mmol) in DMF (50 mL) was added dropwise. The reaction mixture was stirred at room temperature for 24 h. The solvent was removed under reduced pressure as a eutectic mixture with toluene (3x) and dried in vacuo overnight. The resultant solid was redissolved in H2O (150 mL) and extracted with CHCl3 (3×, 150 mL). The combined organic layers were washed with brine (150 mL), dried over Na2SO4, and concentrated under reduced pressure. The residue was purified by column chromatography (SiO2, CH2Cl2:EtOAc = 40:1) to afford either 5a (5.79 g, 8.28 mmol, 82%) or 5b (5.74 g, 8.58 mmol, 85%) as a white foamy solid. Rf = 0.4 (SiO2, CH2Cl2:EtOAc = 40:1). 1H NMR spectrum was consistent with the literature values.

Synthesis of 6a (OMePyN3·3HCl) and 6b (PyN3·3HCl).

These compounds were synthesized by following the reported literature methods with slight modifications.13,49 A solution of either 5a (5.61 g, 8.03 mmol) or 5b (5.37 g, 8.03 mmol) in concentrated H2SO4 (15 mL) was stirred at 100 °C for 4 h. The resultant dark brown solution was cooled down to room temperature, followed by addition of H2O (75 mL). The acidic mixture was extracted with Et2O (1×, 75 mL), and the aqueous layer was neutralized with a 60% m/v solution of NaOH until reaching pH = 14. The product was extracted with CH2Cl2 (3×, 100 mL). The combined organic layers were washed with brine (100 mL), dried over Na2SO4, and concentrated under reduced pressure. The resultant yellow oil was mixed with concentrated HCl (1 mL), followed by CH3CN (5 mL) and Et2O (100 mL). The mixture was sonicated (20 min) and left in the freezer for 2 h. The precipitate was filtered, washed with Et2O, and dried under vacuum to obtain either 6a (2.22 g, 6.42 mmol, 80%) or 6b (2.15 g, 6.82 mmol, 85%) as a white solid. 1H NMR spectrum was consistent with the literature values. X-ray quality crystals of 6a were also obtained by slow evaporation of MeOH as translucid colorless cubes (CCDC #2245877).

Synthesis of 7 and 8.

These compounds were synthesized by following the reported literature methods with slight modifications.15 4b (4.00 g, 8.94 mmol) and TsNHNa (3.44 g, 17.8 mmol) were mixed and suspended in 600 mL of MeCN, followed by reflux under stirring for up to 7 days. The solvent was evaporated under reduced pressure. The resultant solid was washed with cold MeOH and filtrated to obtain a mixture of tosyl-protected dimer and trimer macrocycle. The dimeric macrocycle 8 (1.18 g, 2.01 mmol, 45%) was separated using the procedure reported by Wessel et al., with no further modifications.15 HH NMR spectrum was consistent with the literature values.

Synthesis of 11.

This compound was synthesized by following the reported literature methods with slight modifications.51,52 Compound 10 (9.21 g, 89.3 mmol) was dissolved in H2O (55 mL) and NaOH powder (10.7 g, 268 mmol) was added in small portions during a 30 min time frame, keeping the temperature of the solution at 20 °C. Et2O (55 mL) was added, and the reaction mixture was stirred vigorously. TsCl (51.1 g, 268 mmol) was added while cooling the reaction mixture to 0 °C and then stirred for 2h. A white precipitate was filtered and washed with Et2O (110 mL). Recrystallization from MeOH gave the desired product 11 (50.0 g, 88.4 mmol, 99%) as a white solid. 1H NMR spectrum was consistent with the literature values.

Potentiometric Measurements.

The MnCl2 stock solution was prepared from analytical grade commercial sources, and its concentration was determined by complexometric titration with standardized Na2H2EDTA and an EBT indicator in the presence of ascorbic acid and triethanol-amine. The concentration of the ligand stock solution was determined by pH-potentiometric titration. The protonation constants were calculated with the following setup: 0.2 M carbonate-free NaOH titrant with 2 mM ligand solution at an initial volume of 6 mL. Titrations were performed at 25.0 ± 0.1 °C and an ionic strength of I = 0.15 M NaCl. An inert atmosphere was provided by a constant passage of dry N2 through the sample. The protonation constants of the ligand logKiH are defined as

KiH=HiLi+[Hi1L(i1)+][H+]

where i=1,2, and 3 and [Hi1L(i1)+] and H+ are the equilibrium concentrations of the ligand in different protonation states and hydrogen ions, respectively. Potentiometric titrations were carried out with a Metrohm 888 Titrando workstation and a Metrohm 6.0234.100 combined electrode in the pH range of 1.7–11.8. For the calibration of the electrode, KH-phthalate (pH = 4.008)53 and borate (pH = 9.177)54 buffer standards were used, and the H+ was calculated from the measured pH values by applying the method proposed by Irving et al.55 A solution of approximately 0.01 M HCl was titrated with 0.2 M NaOH solution (I = 0.15 M NaCl), and the differences between the measured and calculated pH values (pH < 2.2) were used to calculated the H+ from the pH values measured in the ligand titrations. The experimental points above pH 11 for the acid–base titration were used to determine the ionic product of water (13.840) in case of our experimental setup. PSEQUAD software was used to calculate the equilibrium constants.56

The protonation constants for the ligands were determined by titrating the ligand solution (acidified with a known volume of a standard HCl solution) with 0.2 M NaOH at 0.15 M NaCl ionic strength in the 1.7–11.8 pH range. Equilibrium points 1 and 2 (EP1 and EP2, respectively) were separated using MnCl2. The logKiH values were calculated from 150 V (mL)−pH data pairs. The initial proton concentration H+0 submitted in PSEQUAD was obtained from the sum of the moles of H+ added from standardized solutions of HCl and ligand solution.

To determine the stability constants of the manganese complexes, potentiometric titrations were carried out at a 1:1 metal to ligand molar ratio, with 2% ligand excess to prevent the hydrolysis of Mn(II), allowing 1 min for the equilibration to occur, with 150 V (mL)−pH data pairs. The initial proton concentration H+0 submitted in PSEQUAD was calculated using the following equation.

H+0=EP1ML+VLVMLΔEPL[NaOH]Vo,ML

where the subindexes L and ML represent values derived from the ligand and metal–ligand titration, respectively. VL and VML is the volume of ligand in each titration. Vo is the initial volume. EP1ML corresponds to the first equilibrium point in the ML titration, and ΔEPL corresponds to the difference between first and second equilibrium points in the ligand titration. This equation allows a more accurate determination of the [H+]0 by taking into account the 2% excess of ligand used, in the case of metal complexes that do not present changes in EP1 between ligand and metal–ligand titration.

Synthesis of Manganese Complexes.

[Mn2(μ-O)2(N4)2][Cl]3, [Mn(PyN3)Cl2][ClO4], and [Mn2(μ-O)2(PyN3)2][ClO4]3 were synthesized following previously reported procedures with no further modifications.8,38

Synthesis of [Mn(OHPyN3)Cl2][ClO4].

An adapted procedure based on work reported elsewhere was used for the synthesis of this compound.8 OHPyN3·3HCl (100 mg, 0.302 mmol) was dissolved in 5 mL of H2O. The pH of the solution was adjusted to 8 with a solution of KOH (1 M). A solution of Mn(ClO4)2·xH2O (75.1 mg, 0.296 mmol) in H2O (5 mL) was added and stirred at room temperature for 12 h. The solution color gradually changed from pale yellow to dark green. The resultant mixture was filtered, and the solvent was removed under reduced pressure. The resultant solid was redissolved in MeCN (20 mL) and stirred at room temperature for 12 h. A bright pink solution was obtained, which was filtered to remove solid impurities and the solvent was removed under reduced pressure to give the desired product [Mn(OHPyN3)Cl2][ClO4] (99.4 mg, 0.222 mmol, 75%) as a pink powder. *The strong propensity of this class of compounds to form μ-oxo dimers and MnO2 under oxidizing conditions, i.e., synthesis or combustion analysis, has continued to result in mixtures of compound despite extensive efforts.

Synthesis of [Mn2(μ-O)2(OHPyN3)2][ClO4]3.

An adapted procedure based on work reported elsewhere was used for the synthesis of this compound.8 OHPyN3·3HCl (100 mg, 0.302 mmol) and Mn(ClO4)2·xH2O (75.1 mg, 0.296 mmol) were dissolved in H2O (10 mL). The pH of the solution was adjusted to 8 with a solution of KOH (1 M) and subsequently stirred at room temperature for 12 h. The solution color gradually changed from pale yellow to dark green. The resultant mixture was filtered, and the solvent was removed under reduced pressure. The resultant solid was redissolved in MeCN (20 mL) and stirred at room temperature for 12 h. A dark green solution was obtained, which was filtered to remove solid impurities and the solvent was removed under reduced pressure to give the desired product [Mn2(μ-O)2(OHPyN3)2][ClO4]3 (58.2 mg, 0.0630 mmol, 42%) as a green powder. The product was redissolved in H2O and filtered (PTFE, 0.45 μm) to obtain X-ray quality crystals by slow evaporation (CCDC #2241016).

Synthesis of [Mn(OMePyN3)F2][BF4].

An adapted procedure based on work reported elsewhere was used for the synthesis of this compound.8 OMePyN3·3HCl (100 mg, 0.289 mmol) was dissolved in 5 mL of H2O. The pH of the solution was adjusted to 7 with a solution of KOH (1 M). A solution of MnCl2·4H2O (56.0 mg, 0.283 mmol) in H2O (5 mL) was added and stirred at room temperature for 12 h. The solution color gradually changed from pale yellow to dark green. The resultant mixture was filtered, and the solvent was removed under reduced pressure. The resultant solid was redissolved in MeCN (20 mL) and stirred at room temperature for 12 h. A bright pink solution was obtained, which was filtered to remove solid impurities and the solvent was removed under reduced. The product was redissolved in MeOH (10 mL), mixed with AgBF4 (55.1 mg, 0.283 mmol), and stirred for 2 h at room temperature to obtain X-ray quality crystals. The pink solution was filtered, and the solvent was removed under reduced pressure to give the desired product [Mn(OMePyN3)F2][BF4] (17.5 mg, 0.042 mmol, 15%) as a pink powder. The product was redissolved in MeOH and filtered (Nylon, 0.45 μm) to obtain X-ray quality crystals by slow diffusion of Et2O. Elemental analysis; [Mn(OMePyN3)F2][BF4] found (calculated): C, 34.87 (34.64); H, 4.88 (4.85); N, 13.05 (13.47) % (CCDC #2205652).

Synthesis of [Mn(Py2N2)Cl2][BF4].

An adapted procedure based on work reported elsewhere was used for the synthesis of this compound.8 Py2N2 (100 mg, 0.416 mmol) was dissolved in 5 mL of H2O. The pH of the solution was adjusted to 7 with a solution of KOH (1 M). A solution of MnCl2·4H2O (80.7 mg, 0.408 mmol) in H2O (5 mL) was added and stirred at room temperature for 12 h. The solution color gradually changed from pale yellow to dark green. The resultant mixture was filtered, and the solvent was removed under reduced pressure. The resultant solid was redissolved in MeCN (20 mL) and stirred at room temperature for 12 h. A bright pink solution was obtained, which was filtered to remove solid impurities and the solvent was removed under reduced pressure. The product was redissolved in MeOH (10 mL), mixed with AgBF4 (79.4 mg, 0.408 mmol), and stirred for 2 h at room temperature to obtain X-ray quality crystals. The pink solution was filtered, and the solvent was removed under reduced pressure to give the desired product [Mn(Py2N2)Cl2][BF4] (145 mg, 0.184 mmol, 45%) as a pink powder. The product was redissolved in MeOH and filtered (Nylon, 0.45 μm) to obtain X-ray quality crystals by slow diffusion of Et2O. Elemental analysis; Mn[(Py2N2)F2][BF4]·1.9MnCl2·3.5MeOH·H2O found (calculated): C, 26.58 (26.63); H, 3.95 (4.09); N, 7.16 (7.10) % (CCDC #2207371).

X-ray Crystallography.

A Leica MZ 75 microscope was used to identify samples suitable for analysis. A Bruker APEX-II CCD diffractometer was employed for crystal screening, unit cell determination, and data collection, which was obtained at 100 K. The Bruker D8 goniometer was controlled using the APEX3 software suite.57 The samples were optically centered with the aid of a video camera so that no translations were observed as the crystal was rotated through all positions. The X-ray radiation employed was generated from a MoKα sealed X-ray tube (λ = 0.71076 Å) with a potential of 50 kV and a current of 30 mA; fitted with a graphite monochromator in the parallel mode (175 mm collimator with 0.5 mm pinholes). The crystal-to-detector distance was set to 50 mm, and the exposure time was 10 s per degree for all data sets at a scan width of 0.5°. The frames were integrated with the Bruker SAINT Software package using a narrow frame algorithm.58 Data were corrected for absorption effects using the multiscan method SADABS.59 Structures were solved using Olex260 with the ShelXS61 solution program using Direct Methods and refined with the SHELXL62 refinement package using least squares minimization. All hydrogen and non-hydrogen atoms were refined using anisotropic thermal parameters. The thermal ellipsoid molecular plots (50%) were produced using Olex2.60

Electrochemistry.

Cyclic voltammetry (CV) was carried out with an EC Epsilon potentiostat (C-3 cell stand) purchased from BASi Analytical Instruments (West Lafayette, IN). A glassy carbon (GC) electrode from BASi (MF-2012), 3 mm in diameter, was polished on a white nylon pad (BASi MF-2058) with different sized diamond polishes (15, 6, and 1 μm) to ensure a mirror-like finish between each measurements. All solutions were bubbled with N2 for at least 15 min prior to experimentation and kept under a N2 blanket.

For electrochemical analysis of [Mn(PyN3)Cl2][ClO4], [Mn(OHPyN3)Cl2][ClO4], [Mn(OMePyN3)F2][BF4], and Mn[(Py2N2)-F2][BF4], a three-electrode cell configuration was used: GC as the working electrode, Ag wire as the quasi-reference electrode housed in a glass tube (7.5 cm × 5.7 mm) and isolated from the analyte with a porous CoralPor tip, and a Pt wire (7.5 cm) as the counter electrode (BASi MW-1032). For each electrochemical analysis, samples were dissolved in anhydrous MeCN containing 0.1 M tetrabutylammonium perchlorate (TBAP) as supporting electrolyte. The concentration of each sample was 1 mg/mL in aliquots of 5 mL.

For electrochemical analysis of [Mn2(μ-O)2(N4)2][Cl]3, [Mn2(μ-O)2(PyN3)2][ClO4]3, and [Mn2(μ-O)2(OHPyN3)2][ClO4]3, a three-electrode cell configuration was used: GC as the working electrode, a Ag/AgCl reference electrode, and a Pt wire (7.5 cm) as the counter electrode (BASi MW-1032). For each electrochemical analysis, samples were dissolved in H2O containing 0.1 M NaClO4 as supporting electrolyte. The concentration of each sample was 1 mg/mL in aliquots of 5 mL.

For electrochemical analysis of Mn/N4, Mn/PyN3, Mn/OHPyN3, and Mn/OMePyN3, a three-electrode cell configuration was used: GC as the working electrode, Ag wire as the quasi-reference electrode housed in a glass tube (7.5 cm × 5.7 mm) and isolated from the analyte with a porous CoralPor tip, and a Pt wire (7.5 cm) as the counter electrode (BASi MW-1032). For each electrochemical analysis, samples were prepared in situ by mixing the corresponding amounts of stock solution of ligand and MnCl2·4H2O (Merck) to achieve a total concentration of 20 mM in 0.5 mL of H2O. Subsequently, the pH of each sample was adjusted to 7.5 with solid KOH and stirred aerobically for 20 min where it was observed a color change from transparent to dark green overtime. Finally, 4.5 mL of anhydrous MeCN was added to each solution to achieve a total concentration of 2 mM in 5 mL aliquots of a mixed solvent H2O:MeCN (1:5). 0.1 M TBAP was used as supporting electrolyte. Attempts to assign these manganese species were performed using UV–vis spectroscopy in our previous study, but there was ambiguity due to the similarity of the ~400 nm region between both species, and the difficulty associated with the stability of the band at ~550 nm, characteristic of the dimer, before additions of H2O2.8

Reaction of Manganese Complexes with H2O2 and Quantification of O2 Evolution.

O2 evolution was quantified as partial pressure (P) by a Clark-type commercial microsensor (Unisense, Denmark), calibrated using N2 and air (PO2 = 0 and 159 mmHg) under atmospheric pressure. The reactions were performed in a sealed pressure vessel (15 mL) equipped with a stirring bar and a three-way manifold valve. Two of the lines (PTFE tubing) were connected to the cell: one was used for N2 and the other to inject the aliquot of H2O2. The pressure vessel was sealed using a PEEK bushing connected to threaded fittings (IDEX, Health and Science) for custom sealing between the gas tubing, microsensor, and reaction vessel. The pressure of the vessel was kept at 1 atm using a snorkel line with minimal loss of headspace during the catalytic studies.

The metal complexes were prepared in buffered aqueous solutions using stock solutions of the ligand of interest, buffer, and a metal solution of MnCl2·4H2O (Merck). Tris(hydroxymethyl) amino methane (Tris) was the buffer selected for the desired pH of the solution (pH = 8). The complexes were prepared using a one-pot approach by adding the corresponding amount of the stock solution of ligand, buffer, and metal salt in a vial to achieve the targeted concentration for the H2O2 studies ([ligand] = 2.04 mM; [Mn] = 2 mM; [buffer] = 50 mM). The vessel was loaded with 1.5 mL of the solution detailed above and purge with N2 gas prior to every measurement. The microsensor signal was read continuously and measured every 0.2 s until a steady signal was obtained for 2 min, and then 0.5 mL of H2O2 solution was injected (600 mM, 50 mM buffer). The signal was recorded as ΔPO2 (mmHg) vs time (min) until a plateau was achieved. Initial concentrations before starting the reaction were [ligand] = 1.53 mM (3.06 μmol), [Mn] = 1.5 mM (3 μmol), and [H2O2] = 150 mM (300 μmol).

Once the reaction was complete, the results were plotted to determine the turnover number (TON). The reaction was stopped when a visible plateau in the PO2 was observed, which happened at roughly 20 min for all experiments. PO2 was converted into [O2] using the Ideal Gas Law. TON was calculated as the ratio between the moles of O2 produced during the time frame selected divided by moles of catalyst used. Kinetic measurements were carried out under the same conditions previously described using the initial rates method (ri). At constant [Cat]0 = 1.5 mM (Cat = any given manganese complex prepared in situ), the reaction exhibited first-order kinetics with respect to [H2O2]0, and the pseudo-first-order rate constant (kobs) was obtained from the slope of plots of ri vs [H2O2]0. Similar results were obtained with isolated complexes, but complexes prepared in situ were preferred since the approach was streamlined and easier to reproduce.

At constants [H2O2]0 = 150 mM, the reaction also exhibited a good linear dependence with respect to [Cat]0, affording a pseudo-first-order rate constant (kobs′) (Figures S29S34). Since the conditions described showed first-order rates with respect to each species (Cat and H2O2), the second-order catalytic constant (k) was obtained by either dividing kobs/[Cat]0 or kobs′/[H2O2]0 and further validated by the slope of either a plot of ri/[H2O2]0 vs [Cat]0 or a plot of ri/[Cat]0 vs [H2O2]0, showing k values in the same order of magnitude. Catalytic experiments were performed at 5, 10, 15, and 25 °C to calculate activation energy (Ea) for each catalyst using the Arrhenius equation.63

Supplementary Material

SI

ACKNOWLEDGMENTS

The authors would like to thank Dr. Benjamin Sherman, Texas Christian University, for the use of the UniSense oxygen probe and related resources. The authors thank Dr. Alexander Lippert, Southern Methodist University, for the use of a mass spectrometer. The authors also acknowledge generous support from National Institutes of General Medical Sciences 2R15GM123463, ACS PRF 65400-ND3, and Welch Foundation P-2063-20210327.

Footnotes

The authors declare no competing financial interest.

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.inorgchem.3c01234

ASSOCIATED CONTENT

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c01234.

Species distribution diagrams, X-ray diffraction data and results, cyclic voltammetry, kinetic plots, and mass spectrometry (PDF)

Accession Codes

CCDC 2205652, 2207371, 2241016, and 2245877 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; fax: +44 1223 336033.

REFERENCES

  • (1).Pereira MM; Dias LD; Calvete MJF Metalloporphyrins: Bioinspired Oxidation Catalysts. ACS Catal. 2018, 8 (11), 10784–10808. [Google Scholar]
  • (2).Chelikani P; Fita I; Loewen PC Diversity of structures and properties among catalases. Cell. Mol. Life Sci. 2004, 61 (2), 192–208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Signorella S; Palopoli C; Ledesma G Rationally designed mimics of antioxidant manganoenzymes: Role of structural features in the quest for catalysts with catalase and superoxide dismutase activity. Coord. Chem. Rev. 2018, 365, 75–102. [Google Scholar]
  • (4).Barynin VV; Whittaker MM; Antonyuk SV; Lamzin VS; Harrison PM; Artymiuk PJ; Whittaker JW Crystal Structure of Manganese Catalase from Lactobacillus plantarum. Structure 2001, 9 (8), 725–738. [DOI] [PubMed] [Google Scholar]
  • (5).Ledesma GN; Anxolabéhère-Mallart E; Sabater L; Hureau C; Signorella SR Functional modeling of the MnCAT active site with a dimanganese(III) complex of an unsymmetrical polydentate N3O3 ligand. J. Inorg. Biochem. 2018, 186, 10–16. [DOI] [PubMed] [Google Scholar]
  • (6).Solís V; Palopoli C; Daier V; Riviere E; Collin F; Moreno DM; Hureau C; Signorella S Tuning the MnII2/MnIII2 redox cycle of a phenoxo-bridged diMn catalase mimic with terminal carboxylate donors. J. Inorg. Biochem. 2018, 182, 29–36. [DOI] [PubMed] [Google Scholar]
  • (7).Brewer SM; Schwartz TM; Mekhail MA; Turan LS; Prior TJ; Hubin TJ; Janesko BG; Green KN Mechanistic Insights into Iron-Catalyzed C–H Bond Activation and C–C Coupling. Organometallics 2021, 40 (15), 2467–2477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Freire DM; Beeri D; Pota K; Johnston HM; Palacios P; Pierce BS; Sherman BD; Green KN Hydrogen peroxide disproportionation with manganese Macrocyclic complexes of cyclen and pyclen. Inorg. Chem. Front. 2020, 7 (7), 1573–1582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Siegbahn PEM A quantum chemical study of the mechanism of manganese catalase. Theor. Chem. Acc. 2001, 105 (3), 197–206. [Google Scholar]
  • (10).Siegbahn PEM Modeling aspects of mechanisms for reactions catalyzed by metalloenzymes. J. Comput. Chem. 2001, 22 (14), 1634–1645. [Google Scholar]
  • (11).Siegbahn PEM Quantum chemical studies of manganese centers in biology. Curr. Opin. Chem. Biol. 2002, 6 (2), 227–235. [DOI] [PubMed] [Google Scholar]
  • (12).Yepremyan A; Mekhail MA; Niebuhr BP; Pota K; Sadagopan N; Schwartz TM; Green KN Synthesis of 12-Membered tetra-aza Macrocyclic Pyridinophanes Bearing Electron-Withdrawing Groups. J. Org. Chem. 2020, 85 (7), 4988–4998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Fan R; Serrano-Plana J; Oloo WN; Draksharapu A; Delgado-Pinar E; Company A; Martin-Diaconescu V; Borrell M; Lloret-Fillol J; Garcia-Espana E; et al. Spectroscopic and DFT Characterization of a Highly Reactive Nonheme Fe(V)-Oxo Intermediate. J. Am. Chem. Soc. 2018, 140 (11), 3916–3928. [DOI] [PubMed] [Google Scholar]
  • (14).Lincoln KM; Gonzalez P; Richardson TE; Julovich DA; Saunders R; Simpkins JW; Green KN A potent antioxidant small molecule aimed at targeting metal-based oxidative stress in neurodegenerative disorders. Chem. Commun. 2013, 49 (26), 2712–2714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Wessel AJ; Schultz JW; Tang F; Duan H; Mirica LM Improved synthesis of symmetrically & asymmetrically N-substituted pyridinophane derivatives. Org. Biomol. Chem. 2017, 15 (46), 9923–9931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Drahos B; Kotek J; Hermann P; Lukes I; Toth E Mn(2+) complexes with pyridine-containing 15-membered Macrocycles: thermodynamic, kinetic, crystallographic, and (1)H/(17)O relaxation studies. Inorg. Chem. 2010, 49 (7), 3224–3238. [DOI] [PubMed] [Google Scholar]
  • (17).Garda Z; Molnar E; Kalman FK; Botar R; Nagy V; Baranyai Z; Brucher E; Kovacs Z; Toth I; Tircso G Effect of the Nature of Donor Atoms on the Thermodynamic, Kinetic and Relaxation Properties of Mn(II) Complexes Formed With Some Trisubstituted 12-Membered Macrocyclic Ligands. Front. Chem 2018, 6, 232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Green KN; Pota K; Tircso G; Gogolak RA; Kinsinger O; Davda C; Blain K; Brewer SM; Gonzalez P; Johnston HM; et al. Dialing in on pharmacological features for a therapeutic antioxidant small molecule. Dalton Trans. 2019, 48 (33), 12430–12439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Mekhail MA; Pota K; Kharel S; Freire DM; Green KN Pyridine modifications regulate the electronics and reactivity of Fe-pyridinophane complexes. Dalton Trans. 2023, 52 (4), 892–901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Brown PL; Ekberg C Hydrolisis of Metal Ions; Wiley-VCH, 2016. [Google Scholar]
  • (21).Kim B; Lingappa UF; Magyar J; Monteverde D; Valentine JS; Cho J; Fischer W Challenges of Measuring Soluble Mn(III) Species in Natural Samples. Molecules 2022, 27 (5), 1661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Pota K; Johnston HM; Madarasi E; Tircso G; Green KN Synthesis and characterization of two piperazine containing Macrocycles and their transition metal complexes. J. Inorg. Biochem. 2023, 241, No. 112124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Li F; Delgado R; Costa J; Drew MG; Felix V Ditopic hexaazamacrocycles containing pyridine: synthesis, protonation and complexation studies. Dalton Trans. 2005, 1, 82–91. [DOI] [PubMed] [Google Scholar]
  • (24).Aime S; Botta M; Geninatti Crich S; Giovenzana GB; Jommi G; Pagliarin R; Sisti M Synthesis and NMR Studies of Three Pyridine-Containing Triaza Macrocyclic Triacetate Ligands and Their Complexes with Lanthanide Ions. Inorg. Chem. 1997, 36 (14), 2992–3000. [DOI] [PubMed] [Google Scholar]
  • (25).Létumier F; Broeker G; Barbe J-M; Guilard R; Lucas D; Dahaoui-Gindrey V; Lecomte C; Thouin L; Amatore C Dichloro(1,4,8,11-tetraazacyclotetradecane)manganese(III) chloride: cis–trans isomerisation evidenced by infrared and electrochemical studies. J. Chem. Soc., Dalton Trans. 1998, 13, 2233–2240. [Google Scholar]
  • (26).Tomczyk D; Andrijewski G; Nowak L; Urbaniak P; Sroczynski ´D. Spectroscopic and electrochemical properties of mononuclear Mn(III) complex and of binuclear di-μ-oxo bridged Mn(III) and Mn(IV) complex with isocyclam. Inorg. Chim. Acta 2012, 390, 70–78. [Google Scholar]
  • (27).Daugherty PA; Glerup J; Goodson PA; Hodgson DJ; Michelsen K; Kleinpeter E Structural and Magnetic Characterization of Manganese(III) Complexes of 1,4,8,11-Tetraazacyclotetradecane (Cyclam). Acta Chem. Scand. 1991, 45 (3), 244–253. [Google Scholar]
  • (28).Suzuki M; Tokura S; Suhara M; Uehara A Dinuclear Manganese(III,IV) and Manganese(IV,IV) Complexes with Tris(2-pyridylmethyl)amine. Chem. Lett. 1988, 17 (3), 477–480. [Google Scholar]
  • (29).Johnston HM; Freire DM; Mantsorov C; Jamison N; Green KN Manganese (III/IV) μ-Oxo Dimers and Manganese (III) Monomers with Tetraaza Macrocyclic Ligands and Historically Relevant Open-Chain Ligands. Eur. J. Inorg. Chem. 2022, 2022 (19), No. e202200039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Hitchcock PB; Lappert MF; Taylor RG Synthesis, chemical behaviour, and structure (crystal and solution) of a fluorouranocene(IV) tetrafluoroborate; X-ray crystal structure of [{Ucp″2(μ-BF4)(μ-F)}2][cp″=η-C5H3(SiMe3)2]. J. Chem. Soc., Chem. Commun. 1984, 16, 1082–1084. [Google Scholar]
  • (31).Reedijk J Formation of Fluoride-Containing Coordination Compounds by Decomposition of Transition-Metal Tetrafluoroborates. Comments Inorg. Chem. 1982, 1 (6), 379–389. [Google Scholar]
  • (32).Cipot J; Wechsler D; McDonald R; Ferguson MJ; Stradiotto M Synthesis and Crystallographic Characterization of New Manganese(I) Complexes of Donor-Functionalized Indenes. Organometallics 2005, 24 (7), 1737–1746. [Google Scholar]
  • (33).Mekhail MA; Pota K; Schwartz TM; Green KN Functionalized pyridine in pyclen-based iron(iii) complexes: evaluation of fundamental properties. RSC Adv. 2020, 10 (52), 31165–31170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Schwartz TM; Burnett ME; Green KN Electronic influence of substitution on the pyridine ring within NNN pincer-type molecules. Dalton Trans. 2020, 49 (7), 2356–2363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Mekhail MA; Smith KJ; Freire DM; Pota K; Nguyen N; Burnett ME; Green KN Increased Efficiency of a Functional SOD Mimic Achieved with Pyridine Modification on a Pyclen-Based Copper(II) Complex. Inorg. Chem. 2023, 62 (14), 5415–5425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (36).Lee W-T; Xu S; Dickie DA; Smith JM A Robust Mn Catalyst for H2O2 Disproportionation in Aqueous Solution. Eur. J. Inorg. Chem. 2013, 2013 (22–23), 3867–3873. [Google Scholar]
  • (37).Goodson PA; Hodgson DJ; Glerup J; Michelsen K; Weihe H Syntheses and characterization of binuclear manganese-(III,IV) and (IV,IV) complexes with 1,4,7,10-tetraazacyclododecane (cyclen). Inorg. Chim. Acta 1992, 197 (2), 141–147. [Google Scholar]
  • (38).Brewer KJ; Liegeois A; Otvos JW; Calvin M; Spreer LO Synthesis and properties of two bimetallic mixed-valence di-μ-oxo manganese complexes with different tetra-aza Macrocyclic ligands. J. Chem. Soc., Chem. Commun. 1988, 17, 1219–1220. [Google Scholar]
  • (39).Kozuch S; Martin JML Turning Over” Definitions in Catalytic Cycles. ACS Catal. 2012, 2 (12), 2787–2794. [Google Scholar]
  • (40).Ledesma GN; Eury H; Anxolabéhère-Mallart E; Hureau C; Signorella SR A new mononuclear manganese(III) complex of an unsymmetrical hexadentate N3O3 ligand exhibiting superoxide dismutase and catalase-like activity: synthesis, characterization, properties and kinetics studies. J. Inorg. Biochem. 2015, 146, 69–76. [DOI] [PubMed] [Google Scholar]
  • (41).Hage R Oxidation catalysis by biomimetic manganese complexes. Recl. Trav. Chim. Pays-Bas 1996, 115 (9), 385–395. [Google Scholar]
  • (42).Chino M; Leone L; Zambrano G; Pirro F; D’Alonzo D; Firpo V; Aref D; Lista L; Maglio O; Nastri F; et al. Oxidation catalysis by iron and manganese porphyrins within enzyme-like cages. Biopolymers 2018, 109 (10), No. e23107. [DOI] [PubMed] [Google Scholar]
  • (43).de Boer JW; Browne WR; Feringa BL; Hage R Carboxylate-bridged dinuclear manganese systems – From catalases to oxidation catalysis. C. R. Chim. 2007, 10 (4–5), 341–354. [Google Scholar]
  • (44).Signorella S; Hureau C Bioinspired functional mimics of the manganese catalases. Coord. Chem. Rev. 2012, 256 (11–12), 1229–1245. [Google Scholar]
  • (45).Pellegatti L; Zhang J; Drahos B; Villette S; Suzenet F; Guillaumet G; Petoud S; Toth E Pyridine-based lanthanide complexes: towards bimodal agents operating as near infrared luminescent and MRI reporters. Chem. Commun. 2008, 48, 6591–6593. [DOI] [PubMed] [Google Scholar]
  • (46).Vuillamy A; Zebret S; Besnard C; Placide V; Petoud S; Hamacek J Functionalized Triptycene-Derived Tripodal Ligands: Privileged Formation of Tetranuclear Cage Assemblies with Larger Ln(III). Inorg. Chem. 2017, 56 (5), 2742–2749. [DOI] [PubMed] [Google Scholar]
  • (47).Bell HJ; Malins LR Peptide macrocyclisation via late-stage reductive amination. Org. Biomol. Chem. 2022, 20 (31), 6250–6256. [DOI] [PubMed] [Google Scholar]
  • (48).Hamada T; Manabe K; Ishikawa S; Nagayama S; Shiro M; Kobayashi S Catalytic asymmetric aldol reactions in aqueous media using chiral bis-pyridino-18-crown-6-rare earth metal triflate complexes. J. Am. Chem. Soc. 2003, 125 (10), 2989–2996. [DOI] [PubMed] [Google Scholar]
  • (49).Pradhan RN; Chakraborty S; Bharti P; Kumar J; Ghosh A; Singh AK Seven coordinate Co(ii) and six coordinate Ni(ii) complexes of an aromatic Macrocyclic triamide ligand as paraCEST agents for MRI. Dalton Trans. 2019, 48 (24), 8899–8910. [DOI] [PubMed] [Google Scholar]
  • (50).Serrano-Plana J; Aguinaco A; Belda R; Garcia-Espana E; Basallote MG; Company A; Costas M Exceedingly Fast Oxygen Atom Transfer to Olefins via a Catalytically Competent Nonheme Iron Species. Angew. Chem., Int. Ed. 2016, 55 (21), 6310–6314. [DOI] [PubMed] [Google Scholar]
  • (51).Bridonneau N; Chamoreau LM; Gontard G; Cantin JL; von Bardeleben J; Marvaud V A high-nuclearity metal-cyanide cluster [Mo6Cu14] with photomagnetic properties. Dalton Trans. 2016, 45 (23), 9412–9418. [DOI] [PubMed] [Google Scholar]
  • (52).Wilson JM; Giordani F; Farrugia LJ; Barrett MP; Robins DJ; Sutherland A Synthesis, characterisation and anti-protozoal activity of carbamate-derived polyazamacrocycles. Org. Biomol. Chem. 2007, 5 (22), 3651–3656. [DOI] [PubMed] [Google Scholar]
  • (53).Manov GG; DeLollis NJ; Acree SF Comparative Liquid-Junction Potentials of Some pH Buffer Standards and the Calibration of pH Meters. J. Res. Natl. Bur. Stand. 1945, 34, 115–127. [Google Scholar]
  • (54).Manov GG; DeLollis SF; DeLollis SF; Lindvall NJ; Acree PW Effect of Sodium Chloride on the Apparent Ionization Constant of Boric Acid and the pH Values of Borate Solutions. J. Res. Natl. Bur. Stand. 1946, 36, 543–558. [Google Scholar]
  • (55).Irving HM; Miles MG; Pettit LD A study of some problems in determining the stoicheiometric proton dissociation constants of complexes by potentiometric titrations using a glass electrode. Anal. Chim. Acta 1967, 38, 475–488. [Google Scholar]
  • (56).Zekany L; Nagypal I PSEQUAD. In Computational Methods for the Determination of Formation Constants, Leggett D, Ed.; Plenum Press, 1985; pp 291–353. [Google Scholar]
  • (57).APEX3 Version 1; BRUKER AXS Inc.: Madison, WI, 2016. [Google Scholar]
  • (58).SAINT; BRUKER AXS Inc.: Madison, WI, 2012. [Google Scholar]
  • (59).SADABS; BRUKER AXS Inc.: Madison, WI, 2001. [Google Scholar]
  • (60).Dolomanov OV; Bourhis LJ; Gildea RJ; Howard JAK; Puschmann H OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42 (2), 339–341. [Google Scholar]
  • (61).Sheldrick GM A short history of SHELX. Acta Crystallogr. A 2008, 64 (1), 112–122. [DOI] [PubMed] [Google Scholar]
  • (62).Sheldrick GM Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71 (1), 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (63).Laidler KJ The development of the Arrhenius equation. J. Chem. Educ. 1984, 61 (6), 494–498. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES