Abstract
The tetradentate mixed imino/amino phenoxide ligand (N-(3,5-di-tert-butylsalicylidene)-N’-(2-hydroxyl-3,5-di-tert-butylbenzyl))-trans-1,2-cyclohexanediamine (salalen) was complexed with CuII, and the resulting Cu complex (2) was characterized by a number of experimental techniques and theoretical calculations. Two quasi-reversible redox processes for 2, as observed by cyclic voltammetry, demonstrated the potential stability of oxidized forms, and also the increased electron-donating ability of the salalen ligand in comparison to the salen analogue. The electronic structure of the one-electron oxidized [2]+ was then studied in detail, and Cu K-edge X-ray Absorption Spectroscopy (XAS) measurements confirmed a CuII-phenoxyl radical complex in solution. Subsequent resonance Raman (rR) and variable temperature 1H NMR studies, coupled with theoretical calculations, showed that [2•]+ is a triplet (S = 1) CuII-phenoxyl radical species, with localization of the radical on the more electron-rich aminophenoxide. Attempted isolation of X-ray quality crystals of [2•]+ afforded [2H]+, with a protonated phenol bonded to CuII, and an additional H-bonding interaction with the SbF6− counterion. Stoichiometric reaction of dilute solutions of [2•]+ with benzyl alcohol showed that the complex reacted in a similar manner as the oxidized CuII-salen analogue, and does not exhibit a substrate-binding pre-equilibrium as observed for the oxidized bisaminophenoxide CuII-salan derivative.
TOC Image

1. Introduction
The incorporation of redox-active ligands into transition metal complexes has garnered significant interest,[1–6] drawing inspiration from enzymatic systems. Pro-radical ligands are essential components in the active sites of many metalloenzymes including cytochrome c oxidase[7] and P450,[8] glyoxal oxidase,[9] and galactose oxidase (GOase).[10] The Cu-containing GOase is the archetypical example, catalyzing the oxidation of primary alcohols to aldehydes with the concomitant reduction of O2 to H2O2.[11, 12] The catalytically active form of GOase incorporates a CuII metal center and a tyrosyl ligand radical, which is cross-linked with a cysteine residue, with each undergoing one-electron reduction during substrate oxidation.[10] The simplicity and reactivity of the active site in GOase has inspired many structural and functional small-molecule models.[13–16]
We, and others, have had a long-standing interest in the electronic structure of oxidized metal salen (salen is a common abbreviation for N2O2 bis-imine bis-phenoxide ligands) complexes,[17–44] in particular the similarity of the coordination sphere to GOase.[15] The highly modular synthesis of salen ligands allows for facile tuning of both steric and electronic properties, and ability to stabilize a wide variety of transition and main group metal complexes in a number of oxidation states.[15, 45–50] Traditionally, the term salen refers to ligands specifically prepared via condensation of two equivalents of salicylaldehyde and ethylenediamine, but now includes ligands with different phenyl ring substituents and diamine backbone.[51] The imine functionalities are reducible to afford the corresponding tetrahydrosalen (salan) analogues,[35, 52, 53] while a half-reduced salen ligand (salalen) can be prepared via reduction of only single imine function.[54–61] The development of synthetic methodologies for asymmetric salens (ligands with different substitution patterns on each of the phenol moieties) allows for the preparation of a diverse library of ligands.[62]
Upon one-electron oxidation, metal-salen complexes can exhibit redox activity at either the ligand or the metal centre, depending on the transition metal used, the peripheral ligand substituents, the solvent or the temperature. Without the use of ortho- and para-phenolate protecting groups, oxidized metal salen complexes are prone to rapid polymerization via radical coupling even at −80°C.[63–67] In the context of this work, we aimed to study the comparative electronic structure and associated reactivity of the one-electron oxidized forms of Cu salen (1), Cu salalen (2), and Cu salan (3), in which the imine functionalities are reduced sequentially (Figure 1).
Figure 1.

Chemical structures of Cu salen (1), Cu salalen (2), and Cu salan (3) of relevance to this work.
We reported previously the one-electron oxidation behavior and reactivity of [1]+, in comparison to [3]+.[24, 52] Initial studies showed that both oxidized complexes react with the model substrate benzyl alcohol to form benzaldehyde, however compound [3]+ reacts more quickly despite being a weaker oxidant (E1/2 of 80 and 450 mV for 3 and 1, respectively, vs. Fc+/Fc). A mechanistic difference was uncovered in which initial substrate binding to [3]+ presumably results in more rapid reactivity. Subsequent in-depth characterization of the electronic structure of [1]+ demonstrated that the complex exists as two essentially isoenergetic valence tautomers at 298 K, a triplet CuII-phenoxyl radical species, and a diamagnetic CuIII species. Thus, the slower reactivity for [1]+, in comparison to [3]+ is presumably due to the presence of the CuIII tautomer, which disfavours axial substrate binding in comparison to the CuII d9 species. Substrate predisposition via substrate binding allows enzymes such as GOase to perform highly selective transformations under mild conditions. We were thus interested to investigate the electronic structure and reactivity of the one-electron oxidized asymmetric Cu salalen complex [2]+, in comparison to [1]+ and [3]+. Metal complexes of the salalen ligand detailed herein have been investigated in catalysis, including the copolymerization of cyclohexene oxide with CO2 (Cr),[55] hydrosilylation (Mo),[60] and asymmetric nitro-aldol chemistry (Cu).[59] In this work we detail the electronic structure of [2]+, and probe the reactivity of this interesting localized CuII-phenoxyl radical species.
2. Experimental
2.1 General Considerations
All solvents and reagents were obtained from commercial sources and used as received unless noted otherwise. The salalen ligand was prepared as reported previously.[54] Data collection and structure solution for the crystal structures of 2 and [2H]+ SbF6− were conducted at the X-ray Crystallographic Laboratory, S146 Kolthoff Hall, Department of Chemistry, University of Minnesota. Details are available in the CIF files which have CCDC numbers 1563639 for 2 and 1563639 for [2H]+ SbF6−. Cyclic voltammetry was performed using a BAS CV-40 potentiometer, a Ag wire reference electrode, a platinum disk working electrode, and a Pt wire counter electrode with 0.1M nBu4NClO4 solutions in CH2Cl2. Ferrocene was used as an internal standard. X-band EPR spectra were collected with a Bruker EMX spectrometer using an ER041XG microwave bridge and an ER4102ST cavity. UV-Vis-NIR spectroscopy was performed using a Cary 500 dual-beam spectrophotometer. Variable temperature UV-Vis spectroscopy was performed using a Cary 50 spectrophotometer with a custom-designed immersable fiber-optic quartz probe with variable path length (1 and 10 mm; Hellma, Inc.). Constant temperatures were maintained either by a cooling bath (Kinetic Systems, New York) or a dry ice/acetone bath. Solution temperatures were directly monitored by insertion of an Omega temperature probe into the solutions. Variable temperature Evans method 1H NMR measurements were performed on a Varian Inova 300 MHz instrument in CD2Cl2, and diamagnetic corrections were estimated using Pascal’s constants. Solvent contraction was accounted for in all variable temperature studies. Analytical services were provided by Desert Analytics (Tucson, AZ). Gas chromatography was performed on an HP5890 gas chromatograph equipped with a 30 m DB-1 column (J&WScientific, Inc.) and a flame ionization detector.
2.2 Synthesis
2.2.1 (N-(3,5-di-tert-butylsalicylidene)-N’-(2-hydroxide-3,5-di-tert-butylbenzyl))-trans-1,2-cyclohexanediamine Copper(II) (2)
(0.823 g, 1.5 mmol) of salalen ligand[54] was dissolved in CH2Cl2 (5 mL), and CuCl2·2H2O (0.256 g, 1.5 mmol) was added as a solution in EtOH (6 mL). With stirring, NEt3 (0.46 mL, 6.0 mmol) was added. The volume was reduced by half after 1 h and H2O (2 mL) was added to precipitate a green solid. The solid was isolated and washed with EtOH and H2O to afford 2 (0.809 g, 88 %). X-ray quality crystals were obtained by slow evaporation of a concentrated CH2Cl2/MeOH solution. HRMS-ESI+ (m/z): [M]+ calculated for (C36H54N2O2Cu), 609.3581; Found, 609.4041. UV-vis-NIR (λmax [cm−1] (ε [M−1 cm−1]): 25,300 (6,250), 17,000 cm−1 (800). Anal. Calcd (found) for C36H54N2O2Cu· H2O: C, 68.81 (69.17); H, 8.98 (8.54); N, 4.46 (4.56).
2.2.2 (N-(3,5-di-tert-butylsalicylidene)-N’-(2-hydroxyl-3,5-di-tert-butylbenzyl))-trans-1,2-cyclohexanediamine Copper(II) hexafluoroantimonate ([2H]+SbF6−
2 (0.222 g, 0.4 mmol) was dissolved in CH2Cl2 (5 mL) and solid AgSbF6 (0.125 g, 0.4 mmol) was added. A purple suspension formed immediately. After 1 h the mixture was filtered through celite and recrystallized by pentane diffusion into the bulk CH2Cl2 solution. UV-vis-NIR (λmax [cm−1] (ε [M−1 cm−1]): 25,000 (6,500), 19,000 (1,700), 17,600 (2,190), 8,600 (1,212). Brown/green crystals suitable for X-ray crystallography were isolated after one week. Anal. Calcd (found) for C36H55N2O2CuSbF6· 1.5 CH2Cl2: C, 46.99 (47.32); H, 6.10 (5.78); N, 2.92 (2.99).
2.3 Resonance Raman Spectroscopy
Resonance Raman spectra were obtained on a SpectraPro-300i spectrometer (Acton Research) with a 2400-groove grating, a Beamlok 2060 Kr ion laser (Spectra-Physics), a holographic supernoch filter (Kaiser Optical Systems), and LN-1100PB CCD detector (Princeton Instruments) cooled with liquid N2. Spectra were collected on solvated samples (1 mM) in spinning cells (2 cm diameter, 1500 rpm) at 213 K, at an excitation wavelength λex = 413.1 nm (20 mW), 90° scattering geometry, and 5 min data accumulation. Peak frequencies were calibrated relative to indene and CCl4 standards (accurate to ± 1 cm−1). During each Raman experiment, UV/Vis spectra were collected simultaneously on a PMA-11 CCD spectrophotometer (Hamamatsu) with a Photal MC-2530 (D2/W2) light source (Otsuka Electronic Co.).
2.4 X-ray Absorption Spectroscopy
Cu K-edge XAS data were collected at the Stanford Synchrotron Radiation Laboratory under dedicated ring conditions of 60–100 mA and 3 GeV. The data were recorded on the wiggler beam line 7-3. A Si (220) monochromator was used for energy selection and detuned (50%) to minimize higher harmonic components of the X-ray beam. The samples were formulated as frozen solutions in CH2Cl2. The sample cell was pre-cooled in liquid N2 before insertion into an Oxford Instruments CF1208 continuous flow liquid helium cryostat, in which the samples were maintained at 10–15 K during data collection. Data were measured in fluorescence mode with N2 filled ionization chambers before and after the sample. Internal energy calibration was performed by simultaneous measurement of the transmission signal through a Cu reference foil. The first inflection point of the copper reference data was aligned to 8980.3 eV. Data of four scans were averaged and were processed by fitting a smooth polynomial to the pre-edge region, which was background subtracted from the entire spectrum. A three region cubic spline was used to model the smooth background above the edge. Data were normalized by subtracting the spline and normalizing the postedge data to 1.0. EXAFS data were fit with the EXAFSPAK[68] program, using theoretical phase and amplitude parameters derived from FeFF 7.0.[69] Absorber-scatterer distances (R, Å) and Debye-Waller factors (σ2, Å2) were varied independently for each component in each fit. Additionally, the E0 (eV) value, or the energy onset of photoionization, was also varied for each fit, but kept to a common value for all components of a given fit, and coordination numbers were kept constant. The intensities and energies of the 1s→3d pre-edge transitions for each complex were quantified by simultaneously fitting pseudo-Voigt line shapes to the data and the second derivative of the data using the EDG_FIT program. The energy position, half-width, and amplitude of each feature were allowed to vary within each fit. A background pseudo-Voigt function was used to model the rising absorption edge shape. The approximate peak area was measured as the product of the peak amplitude and the full width at half-maximum of the pre-edge feature. The stated peak areas are the average of the areas over the four energy ranges used, and standard deviations were calculated to estimate the variability of the fits.
2.5 Calculations
Geometry optimizations were performed using the Gaussian 09 program[70] (revision D.01), the B3LYP[71, 72] functional, a polarized continuum model (PCM) for CH2Cl2 (dielectric ε = 8.94),[73–76] and the 6–31G(d) basis set on all atoms as this double-ζ optimization has afforded good agreement with solid-state structural data of related salen systems.[20, 24, 25, 54, 77] The optimized geometries were confirmed as stationary states using frequency calculations. Single point calculations for energetic analysis were performed using the B3LYP functional and the TZVP basis set of Ahlrichs[78, 79] on all atoms with a PCM for CH2Cl2.
3. Discussion
3.1 Synthesis and Characterization of 2
The CuII salalen complex 2 (Figure 1) was synthesized in high yield by reacting the salalen ligand[54] with CuCl2•2H2O and NEt3 under aerobic conditions. The X-ray structure of 2 (Figure 2) with selected crystallographic data (Table 1) exhibits a slightly distorted square-planar geometry presumably due to the sterically demanding ortho-tBu substituents, as well as the increased flexibility due to the partially-reduced salen ligand scaffold. This ligand asymmetry also results in a non-symmetric coordination sphere (Table 2): a slightly longer Cu-N bond is observed for the Cu-Naminophenoxide (Cu-N(H) = 1.969 Å) in comparison to Cu-Niminophenoxide (Cu-N(=) = 1.937 Å). This pattern is reversed for the Cu-O bond, where the Cu-Oaminophenoxide bond (1.874 Å) is shorter than Cu-Oiminophenoxide (1.893 Å). The distortion and non-symmetric coordination sphere observed for 2 is also observed for the reported Ni analogue.[54]
Figure 2.

ORTEP plot of 2 (50% probability) using POV-Ray, including selected hydrogen atoms. Selected average interatomic distances [Ǻ] and angles [°]: Cu(1)-N(1): 1.969, Cu(1)-N(2): 1.937, Cu(1)-O(1): 1.874, Cu(1)-O(2): 1.893, C(1)-O(1): 1.329, C(20)-O(2): 1.305; Angles: N(1)-Cu(1)-N(2): 84.9, N(2)-Cu(1)-O(2): 92.9, O(1)-Cu(1)-O(2): 89.5, N(1)-Cu(1)-O(1): 92.9, N(1)-Cu(1)-O(2): 176.5, O(1)-Cu(1)-N(2): 175.7.
Table 1.
Selected Crystallographic Data for 2 and [2H]+SbF6−
| 2 | [2H]+SbF6− | |
|---|---|---|
| Formula | C36H54N2O2Cu | C36H55N2O2CuSbF6 |
| Formula weight | 610.36 | 847.14 |
| space group | P 1 | P 1 |
| a (Å) | 10.2792(9) | 9.412(3) |
| b (Å) | 13.6782(12) | 16.187(5) |
| c (Å) | 14.8362(19) | 17.244(5) |
| α (°) | 117.382(2) | 111.733(4) |
| β (°) | 97.513(2) | 106.803(4) |
| γ (°) | 103.518(2) | 90.066(4) |
| V [Å3] | 1731.0(3) | 2318.8(11) |
| Z, Dcalc [g/cm3] | 2 | 1 |
| T (K) | 173 | 173 |
| ρcalcd (g cm−3) | 1.171 | 1.517 |
| λ (Å) | 0.71073 | 0.71073 |
| μ (mm−1) | 0.662 | 1.387 |
| R indicesa with I>2σ(I) (data) | 6793 | 16675 |
| wR2 | 0.1110 | 0.1557 |
| R1 | 0.0435 | 0.0575 |
| Goodness-of-fits on F2 | 1.011 | 1.071 |
Goodness-of-fit on F.
Table 2.
Coordination sphere metrical data for 2, [2]+, and [2H]+.
| Compound | Cu-O (Å) | Cu-O(=)a (Å) | Cu-N(=)a (Å) | Cu-N(H) (Å) | |
|---|---|---|---|---|---|
| 2 | X-rayb | 1.874 | 1.893 | 1.937 | 1.969 |
| XAS | 1.92(1)c | ||||
| Calculatedd | 1.890 | 1.908 | 1.933 | 2.016 | |
|
| |||||
| [2]+ | XAS | 1.92(1)c | |||
| Calculated (S = 1)d | 1.955 | 1.868 | 1.909 | 2.008 | |
|
| |||||
| [2H]+ | X-ray | 2.022 | 1.846 | 1.910 | 1.947 |
| Calculatedd | 2.024 | 1.856 | 1.896 | 1.891 | |
(=) signifies imine portion of ligand.
Average of two values from two molecules in the unit cell.
Average Cu-O/Cu-N bond lengths from EXAFS fit.
See Experimental Section for calculation details.
3.2 Electrochemistry
Redox processes for 2 were probed by cyclic voltammetry (CV) in CH2Cl2 using tetra-n-butyl ammonium perchlorate (nBu4NClO4) as the supporting electrolyte. Complex 2 exhibits two quasi-reversible, one-electron redox processes at 0.15 V and 0.54 V vs. Fc+/Fc (Table 3) that are in line with the previously reported Ni derivative.[54] In general, the redox potentials for 2 are intermediate in comparison to the salen analogue 1, and the fully reduced salan analogue 3 (Figure S1)[24, 52] in accord with the intermediate electron-donating ability of the salalen ligand in comparison to the salen and salan analogues. We have previously reported that [1]+ affords a CuIII complex in the solid state, and that this electronic isomer is in equilibrium with a CuII-phenoxyl radical species in solution. This contrasts the temperature-invariant CuII-phenoxyl radical electronic structure for [3]+.[24] Thus, we were interested to isolate the one-electron oxidized [2]+, determine the electronic structure, and how the electronic structure would influence subsequent reactivity.
Table 3.
Redox Potentials of CuII complexes 1–3 Versus Fc+/Fc at 298 K.a
| Compound | E1/21 (V) | E1/22 (V) | Reference |
|---|---|---|---|
| 1 | 0.45 (0.13) | 0.65 (0.12) | [52] |
| 2 | 0.15 (0.14) | 0.54 (0.14) | This work |
| 3 | 0.08 (0.14) | 0.20 (0.13) | [52] |
Peak-to-peak differences in brackets (|Epa – Epc| in V).
3.3 Chemical oxidation of 2 to [2]+
The chemical oxidation of 2 to afford [2]+ was probed under a number of different conditions. Employing the thianthrene radical oxidant (Th+•SbF6−; 0.86 V vs. Fc+/Fc)[80] in CH2Cl2 solution cleanly afforded [2]+SbF6− as shown by UV-vis-NIR spectroscopy (Figure 3). The electronic absorption spectrum of 2 is typical of a d9, square-planar Cu complex, with a relatively intense charge transfer transition at 25,300 cm−1 (ε = 6,250 M−1 cm−1) and a weaker d-d transition at 17,000 cm−1 (ε = 800 M−1 cm−1, Figure 3).[24, 38, 77] Chemical oxidation by addition of Th+•SbF6− leads to the appearance of three new bands; a shoulder at ca. 19,000 cm−1 (ε = 1,700 M−1 cm−1), and bands at 17,600 cm−1 (ε = 2,190 M−1 cm−1) and 8,600 cm−1 (ε = 1,212 M−1 cm−1). The broad shape and low intensity of the NIR band for [2]+ is suggestive of a localized ligand radical species,[21, 39, 54, 81, 82] however further confirmation of the electronic structure is detailed herein. Subsequent stability studies at 0.1 mM at 298 K revealed very little decomposition over a 15 h period (Figure 3 inset), however at higher concentrations [2]+ was observed to decompose in solution over time (vide infra).
Figure 3.

Electronic spectra of 0.1 mM 2 (black) and [2]+SbF6− (red) in CH2Cl2 at 298 K. [2]+SbF6− was oxidized with Th+•SbF6−. Inset: Decay of 0.1 mM [2]+SbF6− in CH2Cl2 at 298 K over 15 h.
3.4 Continuous Wave X-band Electron Paramagnetic Resonance (EPR)
The X-band EPR spectrum of a frozen solution of 2 collected at 77 K exhibits features consistent with a square-planar d9 CuII center (Figure 4), as observed for both 1 and 3.[52] Oxidation of 2 to [2]+ results in the almost complete loss of an EPR signal, with < 4% of the original intensity by spin quantitation (Figure 4). Three possible electronic structures can account for the loss of EPR signal upon oxidation: 1) an antiferromagnetically-coupled CuII-phenoxyl radical species (S = 0); 2) a ferromagnetically-coupled CuII-phenoxyl radical species (S = 1) with a triplet energy difference outside of the range of X-band EPR; and 3) a CuIII bis-phenoxide (S = 0). We employed additional experimental and theoretical analysis to further probe the electronic structure of [2]+.
Figure 4.

X-band EPR spectra for concentration-matched 2 mM samples of 2 (black) and simulation (red), in 2-MeTHF at 77 K, and [2]+ (blue), in CH2Cl2 at 77 K. Simulation parameters: g‖ = 2.217, g⊥ = 2.046, ACu‖ = 193, ACu⊥ = 31, A values in cm−1. Conditions: frequency = 9.3653 GHz; power = 2mW; modulation amplitude = 20 G.
3.5 Cu K-Edge X-ray Absorption Spectroscopy (XAS)
Cu K-edge XAS[83–85] was used to further probe the metal oxidation state and structure of 2 and [2]+ in frozen CH2Cl2 solution. The Cu K-edge data for 2 and [2]+ (Figure S2) are similar, with only a minor change in the 1s → 3d transition, or pre-edge, at 8979.5 eV for 2, and 8979.7 eV for [2]+, indicating that the CuII oxidation state is maintained in [2]+ upon one-electron oxidation. This is in contrast to the increase in the pre-edge of ca. 1 eV upon oxidation of 1 to [1]+, which is indicative of metal-based oxidation (CuII to CuIII).[24] Extended X-ray absorption fine structure (EXAFS) for 2 and [2]+ are well-fit to four N/O donors of the ligand, with a single shell distance of 1.92 Å for both complexes (Figure S3). The same average value for the N/O donors of 2 and [2]+, presumably reflects the overall asymmetry in the structure, and the fact that EXAFS analysis typically cannot distinguish between first-shell atoms that differ in Z by 1 (e.g. O and N).[86] The average value of 1.92 Å well-matches the average coordination sphere metrical data obtained by other methods (Table 2). Two additional parameters were required to fit the outer shell contributions to the EXAFS data: a shell of single scattering carbon atoms at 2.86 Å and 2.88 Å for 2 and [2]+, respectively, and multiple scattering from these same carbon atoms at 3.12 Å and 3.13 Å for 2 and [2]+, respectively (Table S1). Overall, the Cu K-edge XAS data supports a CuII-phenoxyl radical electronic description for [2]+.
3.6 Resonance Raman (rR) Spectroscopy
Resonance Raman (rR) spectroscopy is an important tool for the characterization phenoxyl radical species.[87] The rR spectra of 2 and [2•]+ (Figure 5) exhibit key differences upon oxidation, with vibrational modes associated with both phenoxyl π → π* and phenolate → CuII LMCT transitions observable in the spectrum of [2•]+. The features at 1495 and 1597 cm−1 are assigned to characteristic phenoxyl radical C-O stretching (ν7a), and Cortho-Cmeta stretching (ν8a) modes respectively.[88–90] The phenolate bands at 1530 and 1627 cm−1, are maintained upon oxidation to [2•]+, consistent with localization of the ligand radical on the Raman timescale.[23, 91] Overall, the rR data supports the presence of both phenolate and phenoxyl ligands based on the observed vibrational modes, and thus localization of the ligand radical in [2•]+.
Figure 5.

Resonance Raman (rR) spectra of (a) 2 and (b) [2•]+ at 213 K in CH2Cl2 (λex = 413 nm).
3.7 Variable-Temperature (VT) 1H NMR and UV-Vis Spectroscopy
With confirmation of a localized CuII-phenoxyl radical structure for [2•]+, we further probed the electronic structure, and any possible temperature-dependence, using variable-temperature (VT) 1H NMR and UV-Vis spectroscopies. Previous work has shown that [1]+ exists in a reversible spin-equilibrium between a CuII-phenoxyl radical species (S = 1) and a high-valent CuIII form (S = 0).[24] The 1H NMR spectrum of [2]+ in CD2Cl2 at 293 K exhibits broad signals over a wide spectral range (Figure S4). Subsequent solution susceptibility measurements by the Evans method[92] indicates a μeff of 2.67 BM (ca. 2 unpaired electrons at 293 K) consistent with a ferromagnetically-coupled CuII-phenoxyl radical species in solution. Cooling this solution to 193 K results in small decrease of the μeff to 2.36 (Figure 6), and this change is reversible upon warming to 293 K. We also probed the temperature-dependence of the UV-vis spectrum of [2•]+, and observe a small increase in the band at 17,500 cm−1 at low temperature (Figure 6). While this data is suggestive of a possible spin-equilibrium in solution, the data could not be reliably modelled over the available temperature range. The changes for [2•]+ are much smaller in comparison to the temperature-dependent susceptibility and UV-vis data reported for [1]+ in solution.[24] In addition, DFT calculations predict the CuIII electronic isomer for [2]+ to be much higher in energy (vide infra).
Figure 6.

Comparison of the VT solution susceptibility by 1H NMR (black circles, CD2Cl2) and 17,500 cm−1 band intensity (red squares, CH2Cl) for [2•]+SbF6−. See experimental section for details.
3.8 Theoretical Calculations on 2 and [2]+
Using the B3LYP[71, 72] functional with a polarized continuum model (PCM)[73–76] for CH2Cl2, a non-symmetric structure is predicted for the doublet 2 with coordination-sphere metrical parameters within ± 0.05 Å of the X-ray data (Table 2). Using the same DFT method, three different electronic configurations for [2]+ were considered: 1) a CuIII complex (S = 0), 2) a CuII complex antiferromagnetically coupled to a phenoxyl radical (broken symmetry, S = 0), or 3) a CuII complex ferromagnetically coupled to a phenoxyl radical (triplet, S = 1). Consistent with the experimental XAS and rR data, the CuII-phenoxyl radical description is favored over the CuIII solution by ca. 16 kcal mol−1. The two CuII-phenoxyl radical electronic structures are nearly isoenergetic with the triplet favored by 0.3 kcal mol−1. The significant magnetic susceptibility of [2]+ is consistent with a predominance of the triplet species in solution. The average coordination distance of the optimized S = 1 solution is within ± 0.02 Å of the experimental EXAFS data (Table 2). The phenoxyl radical is predicted to localize on the more electron-rich aminophenoxide moiety, with the associated magnetic orbitals shown in Figure 7. The phenoxyl radical (π) and Cu(II) dx2−y2 magnetic orbitals are orthogonal in accordance with the triplet spin state.[24, 93] This is consistent with the observation of both phenolate and phenoxyl bands in the rR spectrum of [2•]+. Overall, both experimental and theoretical results support that [2•]+ is a localized CuII-phenoxyl radical species.
Figure 7.

Magnetic orbitals for the triplet (S = 1) electronic solution for [2•]+.
3.9 Reactivity Studies
3.9.1 Decomposition of [2•]+ in solution
In efforts to isolate X-ray quality crystals of [2•]+, bulk oxidation of 2 at ca. 75 mM, using AgSbF6 (0.65 V vs. Fc+/Fc) in CH2Cl2 at 298 K under an inert atmosphere. The resulting purple suspension was filtered and set aside to crystallize, and after one week green/brown crystals suitable for X-ray analysis were isolated. The X-ray structure is interpreted as a protonated phenoxide complex ([2H]+ SbF6−) rather than a CuII-phenoxyl radical species ([2•]+SbF6−) (Figure 8); the Cu-O and Cu-N bond lengths on the iminophenoxide side in [2H]+ contract considerably in comparison to 2, while the protonated phenoxide (Cu-O(H)) lengthens (Figure 8). The DFT-optimized structure of [2H]+ has a Cu-O(H) bond length that matches within ±0.002 Å of the X-ray metrical data. In addition, a H-bonding interaction between the protonated phenoxide and the SbF6− counterion is evident (O1—F3 = 2.727 Å). A frozen solution EPR spectrum of crystalline material exhibits typical features consistent with a square-planar d9 CuII center (Figure S5). Interestingly, a similar protonated species was recently crystallographically-characterized from a solution of a Cu(III) anilido complex.[94]
Figure 8.

ORTEP plot of [2H]+SbF6− (50% probability) using POV-Ray, including selected hydrogen atoms. Selected average interatomic distances [Ǻ] and angles [°]: Cu(1)-N(1): 1.947, Cu(1)-N(2): 1.910, Cu(1)-O(1): 2.022, Cu(1)-O(2): 1.846, C(1)-O(1): 1.379, C(20)-O(2): 1.334; Angles: N(1)-Cu(1)-N(2): 85.7, N(2)-Cu(1)-O(2): 95.7, O(1)-Cu(1)-O(2): 93.0, N(1)-Cu(1)-O(1): 87.9, N(1)-Cu(1)-O(2): 170.0, O(1)-Cu(1)-N(2): 164.5.
Intrigued by this result, a bulk oxidation performed under an inert atmosphere (ca. 50 mM) gave after one week a green/brown material. Subsequent purification using silica gel chromatography afforded a 2:1 ratio of 2 and Cu salen 1 in a total yield of 75 % based on the initial amount of 2. We speculate that the recovered 2 was actually [2H]+, however, deprotonation readily occurs during purification (vide infra). Evolution of the CuII EPR signal for the decomposition of [2•]+ over a 30 h period at a concentration of 3.3 mM (Figure S6) provides additional insights into this transformation. The initial spectrum displays <2 % of the intensity in comparison to a concentration-matched sample of 2, consistent with initial formation of [2•]+. Over 30 h, a CuII signal appears, and reaches a maximum intensity of ca. 80 % relative to a concentration-matched sample of 2. The frozen solution EPR spectrum of the isolated bulk oxidized material clearly shows a broadening of the parallel features, consistent with multiple CuII species in the sample (Figure S5). This EPR pattern can be fit to a 2:1 ratio of [2H]+ (Figure S7), synthesized directly from equimolar amounts of triflic acid and 2, and 1, providing further confirmation for the stoichiometry of the decay process via purification of bulk material. While the decomposition of [2•]+ was not probed further, we note that in order for the 2:1 ratio of [2H]+ and 1 to form, an additional reducing equivalent is needed for electron balance. While the reducing agent was not identified, the ca. 75 % total yield from the bulk reaction points towards other Cu-ligated species that were not isolated.
3.9.2 Reactivity with Benzyl Alcohol
The stoichiometric reaction of [2•]+ with the standard substrate benzyl alcohol was compared to the previously reported data for [1]+ and [3]+.[52] While the first-order rate constant for [1]+ varies linearly over the range [alcohol] = 0.05 – 1.0 M, providing a second-order rate constant of k2 = 1.3 ± 0.1 × 10−3 M−1 s−1 at 298 K, the reaction rate of [3]+ saturates at high substrate concentration, and is well-fit by a kobs expression including a substrate-binding pre-equilibrium (Figure 9). The parameters obtained for [3]+ are K = 3.0 ± 0.2 M−1 and k1 = 5.2 ± 0.1 × 10−3 s−1, and the faster rate observed for [3]+ vs [1]+ is attributed to the substrate binding and preorganization that bears similarity to galactose oxidase. Reaction rate analysis for [2]+ shows that the first-order rate constant varies linearly over the range [alcohol] = 0.05 – 1.0 M, similarly to [1]+, providing a slightly faster second-order rate constant of k2 = 1.8 ± 0.1 × 10−3 M−1 s−1 at 298 K (Figure 9). The first-order decay of [2•]+ in the presence of benzyl alcohol, monitored by UV-vis, directly correlates with the appearance of benzaldehyde as measured by GC analysis. Product yields are consistent with [2•]+ reacting as a one-electron oxidant with the following reaction stoichiometry:
Figure 9.

Dependence of the pseudo-first-order rate constant kobs on benzyl alcohol concentration for [1]+ (circles, shown with linear fit),[52] [2•]+ (triangles, shown with linear fit), and [3]+ (squares, shown with saturation fit)[52] at 298 K in CH2Cl2. The concentration of [2•]+ was 0.1 mM in this study, thereby minimizing any decomposition over the time period of the experiment.
Overall, it appears that [2•]+ reacts with benzyl alcohol in a similar manner to [1]+, suggesting that even though the localized CuII-phenoxyl radical electronic structure matches that for [3]+, the rigidity of the one imine arm limits the substrate binding step. In addition, the reaction profile for [2•]+ is similar to that reported for the para-OMe substituted analogue of 4 (Figure 10), which also exhibits a localized CuII-phenoxyl radical electronic structure, and a rigid salen ligand framework.[38] Interestingly, the reactivity of 5 (Figure 10) with ethanol and methanol proceeds via a substrate-binding mechanism, similarly to [3]+.[95] The rate constants are much slower in this case, reflecting the higher C-H bond strength of these substrates. An additional comparison of the benzyl alcohol single-turnover reactivity of [1–3]+ with the Cu complexes of ligands HL1 and H2L2 (the benzyl alcohol single-turnover reactivity of [1–3]+ with the Cu complexes of ligands HL1 and) can be made.[96–99] The second order rate constant for the reaction of [CuII(L1•)(NO3−)]+ with benzyl alcohol (k2 = 4.8 × 10−3 M−1 s−1) is similar to those determined for [1–3]+, while the comparative reaction rate for [CuII(HL2•)(OAc)]+ is 10-fold faster (k2 = 2.7 × 10−2 M−1 s−1). The reasons for the rate acceleration for the latter system are unknown.[98] The 1:1 stoichiometry of the reaction of benzyl alcohol with [CuII(L1•)(NO3−)]+ or [CuII(HL2•)(OAc)]+ reflects the accessibility of both phenoxyl and CuII oxidizing equivalents, while only the phenoxyl is accessible for [1–3]+. The square-planar geometry favoured by the salen-type ligands in 1–3 disfavours reduction to the CuI state.
Figure 10.

Representative compounds that have been investigated for reactivity with benzyl alcohol under similar single-turnover reactivity conditions to [1–3]+. Note one resonance form is shown for [4•]+ and [5•]+.
4. Summary
In this work we have characterized the localized CuII-phenoxyl radical electronic structure of [2•]+. On the basis of both experimental data and theoretical calculations, we show that the ligand radical is localized on the more electron-rich aminophenoxide. Unlike the oxidized Ni analogue,[54] [2•]+ decomposes slowly in solution at RT, and a decay product [2H]+ is formed. Further analysis is consistent with the formation of [2H]+ and Cu salen 1 in a 2:1 ratio, however the mechanistic details have yet to be studied in detail. [2•]+ is however stable at low concentrations (ca. 0.1 mM) at RT, and reactivity studies with benzyl alcohol show that this complex reacts similarly to [1]+, and does not exhibit a substrate-binding pre-equilibrium.
Supplementary Material
We characterize a localized ligand radical via oxidation of an asymmetric Cu salan complex
The oxidized Cu salan complex decomposes over time to a protonated phenoxide and a Cu salen in a 2:1 ratio
Under single-turn-over conditions the oxidized Cu salan complex oxidizes benzyl alcohol without a substrate-binding pre-equilibrium
Acknowledgments
This work was supported by a NIH grant (GM-120187) to TDPS, TS thanks NSERC for a postdoctoral scholarship. Westgrid is thanked for access to computational resources. We acknowledge Prof. Yoshinori Naruta and Prof. Fumito Tani of Kyushu University for measurement of the resonance Raman spectra, and this work was supported in part by the Cooperative Research Program of “Network Joint Research Center for Materials and Devices” (Institute for Materials Chemistry and Engineering, Kyushu University). ECW thanks CSU Chico College of Natural Sciences for start-up funding. Dr. Benjamin Kucera is acknowledged for assistance with X-ray crystallography. Use of the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, is supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the DOE Office of Biological and Environmental Research, and by the National Institutes of Health, National Institute of General Medical Sciences (including P41GM103393). The contents of this publication are solely the responsibility of the authors and do not necessarily represent the official views of NIGMS or NIH.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Chirik PJ. Inorg Chem. 2011;50:9737–9740. doi: 10.1021/ic201881k. [DOI] [PubMed] [Google Scholar]
- 2.Chirik PJ, Wieghardt K. Science. 2010;327:794–795. doi: 10.1126/science.1183281. [DOI] [PubMed] [Google Scholar]
- 3.Praneeth VKK, Ringenberg MR, Ward TR. Angew Chem Int Ed. 2012;51:10228–10234. doi: 10.1002/anie.201204100. [DOI] [PubMed] [Google Scholar]
- 4.Luca OR, Crabtree RH. Chem Soc Rev. 2013;42:1440–1459. doi: 10.1039/c2cs35228a. [DOI] [PubMed] [Google Scholar]
- 5.Kaim W. Eur J Inorg Chem. 2012:343–348. [Google Scholar]
- 6.Lyaskovskyy V, de Bruin B. ACS Catal. 2012;2:270–279. [Google Scholar]
- 7.Rogers MS, Dooley DM. Current Opinion in Chemical Biology. 2003;7:189–196. doi: 10.1016/s1367-5931(03)00024-3. [DOI] [PubMed] [Google Scholar]
- 8.Rittle J, Green MT. Science. 2010;330:933–937. doi: 10.1126/science.1193478. [DOI] [PubMed] [Google Scholar]
- 9.Whittaker MM, Kersten PJ, Nakamura N, Sandersloehr J, Schweizer ES, Whittaker JW. J Biol Chem. 1996;271:681–687. doi: 10.1074/jbc.271.2.681. [DOI] [PubMed] [Google Scholar]
- 10.Whittaker JW. Chem Rev. 2003;103:2347–2363. doi: 10.1021/cr020425z. [DOI] [PubMed] [Google Scholar]
- 11.Whittaker MM, DeVito VL, Asher SA, Whittaker JW. J Biol Chem. 1989;264:7104–7106. [PubMed] [Google Scholar]
- 12.Whittaker MM, Whittaker JW. J Biol Chem. 1990;265:9610–9613. [PubMed] [Google Scholar]
- 13.Que L, Jr, Tolman WB. Nature. 2008;455:333–340. doi: 10.1038/nature07371. [DOI] [PubMed] [Google Scholar]
- 14.Swiegers GF, Chen J, Wagner P. Bioinspiration and Biomimicry in Chemistry. John Wiley & Sons, Inc.; 2012. Bioinspired Catalysis; pp. 165–208. [Google Scholar]
- 15.Lyons CT, Stack TDP. Coord Chem Rev. 2012;257:528–540. doi: 10.1016/j.ccr.2012.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Thomas F. Eur J Inorg Chem. 2007:2379–2404. [Google Scholar]
- 17.Clarke RM, Herasymchuk K, Storr T. Coord Chem Rev. 2017:CCR112523. In Press. [Google Scholar]
- 18.Benisvy L, Kannappan R, Song YF, Milikisyants S, Huber M, Mutikainen I, Turpeinen U, Gamez P, Bernasconi L, Baerends EJ, Hartl F, Reedijk J. Eur J Inorg Chem. 2007:637–642. [Google Scholar]
- 19.de Bellefeuille D, Askari MS, Lassalle-Kaiser B, Journaux Y, Aukauloo A, Orio M, Thomas F, Ottenwaelder X. Inorg Chem. 2012;51:12796–12804. doi: 10.1021/ic301684h. [DOI] [PubMed] [Google Scholar]
- 20.Asami K, Tsukidate K, Iwatsuki S, Tani F, Karasawa S, Chiang L, Storr T, Thomas F, Shimazaki Y. Inorg Chem. 2012;51:12450–12461. doi: 10.1021/ic3018503. [DOI] [PubMed] [Google Scholar]
- 21.Chiang L, Kochem A, Jarjayes O, Dunn TJ, Vezin H, Sakaguchi M, Ogura T, Orio M, Shimazaki Y, Thomas F, Storr T. Chem Eur J. 2012;18:14117–14127. doi: 10.1002/chem.201201410. [DOI] [PubMed] [Google Scholar]
- 22.Shimazaki Y, Arai N, Dunn TJ, Yajima T, Tani F, Ramogida CF, Storr T. Dalton Trans. 2011;40:2469–2479. doi: 10.1039/c0dt01574a. [DOI] [PubMed] [Google Scholar]
- 23.Shimazaki Y, Stack TDP, Storr T. Inorg Chem. 2009;48:8383–8392. doi: 10.1021/ic901003q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Storr T, Verma P, Pratt RC, Wasinger EC, Shimazaki Y, Stack TDP. J Am Chem Soc. 2008;130:15448–15459. doi: 10.1021/ja804339m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Storr T, Wasinger EC, Pratt RC, Stack TDP. Angew Chem Int Ed. 2007;46:5198–5201. doi: 10.1002/anie.200701194. [DOI] [PubMed] [Google Scholar]
- 26.Rotthaus O, Jarjayes O, Philouze C, Del Valle CP, Thomas F. Dalton Trans. 2009:1792–1800. doi: 10.1039/b811702k. [DOI] [PubMed] [Google Scholar]
- 27.Rotthaus O, Jarjayes O, Thomas F, Philouze C, Saint-Aman E, Pierre JL. Dalton Trans. 2007:889–895. doi: 10.1039/b612068g. [DOI] [PubMed] [Google Scholar]
- 28.Rotthaus O, Jarjayes O, Del Valle CP, Philouze C, Thomas F. Chem Commun. 2007:4462–4464. doi: 10.1039/b710027b. [DOI] [PubMed] [Google Scholar]
- 29.Rotthaus O, Thomas F, Jarjayes O, Philouze C, Saint-Aman E, Pierre JL. Chem Eur J. 2006;12:6953–6962. doi: 10.1002/chem.200600258. [DOI] [PubMed] [Google Scholar]
- 30.Rotthaus O, Jarjayes O, Thomas F, Philouze C, Del Valle CP, Saint-Aman E, Pierre JL. Chem Eur J. 2006;12:2293–2302. doi: 10.1002/chem.200500915. [DOI] [PubMed] [Google Scholar]
- 31.Kochem Al, Kanso H, Baptiste B, Arora H, Philouze C, Jarjayes O, Vezin H, Luneau D, Orio M, Thomas F. Inorg Chem. 2012;51:10557–10571. doi: 10.1021/ic300763t. [DOI] [PubMed] [Google Scholar]
- 32.Kochem A, Jarjayes O, Baptiste B, Philouze C, Vezin H, Tsukidate K, Tani F, Orio M, Shimazaki Y, Thomas F. Chem Eur J. 2012;18:1068–1072. doi: 10.1002/chem.201102882. [DOI] [PubMed] [Google Scholar]
- 33.Shimazaki Y, Yajima T, Tani F, Karasawa S, Fukui K, Naruta Y, Yamauchi O. J Am Chem Soc. 2007;129:2559–2568. doi: 10.1021/ja067022r. [DOI] [PubMed] [Google Scholar]
- 34.Shimazaki Y, Tani F, Fukui K, Naruta Y, Yamauchi O. J Am Chem Soc. 2003;125:10512–10513. doi: 10.1021/ja035806o. [DOI] [PubMed] [Google Scholar]
- 35.Pratt RC, Lyons CT, Wasinger EC, Stack TDP. J Am Chem Soc. 2012;134:7367–7377. doi: 10.1021/ja211247f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Verma P, Pratt RC, Storr T, Wasinger EC, Stack TDP. Proc Natl Acad Sci. 2011;108:18600–18605. doi: 10.1073/pnas.1109931108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Orio M, Philouze C, Jarjayes O, Neese F, Thomas F. Inorg Chem. 2010;49:646–658. doi: 10.1021/ic901846u. [DOI] [PubMed] [Google Scholar]
- 38.Orio M, Jarjayes O, Kanso H, Philouze C, Neese F, Thomas F. Angew Chem Int Ed. 2010;49:4989–4992. doi: 10.1002/anie.201001040. [DOI] [PubMed] [Google Scholar]
- 39.Kurahashi T, Fujii H. J Am Chem Soc. 2011;133:8307–8316. doi: 10.1021/ja2016813. [DOI] [PubMed] [Google Scholar]
- 40.Lecarme L, Chiang L, Philouze C, Jarjayes O, Storr T, Thomas F. Eur J Inorg Chem. 2014:3479–3487. [Google Scholar]
- 41.Dunn TJ, Webb MI, Hazin K, Verma P, Wasinger EC, Shimazaki Y, Storr T. Dalton Trans. 2013;42:3950–3956. doi: 10.1039/c2dt32632a. [DOI] [PubMed] [Google Scholar]
- 42.Dunn TJ, Chiang L, Ramogida CF, Hazin K, Webb MI, Katz MJ, Storr T. Chem Eur J. 2013;19:9606–9618. doi: 10.1002/chem.201300798. [DOI] [PubMed] [Google Scholar]
- 43.Kurahashi T, Fujii H. Inorg Chem. 2013;52:3908–3919. doi: 10.1021/ic302677f. [DOI] [PubMed] [Google Scholar]
- 44.de Bellefeuille D, Orio M, Barra AL, Aukauloo A, Journaux Y, Philouze C, Ottenwaelder X, Thomas F. Inorg Chem. 2015;54:9013–9026. doi: 10.1021/acs.inorgchem.5b01285. [DOI] [PubMed] [Google Scholar]
- 45.McGarrigle EM, Gilheany DG. Chem Rev. 2005;105:1563–1602. doi: 10.1021/cr0306945. [DOI] [PubMed] [Google Scholar]
- 46.Canali L, Sherrington DC. Chem Soc Rev. 1999;28:85–93. [Google Scholar]
- 47.Darensbourg DJ. Chem Rev. 2007;107:2388–2410. doi: 10.1021/cr068363q. [DOI] [PubMed] [Google Scholar]
- 48.Irie R, Noda K, Ito Y, Matsumoto N, Katsuki T. Tetrahedron Lett. 1990;31:7345–7348. [Google Scholar]
- 49.Jacobsen EN, Zhang W, Muci AR, Ecker JR, Deng L. J Am Chem Soc. 1991;113:7063–7064. [Google Scholar]
- 50.Cozzi PG. Chem Soc Rev. 2004;33:410–421. doi: 10.1039/b307853c. [DOI] [PubMed] [Google Scholar]
- 51.Clarke RM, Storr T. Dalton Trans. 2014;43:9380–9391. doi: 10.1039/c4dt00591k. [DOI] [PubMed] [Google Scholar]
- 52.Pratt RC, Stack TDP. J Am Chem Soc. 2003;125:8716–8717. doi: 10.1021/ja035837j. [DOI] [PubMed] [Google Scholar]
- 53.Saint-Aman E, Ménage S, Pierre JL, Defrancq E, Gellon G. New J Chem. 1998;22:393–394. [Google Scholar]
- 54.Storr T, Verma P, Shimazaki Y, Wasinger EC, Stack TDP. Chem Eur J. 2010;16:8980–8983. doi: 10.1002/chem.201001401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Nakano K, Nakamura M, Nozaki K. Macromolecules. 2009;42:6972–6980. [Google Scholar]
- 56.Berkessel A, Brandenburg M, Leitterstorf E, Frey J, Lex J, Schafer M. Adv Synth Catal. 2007;349:2385–2391. [Google Scholar]
- 57.Lansing M, Engler H, Leuther TM, Neudörfl JM, Berkessel A. ChemCatChem. 2016;8:3706–3709. [Google Scholar]
- 58.Talsi EP, Rybalova TV, Bryliakov KP. ACS Catal. 2015;5:4673–4679. [Google Scholar]
- 59.Kannan M, Punniyamurthy T. Tetrahedron Asymm. 2014;25:1331–1339. [Google Scholar]
- 60.Ziegler JE, Du G, Fanwick PE, Abu-Omar MM. Inorg Chem. 2009;48:11290–11296. doi: 10.1021/ic901794h. [DOI] [PubMed] [Google Scholar]
- 61.Fujita H, Uchida T, Irie R, Katsuki T. Chem Lett. 2007;36:1092–1093. [Google Scholar]
- 62.Campbell EJ, Nguyen ST. Tetrahedron Lett. 2001;42:1221–1225. [Google Scholar]
- 63.Audebert P, Capdevielle P, Maumy M. New J Chem. 1991;15:235–237. [Google Scholar]
- 64.Pasini A, Bernini E, Scaglia M, DeSantis G. Polyhedron. 1996;15:4461–4467. [Google Scholar]
- 65.Abuelwafa SM, Issa RM, McAuliffe CA. Inorg Chim Acta. 1985;99:103–106. [Google Scholar]
- 66.Böttcher A, Elias H, Jäger EG, Langfelderova H, Mazur M, Müller L, Paulus H, Pelikan P, Rudolph M, Valko M. Inorg Chem. 1993;32:4131–4138. [Google Scholar]
- 67.Freire C, de Castro B. J Chem Soc Dalton Trans. 1998:1491–1497. [Google Scholar]
- 68.Rehr JJ, Deleon JM, Zabinsky SI, Albers RC. J Am Chem Soc. 1991;113:5135–5140. [Google Scholar]
- 69.George GN. EXAFSPAK & EDG_FIT. Stanford Synchrotron Radiation Laboratory, Stanford Linear Accelerator Center; Stanford University Stanford, CA; 2000. [Google Scholar]
- 70.Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery J, A J, Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam NJ, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas Ö, Foresman JB, Ortiz JV, Cioslowski J, Fox DJ. Gaussian 09, Revision D.01. Gaussian, Inc.; Wallingford, CT: 2009. [Google Scholar]
- 71.Becke AD. J Chem Phys. 1993;98:5648–5652. [Google Scholar]
- 72.Stephens PJ, Devlin FJ, Chabalowski CF, Frisch MJ. J Phys Chem. 1994;98:11623–11627. [Google Scholar]
- 73.Barone V, Cossi M, Tomasi J. J Comput Chem. 1998;19:404–417. [Google Scholar]
- 74.Tomasi J, Mennucci B, Cances E. J Mol Struct. 1999;464:211–226. [Google Scholar]
- 75.Barone V, Cossi M, Tomasi J. J Chem Phys. 1997;107:3210–3221. [Google Scholar]
- 76.Miertus S, Scrocco E, Tomasi J. Chem Phys. 1981;55:117–129. [Google Scholar]
- 77.Chiang L, Herasymchuk K, Thomas F, Storr T. Inorg Chem. 2015;54:5970–5980. doi: 10.1021/acs.inorgchem.5b00783. [DOI] [PubMed] [Google Scholar]
- 78.Schafer A, Horn H, Ahlrichs R. J Chem Phys. 1992;97:2571–2577. [Google Scholar]
- 79.Schafer A, Huber C, Ahlrichs R. J Chem Phys. 1994;100:5829–5835. [Google Scholar]
- 80.Connelly NG, Geiger WE. Chem Rev. 1996;96:877–910. doi: 10.1021/cr940053x. [DOI] [PubMed] [Google Scholar]
- 81.Robin MB, Day P. Adv Inorg Chem Radiochem. 1967;10:247. [Google Scholar]
- 82.D’Alessandro DM, Keene FR. Chem Soc Rev. 2006;35:424–440. doi: 10.1039/b514590m. [DOI] [PubMed] [Google Scholar]
- 83.DuBois JL, Mukherjee P, Stack TDP, Hedman B, Solomon EI, Hodgson KO. J Am Chem Soc. 2000;122:5775–5787. [Google Scholar]
- 84.Kau LS, Spira-Solomon DJ, Penner-Hahn JE, Hodgson KO, Solomon EI. J Am Chem Soc. 1987;109:6433–6442. [Google Scholar]
- 85.Westre TE, Kennepohl P, DeWitt JG, Hedman B, Hodgson KO, Solomon EI. J Am Chem Soc. 1997;119:6297–6314. [Google Scholar]
- 86.Scott RA. Methods Enzymol. 1985;117:414–459. [Google Scholar]
- 87.Chaudhuri P, Wieghardt K. Prog Inorg Chem. 2001;50:151–216. [Google Scholar]
- 88.Sokolowski A, Leutbecher H, Weyhermüller T, Schnepf R, Both E, Bill E, Hildebrandt P, Wieghardt K. Journal of Biological Inorganic Chemistry. 1997;2:444–453. [Google Scholar]
- 89.Sokolowski A, Müller J, Weyhermüller T, Schnepf R, Hildebrandt P, Hildenbrand K, Bothe E, Wieghardt K. J Am Chem Soc. 1997;119:8889–8900. [Google Scholar]
- 90.Mukherjee A, McGlashen ML, Spiro TG. J Phys Chem. 1995;99:4912–4917. [Google Scholar]
- 91.Schnepf R, Sokolowski A, Müller J, Bachler V, Wieghardt K, Hildebrandt P. J Am Chem Soc. 1998;120:2352–2364. [Google Scholar]
- 92.Evans DF. J Chem Soc. 1959:2003–2005. [Google Scholar]
- 93.Müller J, Weyhermüller T, Bill E, Hildebrandt P, Ould-Moussa L, Glaser T, Wieghardt K. Angew Chem Int Ed. 1998;37:616–619. doi: 10.1002/(SICI)1521-3773(19980316)37:5<616::AID-ANIE616>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
- 94.Thomas F, Kochem A, Molloy JK, Gellon G, Leconte N, Philouze C, Jarjayes O, Berthiol F. Chem Eur J. ASAP; [DOI] [PubMed] [Google Scholar]
- 95.Chaudhuri P, Hess M, Müller J, Hildenbrand K, Bill E, Weyhermüller T, Wieghardt K. J Am Chem Soc. 1999;121:9599–9610. [Google Scholar]
- 96.Itoh S, Taki M, Takayama S, Nagatomo S, Kitagawa T, Sakurada N, Arakawa R, Fukuzumi S. Angew Chem Int Ed. 1999;38:2774–2776. doi: 10.1002/(sici)1521-3773(19990917)38:18<2774::aid-anie2774>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
- 97.Taki M, Kumei H, Itoh S, Fukuzumi S. J Inorg Biochem. 2000;78:1–5. doi: 10.1016/s0162-0134(99)00198-1. [DOI] [PubMed] [Google Scholar]
- 98.Thomas F, Gellon G, Gautier-Luneau I, Saint-Aman E, Pierre JL. Angew Chem Int Ed. 2002;41:3047–3050. doi: 10.1002/1521-3773(20020816)41:16<3047::AID-ANIE3047>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
- 99.Pratt RC, Stack TDP. Inorg Chem. 2005;44:2367–2375. doi: 10.1021/ic048695i. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
