ACS Publications. Most Trusted. Most Cited. Most Read
Transient Formation and Reactivity of a High-Valent Nickel(IV) Oxido Complex
My Activity

Figure 1Loading Img
  • Open Access
Article

Transient Formation and Reactivity of a High-Valent Nickel(IV) Oxido Complex
Click to copy article linkArticle link copied!

View Author Information
Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, Faculty of Science and Engineering, University of Groningen, Nijenborgh 4, 9747AG, Groningen, The Netherlands
IQCC & Departament de Química, Universitat de Girona, Campus Montilivi (Ciències), 17003 Girona, Spain
# ICREA, Pg. Lluís Companys 23, 08010 Barcelona, Spain
§ Sustainable Materials Characterisation, Van’t Hoff Institute for Molecular Sciences, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
Open PDFSupporting Information (4)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2017, 139, 25, 8718–8724
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.7b04158
Published June 5, 2017

Copyright © 2017 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Abstract

Click to copy section linkSection link copied!

A reactive high-valent dinuclear nickel(IV) oxido bridged complex is reported that can be formed at room temperature by reaction of [(L)2Ni(II)2(μ-X)3]X (X = Cl or Br) with NaOCl in methanol or acetonitrile (where L = 1,4,7-trimethyl-1,4,7-triazacyclononane). The unusual Ni(IV) oxido species is stabilized within a dinuclear tris-μ-oxido-bridged structure as [(L)2Ni(IV)2(μ-O)3]2+. Its structure and its reactivity with organic substrates are demonstrated through a combination of UV–vis absorption, resonance Raman, 1H NMR, EPR, and X-ray absorption (near-edge) spectroscopy, ESI mass spectrometry, and DFT methods. The identification of a Ni(IV)-O species opens opportunities to control the reactivity of NaOCl for selective oxidations.

Copyright © 2017 American Chemical Society

Introduction

Click to copy section linkSection link copied!

Metalloenzymes are central to the functioning of biological systems, especially in oxidative transformations and protection against reactive oxygen species. (1) Fe-, Cu-, and Mn-dependent metalloenzymes have been studied extensively and have stimulated the design and synthesis of structural and functional model complexes, especially in the search for synthetic analogues of reactive bioinorganic intermediates. (2) Recently, attention has turned to synthetic nickel-based complexes due to both enzymes such as NOD (nickel oxide dismutase) (3) and their potential in the activation of small molecules, including H2O2, (4) mCPBA, (5) and NaOCl. (6) In the latter case, such complexes open the possibility to achieve selective alkane chlorination and alkane and alkene oxygenation.
Recently, several high-valent (Ni(III) (4, 7) and Ni(IV) (6c, 8)) intermediates were identified spectroscopically. In contrast to organometallic Ni(IV) complexes, (8) the formation of Ni(III) and, more so, Ni(IV) oxido complexes, although inferred, is controversial due to the implications of the oxo-wall premise. (9) Nevertheless, several Ni(II) and Ni(III) oxido or peroxido complexes, formed with O2 and H2O2, have been characterized already at low temperature. (24-27)
Here we show that a novel room-temperature-stable dinuclear Ni(IV) oxido complex (3), [(L)2Ni(IV)2(μ-O)3]2+ (where L = 1,4,7-trimethyl-1,4,7-triazacyclononane), can be generated by reaction of the Ni(II) complexes (1, 2) with NaOCl (Scheme 1). The high oxidation state of 3 is stabilized by the tris-μ-oxido-bridged dinuclear structure. Furthermore, the reactivity of 3 toward direct C–H oxidation of organic substrates is demonstrated.

Scheme 1

Scheme 1. Formation of 3 from 1 and NaOCl
Complex 3 is obtained within seconds of addition of near stoichiometric amounts of NaOCl to [(L)2Ni(II)2(μ-X)3]X·(H2O)n (where X = Cl (1) or Br (2), n = 5 or 7) (10) in methanol or acetonitrile at room temperature. The structure of 3 was elucidated through a combination of UV–vis absorption and resonance Raman spectroscopy and ESI mass spectrometry, supported by isotope labeling, crossover experiments, and DFT methods.

Results and Discussion

Click to copy section linkSection link copied!

Complexes 1 and 2 were prepared by methods analogous to those reported earlier (see SI for details). (10) The Raman spectra of 1 in the solid state (Figure S1) and in solution (i.e., methanol, acetonitrile, Figure S2) indicate that the complexes retain their structure; that is, the ligand (L) remains bound to Ni(II) upon dissolution. However, ESI mass spectrometry indicates that 1 can form mononuclear complexes with three (176.3 m/z: 4 [(L)Ni(II)(CH3CN)3)]2+) or two acetonitrile ligands (155.9 m/z: 5, ([(L)Ni(II)(CH3CN)2)]2+) by substitution of the chlorido ligands (Figure S3). Indeed, mixtures of 1 and 2 show rapid exchange of Cl and Br ligands (Figure S4). The 1H NMR spectrum of 1 in acetonitrile shows paramagnetically shifted and broadened signals at ca. 60, 90, and 120 ppm (Figure S5), and its UV–vis absorption spectrum in methanol and acetonitrile (Figure S6) shows bands at 384, 635, and 1014 nm. The DFT data (vide infra) indicate antiferromagnetically coupled Ni(II) ions in 1 and 2, consistent with observations by Wieghardt et al. (11) The C3h-symmetric DFT structure for both 1 and 2 is in excellent agreement with the X-ray structure, (11) with differences of ca. 0.01–0.02 Å between DFT and X-ray for Ni–N, Ni–Cl/Br, and Ni–Ni distances (Table S1). The antiferromagnetically coupled state lies lowest in energy (including COSMO solvation (12) and ZORA (13) relativistic effects, Table S2), (14, 15) in excellent agreement with experiment. (11) In this state, the locally spin-polarized triplet (S = 1) Ni(II) ions couple antiferromagnetically (AFM) to reach overall an open-shell singlet state. As a result, spin-up spin-density is observed around one Ni and spin-down around the other (Figure 1). This AFM state is only 0.3 kcal·mol–1 lower than the ferromagnetically (FM) coupled (S = 2) state (see SI section 7 for the corresponding spin-density plot). Both states would be consistent with the paramagnetic 1H NMR shifts (vide supra), but the AFM state is found to be slightly lower by both theory and experiment. The diamagnetic closed-shell singlet state, in which the Ni(II) ions are now locally in a closed-shell (S = 0) state, is higher in energy than the AFM state by 37.7 kcal·mol–1 (1) and 35.0 kcal·mol–1 (2). Finally, also an overall intermediate spin state (S = 1) was obtained, 25.5 kcal·mol–1 (1) and 23.4 kcal·mol–1 (2) higher in energy than the AFM state, where spin-density corresponding to the single occupation of an antibonding dx2y2 orbital was observed on both Ni ions. Finally, the experimental NMR and UV–vis absorption spectra of the AFM state are similar to those of related Ni(II) complexes (Figure S7). (16)

Figure 1

Figure 1. Spin-density plot (S12g/TZ2P) for (antiferromagnetically coupled) 1 with spin-up spin-density shown in blue (around Nia, left) and spin-down spin-density in red (around Nib, right).

Addition of NaOCl to 1 or 2 in methanol or in acetonitrile leads to a rapid increase in absorbance at 363 and 612 nm due to formation of 3 (Figure 1). In methanol, the visible absorption band decreases with a t1/2 of ca. 50 s at 20 °C (Figure S8), and the absorption spectrum after 20 min is similar to the initial spectrum, with only a minor shift from 388 to 378 nm (Figure 2). Notably, the rate of decay of 3 in methanol was substantially lower than the rate of the direct reaction of NaOCl with methanol in the absence of 1 (Figure S8), indicating that the formation of 3 competes with oxidation of methanol by NaOCl and that 3 is less reactive in the oxidation of methanol than NaOCl.

Figure 2

Figure 2. UV–vis absorption spectrum of 1 (3.5 mM) in methanol before (black) and after (red 6 s, blue 48 s, green 463 s) addition of 11 equiv of NaOCl(aq) at 293 K. Inset: Expansion of the NIR region. The data indicate that for 3 ε612 nm > 715 M–1·cm–1.

In both the presence and absence of 1, near-quantitative (cf. NaOCl) oxidation of methanol to formaldehyde occurs (1.5:1, Figure S9), indicating that although at least 2 equiv of NaOCl are consumed in forming 3 from 1, these oxidation equivalents are still available for subsequent oxidation of methanol. Further additions of NaOCl in the presence of 1 resulted in the reappearance of the 612 nm absorption (Figure S10), confirming the integrity of the catalyst under reaction conditions. Similar changes were observed with NaOBr (Figure S11), which indicates that the same intermediate is formed with both oxidants (vide infra). Addition of H2O2 or purging with O2 (Figures S12 and S13) did not result in the appearance of 3.
Intermediate 3 forms upon addition of NaOCl to 1 also in acetonitrile (absorbance band at 612 nm), but persists for a substantially longer time period than in methanol, with a t1/2 of ca. 10 min at 20 °C and over 6 h at −15 °C (Figures S14 and S15), enabling characterization by resonance Raman spectroscopy, XANES, XES, and ESI mass spectrometry and reactivity with other substrates to be studied (vide infra). The maximum transient absorbance at 612 nm was obtained with 3–4 equiv of NaOCl (Figure S16) and is accompanied by an increase in oxidation state as confirmed by XANES and XES (Figure 3 and Figure S17).

Figure 3

Figure 3. (Top) Ni K XANES before (blue) and after (yellow) addition of 4 equiv of NaOCl to 1 (30 mM) in acetonitrile. (Bottom) Ni K edge XANES for structures 1, 3, and 3a, simulated using FEFF9.0 (18) using coordinates available from DFT-optimized structures (see the SI).

The Ni K edge X-ray absorption near edge structure (XANES) as well as the X-ray emission (XES) spectra were recorded for 1 (30 mM) in acetonitrile before and after addition of NaOCl. Spectra were acquired over 5 min after addition when the concentration of 3 (generated from 1) was substantial, although Ni(II) species are present also. The Ni K edge XANES spectra of both 1 and the mixture of species formed after addition of NaOCl show a nondescript edge, i.e., no significant pre- edge features, which indicates a near-octahedral, six-coordinated, geometry around the Ni center for all species present (consistent with the structures 1, 3, and 3a).
The edge position is a function of the ligands present and geometry as well as oxidation state, and hence a direct conclusion based on XANES alone cannot be drawn. For example, for purely oxidic Ni systems, an energy shift of about 1.8 eV per oxidation state is reported, with a much lower shift for sulfur-based systems. (17) A conclusive statement as to the oxidation state requires the availability of a series of reference compounds with similar ligands and geometries (data for closely related structures are not yet available in the literature).
The energy shift of about 2 eV observed in the experimental spectra is consistent with an increase in oxidation state, but it will also reflect the change from chlorido to oxido ligands. Furthermore, the mixture of species available (including Ni(II)) reduces the magnitude of the observed shift and complicates interpretation.
The simulated XANES for complexes 1, 3, and 3a are displayed in Figure 3. (18) All XANES spectra have only a pre-edge feature at low energy and an otherwise featureless edge, consistent with six-coordinated geometries. A clear difference in the spectra of the Ni(II) 1 of 5 eV with the Ni(III) 3a and Ni(IV) 3a species is calculated, i.e., upon oxidation and change of ligands from Cl to O. It is, however, also clear that XANES data will not allow for distinguishing the Ni(III) and Ni(IV) species, i.e., complexes 3 and 3a. Furthermore, the Ni K edge XANES of 1 undergoes only a few electronvolts shift in the presence of NaOCl and not the full 5 eV shift, as it is a mixture of Ni(II) and nickel in higher oxidation states, resulting in only an average shift, i.e., between 2 and 3 eV, reflecting an overall average increase in oxidation state. The XES data at the Kβ1,3 edge also display an overall shift of a few electronvolts; however, the emission lines are sensitive to both oxidation state and ligand type also.
The three broad signals in the 1H NMR spectrum of 1 in acetonitrile decrease upon addition of NaOCl, suggesting the formation of a diamagnetic species, and then recover concomitant with the increase and decrease in absorbance at 612 nm (Figure S18). Samples flash frozen to 77 K at any time, however, did not indicate the presence of a mononuclear Ni(III) species by X-band EPR spectroscopy.
Generation of 3 by electrochemical oxidation of 1 was explored also. The cyclic voltammetry of 1 shows irreversible redox waves at 1.52, 1.35, and 1.05 V and an irreversible redox wave at 0.58 V on the return cycle. At higher scan rates (up to 10 V s–1) the initial oxidation wave at 1.05 V shows some evidence of chemical reversibility; however, the shifts in Ip,a (which are corrected for IRu) indicate that the oxidation is electrochemically irreversible also (Figure S19). Addition of NaOCl to 1 results (after 2 min) in a shift in the oxidation wave to 1.16 V, indicative of ligand exchange, e.g., CH3CN replacing Cl. It should be noted, however, that the redox chemistry of NaOCl, concentration polarization, and the likely complex series of dis- and comproportionation reactions between 1 and 3 preclude the observation of a redox wave assignable to 3. One hour after addition of NaOCl, a further shift is observed, indicating ligand exchange to form a Ni(II) complex similar to 4 (Figure S20). Notably the voltammetry and spectroelectrochemistry were not affected substantially by the addition of water and NaCl (i.e., at concentrations present under reaction conditions).
The spectroelectrochemistry of 1 in acetonitrile shows that oxidation at 1.2 V leads to essentially no change in the UV absorption (Figure S21a). On the second cycle the redox wave is shifted to >1.58 V; again relatively little change in absorbance is observed, and after reduction below 0.86 V only minor shifts in absorbance at ca. 300 nm are observed, consistent with ligand exchange (Figure S21b). Bulk electrolysis of 1 shows an increase in absorption at 344 nm; however, the EPR spectrum (X-band 77 K) of this sample was silent, and hence oxidation leads to ligand exchange to a Ni(II) complex with a more positive redox potential rather than formation of a Ni(III) or Ni(IV) complex.
Attempts to isolate 3 by flash precipitation with KPF6 yielded the mononuclear Ni(II) compound [(L)Ni(II)(CH3CN)3]2+ (4) instead, reported earlier by Tak et al. as the B(Ph)4 salt, (16) in which the chlorido ligands are replaced by acetonitrile ligands to form a mononuclear complex (see the SI and Scheme 2). Notably, addition of NaOCl to 4 in acetonitrile results in the same visible absorption spectrum as obtained with 1 (Figure S23), indicating that the formation of 3 is not dependent on the initial form of the LNi(II) complex (with Cl/Br/CH3CN) and is consistent with the rapid equilibration of these species in solution. Furthermore, addition of NaOCl (4–10 equiv) to a 1:1 mixture of NiCl2 and the ligand (L) in acetonitrile results in the appearance of the bands at λmax 363 and 612 nm (Figure S24), consistent with a maximum 50% conversion to 3. In the absence of ligand (i.e., only NiCl2·6H2O) the band at 612 nm was not observed (Figure S25).

Scheme 2

Scheme 2. Formation of 3 from 1 and NaOCl and Subsequent Decay to 4 in Acetonitrile
In earlier reports, (6) the presence of acetic acid was necessary for the formation of high-valent nickel complexes with NaOCl. Addition of acetic acid (4.5 equiv) to 3 in acetonitrile (generated by addition of 4.5 equiv of NaOCl to 1 (0.9 mM)) did not affect the absorbance at 612 nm significantly (Figure S26a). In contrast, addition of acetic acid (4.5 equiv) prior to addition of NaOCl to 1 (0.9 mM) precluded the appearance of the 612 nm band and hence formation of 3 (Figure S26b). (19)
The ESI mass spectra of 1 in acetonitrile show signals (m/z) assignable to [Ni(II)(L)(CH3CN)2]2+ (155.9 m/z), [Ni(II)(L)(CH3CN)3]2+ (176.3 m/z), [Ni(II)(L) (Cl)]+ (264.1 m/z), and [Ni(II)2(L)2(Cl)3]+ (565.2 m/z) (Figure S27). The spectrum obtained from a solution containing 3 shows an additional strong signal at 253.3 m/z with an isotope distribution consistent with two Ni centers (Figure S28), regardless of whether it was generated with NaOBr or NaOCl and with 2 in place of 1 (Figures S29–S33). Notably, however, the signal increased by 3 m/z units with Na18OCl (Figure S31). The m/z signal at 253.3 is therefore consistent with structures such as the peroxy-bridged 3a, (20) [(L)2Ni(III)2(μ-O)(μ-O-O)]2+, the mono-μ-oxo-bis-terminal-oxo 3b, [(L)-(O)═Ni(IV)-O-Ni(IV)═(O)-L]2+, and the tri-μ-oxido-bridged 3, [((L)2Ni(IV)2(μ-O)3]2+ (Scheme 3), with the latter structure 3 favored on the basis of Raman spectroscopy (vide infra) and DFT. (21)

Scheme 3

Scheme 3. Structures (3, 3a, 3b) Consistent with ESI Mass Spectral Data and Calculated Driving Forces for Their Formation from 1
All possible spin states were explored by DFT methods for 3, 3a, and 3b. As expected for the d6 Ni(IV) ions in 3, the lowest energy (15) state corresponds to locally closed-shell singlet (S = 0) Ni(IV) ions. For consistency, we explored the possibility of other spin states and found several higher lying spin states, where, for example, the Ni(IV) ions had locally a triplet (S = 1) state that coupled to form overall a quintet (S = 2) state. However, these other spin states are >27 kcal·mol–1 higher in energy (see Table S2) and will not be discussed any further. For the d7 Ni(III) ions in 3a, the lowest energy (15) state corresponds to a doublet on each of the metals, which can be FM (S = 1) or AFM (open-shell singlet) coupled. The latter open-shell singlet state is lower in (Gibbs free) energy than the triplet by 2.1 kcal·mol–1, and 5.1 kcal·mol–1 lower than the diamagnetic closed-shell singlet state. All other spin states for 3a are >10 kcal·mol–1 higher in (Gibbs free) energy. The lowest energy for the d6 Ni(IV) ions in 3b corresponds to an AFM state, where each of the Ni(IV) ions is found locally in a triplet (S = 1) state. The AFM state is lower than the FM state in 3b by 1.2 kcal·mol–1, with the other spin states higher in energy by >10 kcal·mol–1.
The geometrical parameters for compound 3, which although isostructural to [(L)2Mn(IV)2(μ-O)3]2+, (22) are found to be somewhat different. The most prominent feature in [(L)2Mn(IV)2(μ-O)3]2+ is a short Mn(IV)–Mn(IV) distance of 2.30 Å, which is absent in 3, where the Ni(IV)–Ni(IV) is instead 2.46 Å; the latter distance is more similar to the bis-μ-oxo variant [L2MnIV2(O)2(μ-O)2] (2.62 Å). (22) Furthermore, the NiIV–N distance of 2.02 Å in 3 is substantially shorter than in [(L)2Mn(IV)2(μ-O)3]2+ (2.11 Å). The NiIV–(μ-O) distance in 3 of 1.85 Å is however similar to that in the MnIV analogue (1.82 Å). Overall, the DFT structure for 3 and the X-ray structure for [(L)2Mn(IV)2(μ-O)3]2+ are similar, which is apparent when superimposed (Figure S34).

Figure 4

Figure 4. Raman spectra (λexc 532 nm) of 1 (3.5 mM) in acetonitrile (a) and with (b) 4.5 equiv of NaOBr, (c) 4.5 equiv of Na16OCl, and (d) 4.5 equiv of Na18OCl. *Solvent band. #Raman band from quartz.

The Raman spectrum of 3, with excitation resonant with the visible absorption band, shows enhanced Raman scattering at 801, 631, and 521 cm–1 (Figure 3), with only the band at 631 cm–1 (Δ[18O] = 32 cm–1) affected by the use of Na18OCl. The bands are unaffected by use of OBr in place of OCl. The DFT-calculated Raman spectrum for [(L)2Ni(IV)2(μ-16/18O)3]2+ (3) shows a penta-atomic symmetric stretching Ni–(O)3–Ni mode at 638 cm–1, which shifted to 609 cm–1 upon isotope labeling (i.e., Δ[18O] = 29 cm–1, Figure S35). Mixed labeling (i.e., varying ratios of 16O and 18O) experiments show the series of four bands expected for the four isotopologues and correspond well with the DFT-calculated shifts (Figure 5).

Figure 5

Figure 5. Top: Calculated spectra for 3 with various degrees of 18O substitution (16O3 (blue), 16O218O (red), 16O18O2 (green), 18O3 (purple)). Bottom: Resonance Raman spectra of 3 generated from 1 (4 mM) in methanol by addition of 4 equiv of NaOCl/H2O with (blue) 100% 16O, (red) 50% 18O, (green) 34% 18O, and (purple) 26% 18O.

The band at 801 cm–1 corresponds to a symmetric Ni–N stretching (calcd by DFT at 782 cm–1, at the same position upon 18O labeling), while the band at 521 cm–1 involves mainly a combination of Ni–N and Ni–O stretching (calcd by DFT at 517 cm–1 and at 516 cm–1 upon 18O labeling). All of the vibrations of the normal modes are available in the SI. The modes observed for 3 are similar in energy to those of the isostructural complex [(L)2Mn(IV)2(μ-O)3]2+, which shows a Mn–(O)3–Mn stretch at 701 cm–118O = 33 cm–1). (23) Furthermore, the spectrum is similar to that reported by Riordan and co-workers for [((PhTttBu)Ni(III))2(μ-O)2], with a band at 585 cm–1 (Δ[18O] = 30 cm–1), (24) and by Fukuzumi and co-workers for [(L′Ni(III))2(μ-O)2]2+, where L′ = N,N-bis[2-(2-pyridyl)ethyl]-2-phenylethylamine, with a band at 612 cm–1 (Δ[18O] = 32 cm–1). (25) The higher Raman shift for 3 is consistent with an increase in the oxidation state from III to IV.
The Ni–O–O–Ni stretching modes in 3a are expected at higher wavenumbers compared to Ni–O–Ni modes; see, for example, the O–O stretch vibrations reported by Riordan and co-workers ([(Ni(tmc))2(μ-O-O)] at 778 cm–1; Δ[18O] = 43 cm–1) (26) and Gade and co-workers ([(Ni(iso-pmbox))2(μ-O-O)] at 742 cm–1; Δ[18O] = 36 cm–1). (27) This conclusion is supported by the DFT-calculated IR spectrum for 3a (Figure S36), which shows an O–O stretching band (Δ[18O] in parentheses) at 892 (841) cm–1, a Ni–N stretching mode at 768 (768) cm–1, and Ni–O bending/stretching modes at 663 (633), 601 (594), 585 (571), and 536 (516) cm–1. Of these modes, the 663 and 768 cm–1 modes are strongly IR active. Therefore, given that in the present system only the band at 631 cm–1 is affected by the use of 18OCl and that the DFT-calculated IR spectrum of 3a (Figure S36) indicates bands at ca. 660 and 770 cm–1 only and not a band at ca. 801 cm–1, the peroxy species can be discarded. The same is true for 3b, which shows terminal-oxo Ni–O stretches at 742 cm–1 (shifting to 718 cm–1 upon 18O labeling, i.e., Δ18O = 24 cm–1), 700 cm–118O = 28 cm–1), and 689 cm–118O = 26 cm–1); the Ni–(μ-O)–Ni stretch is found at 653 cm–118O = 17 cm–1). No bands are observed around 521 cm–1, and all bands show significant isotope effects. Therefore, although the ESI mass spectral data could correspond also to a peroxy-bridged (3a) or the mono-μ-oxo species (3b), the most appropriate structural assignment for the high-valent nickel species is 3 based on the vibrational spectra.
DFT calculations provide further support for this assignment in the thermochemistry of the reaction of 1 and 2 with NaOCl to form 3. Mass spectral data indicate that the reactive intermediate 3 has the composition [(L)2Ni2O3]2+, and the absence of EPR (X-band) signals at 77 K at any time suggests that mononuclear Ni(III) complexes are not present to a significant extent. Hence, geometry optimizations were performed with all spin multiplicities for 1, 2, 3, and 3a, as well as possible mononuclear Ni(II) complexes, e.g., [(L)Ni(II)(CH3CN)3]2+ (4) and [(L)Ni(II) (CH3CN)2]2+ (5) (Table S2). Antiferromagnetically coupled dinuclear species were found as lowest energy for 1, 2, 3a, and 3b, while a closed-shell spin state was found for 3 (vide supra); a high-spin Ni(II) (S = 1) ground state was found for 4 and 5. The reactions of 1 with NaOCl to form 3, 3a, or 3b (Scheme 3) were calculated to be exergonic by −92.5, −88.1, and −50.3 kcal·mol–1, respectively; hence, 3 is 4.46 kcal·mol–1 more stable than 3a and 42.2 kcal·mol–1 than 3b, in terms of Gibbs energy (in electronic energy: 7.50 and 46.11 kcal·mol–1, respectively for 3a and 3b), Table S3.
Further spectroscopic evidence for the formation of 3, and not 3a or 3b, is obtained by mixed labeling experiments. In these experiments we applied ratios of pure 16O (3:0), pure 18O (0:3), and 1:2/2:1 mixtures of these, such that we would have four different distributions with on average the incorporation of 0, 1, 2, and 3 labeled oxygens into the complex. The observed isotope shifts observed in our Raman spectra with these mixed labeling experiments match perfectly with the corresponding DFT isotope shifts; that is, the 631 cm–1 peak (638 cm–1 DFT) shifts to 623 cm–1 (630 cm–1 DFT), to 613 cm–1 (622 cm–1 DFT), and finally to 599 (609 cm–1 DFT).
Overall, the spectroscopic and computational data are consistent with the assignment of the intermediate as 3 ([(L)2Ni(IV)2(μ-16O)3]2+). The mechanism by which 3 forms from Ni(II) complexes undoubtedly involves multiple elementary steps. However, the coordination of OCl to Ni(II) is expected to be facile given the rapid exchange of Cl, Br, and CH3CN ligands. Heterolytic cleavage of Ni(II)–O–Cl to form a transient intermediate Ni(IV) species and Cl is presumably followed by formation of a (μ-O)3-bridged Ni(IV) dimer, which does not show antiferromagnetic coupling, in contrast with the equivalent manganese complex, but instead corresponds to diamagnetic closed-shell Ni(IV) ions.
Finally, although 3 undergoes rapid self-decay in methanol to yield formaldehyde, in acetonitrile it is relatively stable, allowing for its reactivity with organic substrates to be assessed. The addition of ca. 4 equiv of substrate (e.g., xanthene, 9,10-dihydroanthracene, and fluorene) resulted in a complete loss of absorbance at 612 nm within ca. 6 min (70, 150, and 350 s, respectively, Figure S37). At −15 °C, 3 is stable for over 6 h (Figure S38),; however, addition of 4 equiv of xanthene resulted in a rapid loss in absorbance at 612 nm (within ca. 275 s, Figure S39). Addition of 50 equiv of fluorene to 3 resulted in the disappearance of the signal at m/z 253.3 and a recovery of the signals of 1 and 4 (Figure S40). Hence, although substrates react directly with NaOCl, 3 engages in C–H oxidation also (Figure S39).

Conclusions

Click to copy section linkSection link copied!

Complex 3 represents the first example of a Ni(IV)-oxido-bridged dimer. Its generation from NaOCl and subsequent reaction with organic substrates opens up the possibility to use NaOCl as a terminal oxidant. The intermediacy of a transition metal catalyst opens the possibility of engaging in selective oxidations, thereby taming the reactivity of this potent oxidant.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b04158.

  • Animated gifs for selected vibrational modes between 517 and 893 cm−1 of 3 (ZIP)

  • Animated gifs for selected vibrational modes between 536 and 892 cm−1 of 3a (ZIP)

  • Animated gifs for vibrational modes at 653, 689, 700, 742 cm−1 of 3b (ZIP)

  • Details of synthesis and charcaterization of complexes 1, 2, and 4, UV–vis absorption, (resonance) Raman, NMR, EPR spectroscopy, ESI-MS, XANES, XES, and computational data (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Sandeep K. Padamati - Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, Faculty of Science and Engineering, University of Groningen, Nijenborgh 4, 9747AG, Groningen, The Netherlands
    • Davide Angelone - Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, Faculty of Science and Engineering, University of Groningen, Nijenborgh 4, 9747AG, Groningen, The NetherlandsIQCC & Departament de Química, Universitat de Girona, Campus Montilivi (Ciències), 17003 Girona, Spain
    • Apparao Draksharapu - Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, Faculty of Science and Engineering, University of Groningen, Nijenborgh 4, 9747AG, Groningen, The Netherlands
    • Gloria Primi - Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, Faculty of Science and Engineering, University of Groningen, Nijenborgh 4, 9747AG, Groningen, The Netherlands
    • David J. Martin - Sustainable Materials Characterisation, Van’t Hoff Institute for Molecular Sciences, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The NetherlandsOrcidhttp://orcid.org/0000-0002-3549-4202
    • Moniek Tromp - Sustainable Materials Characterisation, Van’t Hoff Institute for Molecular Sciences, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
  • Notes
    The authors declare no competing financial interest.

Acknowledgment

Click to copy section linkSection link copied!

The Ubbo Emmius fund of the University of Groningen, the European Research Council (StG, no. 279549, W.R.B.), NWO for a VIDI grant (723.014.010, D.J.M. and M.T.), The Netherlands Ministry of Education, Culture and Science (Gravity program 024.001.035, W.R.B.), MINECO (CTQ2014-59212-P and CTQ2015-70851-ERC, M.S.), GenCat (2014SGR1202, M.S.), FEDER (UNGI10-4E-801, M.S.), and COST action CM1305 “ECOSTBio” (W.R.B., COST-STSM-CM1305-29045) are acknowledged for financial support. A. van Dam (ERIBA) is thanked for assistance with mass spectrometry.

References

Click to copy section linkSection link copied!

This article references 27 other publications.

  1. 1
    (a) Costas, M.; Mehn, M. P.; Jensen, M. P.; Que, L., Jr. Chem. Rev. 2004, 104, 939 986 DOI: 10.1021/cr020628n
    (b) Solomon, E. I.; Brunold, T. C.; Davis, M. I.; Kemsley, J. N.; Lee, S.-K.; Lehnert, N.; Neese, F.; Skulan, A. J.; Yang, Y.-S.; Zhou, J. Chem. Rev. 2000, 100, 235 350 DOI: 10.1021/cr9900275
    (c) Punniyamurthy, T.; Velusamy, S.; Iqbal, J. Chem. Rev. 2005, 105, 2329 2363 DOI: 10.1021/cr050523v
  2. 2
    (a) Decker, A.; Solomon, E. I. Curr. Opin. Chem. Biol. 2005, 9, 152 163 DOI: 10.1016/j.cbpa.2005.02.012
    (b) Bertini, I.; Gray, H. B.; Stiefel, E. I.; Valentine, J. S. Biological.Inorganic Chemistry. Structure & Reactivity; University Science Books: Sausalito, CA, 2007.
    (c) Meunier, B. Chem. Rev. 1992, 92, 1411 1456 DOI: 10.1021/cr00014a008
    (d) Nam, W. Acc. Chem. Res. 2015, 48, 2415 2423 DOI: 10.1021/acs.accounts.5b00218
    (e) Oloo, W. N.; Que, L. Acc. Chem. Res. 2015, 48, 2612 2621 DOI: 10.1021/acs.accounts.5b00053
  3. 3
    Shearer, J. Acc. Chem. Res. 2014, 47, 2332 2341 DOI: 10.1021/ar500060s
  4. 4
    (a) Cho, J.; Kang, Y.; Liu, L. V.; Sarangi, R.; Solomon, E. I.; Nam, W. Chem. Sci. 2013, 4, 1502 1508 DOI: 10.1039/c3sc22173c
    (b) Honda, K.; Cho, J.; Matsumoto, T.; Roh, J.; Furutachi, H.; Tosha, T.; Kubo, M.; Fujinami, S.; Ogura, T.; Kitagawa, T.; Suzuki, M. Angew. Chem., Int. Ed. 2009, 48, 3304 3307 DOI: 10.1002/anie.200900222
    (c) Tano, T.; Doi, Y.; Inosako, M.; Kunishita, A.; Kubo, M.; Ishimaru, H.; Ogura, T.; Sugimoto, H.; Itoh, S. Bull. Chem. Soc. Jpn. 2010, 83, 530 538 DOI: 10.1246/bcsj.20090346
    (d) Kunishita, A.; Doi, Y.; Kubo, M.; Ogura, T.; Sugimoto, H.; Itoh, S. Inorg. Chem. 2009, 48, 4997 5004 DOI: 10.1021/ic900059m
    (e) Morimoto, Y.; Bunno, S.; Fujieda, N.; Sugimoto, H.; Itoh, S. J. Am. Chem. Soc. 2015, 137, 5867 5870 DOI: 10.1021/jacs.5b01814
  5. 5
    (a) Corona, T.; Pfaff, F. F.; Acua-Pares, F.; Draksharapu, A.; Whiteoak, C. J.; Martin-Diaconescu, V.; Lloret-Fillol, J.; Browne, W. R.; Ray, K.; Company, A. Chem. - Eur. J. 2015, 21, 15029 15038 DOI: 10.1002/chem.201501841
    (b) Pfaff, F. F.; Heims, F.; Kundu, S.; Mebs, S.; Ray, K. Chem. Commun. 2012, 48, 3730 3732 DOI: 10.1039/c2cc30716b
  6. 6
    (a) Draksharapu, A.; Codolá, Z.; Gómez, L.; Lloret-Fillol, J.; Browne, W. R.; Costas, M. Inorg. Chem. 2015, 54, 10656 10666 DOI: 10.1021/acs.inorgchem.5b01463
    (b) Pirovano, P.; Farquhar, E. R.; Swart, M.; McDonald, A. R. J. Am. Chem. Soc. 2016, 138, 14362 14370 DOI: 10.1021/jacs.6b08406
    (c) Corona, T.; Draksharapu, A.; Padamati, S. K.; Gamba, I.; Martin-Diaconescu, V.; Acuña-Parés, F.; Browne, W. R.; Company, A. J. Am. Chem. Soc. 2016, 138, 12987 12996 DOI: 10.1021/jacs.6b07544
  7. 7
    Corona, T.; Company, A. Chem. - Eur. J. 2016, 22, 13422 13429 DOI: 10.1002/chem.201602414
  8. 8
    (a) Camasso, N. M.; Sanford, M. S. Science 2015, 347, 1 7 DOI: 10.1126/science.aaa4526
    (b) Riordan, C. G. Science 2015, 347, 1203 DOI: 10.1126/science.aaa7553
  9. 9
    (a) Gray, H. B.; Hare, C. R. Inorg. Chem. 1962, 1, 363 368 DOI: 10.1021/ic50002a034
    (b) O’Halloran, K. P.; Zhao, C.; Ando, N. S.; Schultz, A. J.; Koetzle, T. F.; Piccoli, P. M. B.; Hedman, B.; Hodgson, K. O.; Bobyr, E.; Kirk, M. L.; Knottenbelt, S.; Depperman, E. C.; Stein, B.; Anderson, T. M.; Cao, R.; Geletii, Y. V.; Hardcastle, K. I.; Musaev, D. G.; Neiwert, W. A.; Fang, X.; Morokuma, K.; Wu, S.; Kögerler, P.; Hill, C. L. Inorg. Chem. 2012, 51, 7025 7031 DOI: 10.1021/ic2008914
  10. 10
    Wieghardt, K.; Schmidt, W.; Herrmann, W.; Kueppers, H. J. Inorg. Chem. 1983, 22, 2953 2956 DOI: 10.1021/ic00162a037
  11. 11
    Bossek, U.; Nühlen, D.; Bill, E.; Glaser, T.; Krebs, C.; Weyhermüller, T.; Wieghardt, K.; Lengen, M.; Trautwein, A. X. Inorg. Chem. 1997, 36, 2834 2843 DOI: 10.1021/ic970119h
  12. 12
    (a) Klamt, A.; Schüürmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 799 805 DOI: 10.1039/P29930000799
    (b) Pye, C. C.; Ziegler, T. Theor. Chem. Acc. 1999, 101, 396 408 DOI: 10.1007/s002140050457
    (c) Swart, M.; Rösler, E.; Bickelhaupt, F. M. Eur. J. Inorg. Chem. 2007, 2007, 3646 3654 DOI: 10.1002/ejic.200700228
  13. 13
    van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597 4610 DOI: 10.1063/1.466059
  14. 14
    (a) Swart, M. Chem. Phys. Lett. 2013, 580, 166 171 DOI: 10.1016/j.cplett.2013.06.045
    (b) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098 3100 DOI: 10.1103/PhysRevA.38.3098
    (c) Perdew, J. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33, 8822 8824 DOI: 10.1103/PhysRevB.33.8822
  15. 15

    At S12g/TZ2P//BP86-D3/TDZP.

  16. 16
    Tak, H.; Lee, H.; Kang, J.; Cho, J. Inorg. Chem. Front. 2016, 3, 157 163 DOI: 10.1039/C5QI00206K
  17. 17
    (a) Sun, Y.-K; Kim, M. G.; Kang, S. H.; Amine, K. J. Mater. Chem. 2003, 13, 319 322 DOI: 10.1039/b209379k
    (b) Gu, W.; Wang, H.; Wang, K. Dalton Trans. 2014, 43, 6406 6413 DOI: 10.1039/c4dt00308j
  18. 18
    Rehr, J. J.; Kas, J. J.; Vila, F. D.; Prange, M. P.; Jorissen, K. Phys. Chem. Chem. Phys. 2010, 12, 5503 5513 DOI: 10.1039/b926434e
  19. 19

    In the absence of Ni(II) or 1, NaOCl is stable in CH3CN.

  20. 20

    This complex is reminiscent of the (μ-η22-disulfido)dinickel(II) complexes with 6-methyl-TPA ligands reported by Itoh et al.; see:

    Inosako, M.; Kunishita, A.; Kubo, M.; Ogura, T.; Sugimoto, T.; Itoh, S. Dalton Trans. 2009, 43, 9410 9147 DOI: 10.1039/b910237j
  21. 21

    The formation of species such as (O═Ni(IV)–O–Ni(IV)═O) is highly unlikely, as it is calculated to lie 45 to 58 kcal mol–1 (depending on functional used, e.g., BP86-D3 and S12g) higher in energy than even the peroxy species 3a. Furthermore, DFT calculations indicate a high tendency for a species to converge to the structure of the 3a species. In addition ligand hydroxylation can be discounted on the basis of the recovery of the Ni(II) complexes, e.g., 1 and 4, once 3 has reacted with solvent or substrate.

  22. 22
    Wieghardt, K.; Bossek, U.; Nuber, B.; Weiss, J.; Bonvoisin, J.; Corbella, M.; Vitols, S. E.; Girerd, J. J. J. Am. Chem. Soc. 1988, 110, 7398 7411 DOI: 10.1021/ja00230a021
  23. 23
    (a) Hage, R.; Krijnen, B.; Warnaar, J. B.; Hartl, F.; Stufkens, D. J.; Snoeck, T. L. Inorg. Chem. 1995, 34, 4973 4978 DOI: 10.1021/ic00124a010
    (b) Angelone, D.; Abdolahzadeh, S.; de Boer, J. W.; Browne, W. R. Eur. J. Inorg. Chem. 2015, 21, 3532 3542 DOI: 10.1002/ejic.201500195
  24. 24
    Mandimutsira, B. S.; Yamarik, J. L.; Brunold, T. C.; Gu, W.; Cramer, S. P.; Riordan, C. G. J. Am. Chem. Soc. 2001, 123, 9194 9195 DOI: 10.1021/ja016209+

    PhTttBu = phenyltris((tert-butylthio)methyl)borate.

  25. 25
    Itoh, S.; Bandoh, H.; Nakagawa, M.; Nagatomo, S.; Kitagawa, T.; Karlin, K. D.; Fukuzumi, S. J. Am. Chem. Soc. 2001, 123, 11168 11178 DOI: 10.1021/ja0104094
  26. 26
    Kieber-Emmons, M. T.; Schenker, R.; Yap, G. P. A.; Brunold, T. C.; Riordan, C. G. Angew. Chem. 2004, 116, 6884 6886 DOI: 10.1002/ange.200460747

    tmc = 1,4,8,11-tetramethyl-1,4,8,11-tetraazadodecane.

  27. 27
    Rettenmeier, C. A.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2015, 54, 4880 4884 DOI: 10.1002/anie.201500141

    iso-pmbox = bis(oxazolinylmethylidene)pyrrolidine.

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 51 publications.

  1. Avery LeComte, Rachel Sailer, Samyadeb Mahato, Warren VandeVen, Wen Zhou, Alisa R. Paterson, Morgane Desmau, Amani M. Ebrahim, Fabrice Thomas, Linus Chiang. Synthesis and Characterization of Co Complexes Coordinated by a Tetraanionic Bis(amidateanilido) Ligand. Inorganic Chemistry 2025, 64 (12) , 5986-5995. https://doi.org/10.1021/acs.inorgchem.4c05005
  2. Deepika G. Karmalkar, Hyeongtaek Lim, Mahesh Sundararajan, Yong-Min Lee, Mi Sook Seo, Dae Young Bae, Xiaoyan Lu, Britt Hedman, Keith O. Hodgson, Won-Suk Kim, Eunsung Lee, Edward I. Solomon, Shunichi Fukuzumi, Wonwoo Nam. Synthesis, Structure, and Redox Reactivity of Ni Complexes Bearing a Redox and Acid–Base Non-innocent Ligand with NiII, NiIII, and NiIV Formal Oxidation States. Journal of the American Chemical Society 2025, 147 (5) , 3981-3993. https://doi.org/10.1021/jacs.4c11751
  3. Bach X. Nguyen, Warren VandeVen, Gregory A. MacNeil, Wen Zhou, Alisa R. Paterson, Charles J. Walsby, Linus Chiang. High-Valent Ni and Cu Complexes of a Tetraanionic Bis(amidateanilido) Ligand. Inorganic Chemistry 2023, 62 (37) , 15180-15194. https://doi.org/10.1021/acs.inorgchem.3c02358
  4. Jibo Zhang, Jay R. Winkler, Harry B. Gray, Bryan M. Hunter. Mechanism of Nickel–Iron Water Oxidation Electrocatalysts. Energy & Fuels 2021, 35 (23) , 19164-19169. https://doi.org/10.1021/acs.energyfuels.1c02674
  5. Ivan S. Golovanov, Roman S. Malykhin, Vladislav K. Lesnikov, Yulia V. Nelyubina, Valentin V. Novikov, Kirill V. Frolov, Andrey I. Stadnichenko, Evgeny V. Tretyakov, Sema L. Ioffe, Alexey Yu. Sukhorukov. Revealing the Structure of Transition Metal Complexes of Formaldoxime. Inorganic Chemistry 2021, 60 (8) , 5523-5537. https://doi.org/10.1021/acs.inorgchem.0c03362
  6. Norman Zhao, Alexander S. Filatov, Jiaze Xie, Ethan A. Hill, Andrey Yu. Rogachev, John S. Anderson. Generation and Reactivity of a NiIII2(μ-1,2-peroxo) Complex. Journal of the American Chemical Society 2020, 142 (52) , 21634-21639. https://doi.org/10.1021/jacs.0c10958
  7. Franck Le Vaillant, Edward J. Reijerse, Markus Leutzsch, Josep Cornella. Dialkyl Ether Formation at High-Valent Nickel. Journal of the American Chemical Society 2020, 142 (46) , 19540-19550. https://doi.org/10.1021/jacs.0c07381
  8. Jorn D. Steen, Stepan Stepanovic, Mahsa Parvizian, Johannes W. de Boer, Ronald Hage, Juan Chen, Marcel Swart, Maja Gruden, Wesley R. Browne. Lewis versus Brønsted Acid Activation of a Mn(IV) Catalyst for Alkene Oxidation. Inorganic Chemistry 2019, 58 (21) , 14924-14930. https://doi.org/10.1021/acs.inorgchem.9b02737
  9. Juan Chen, Apparao Draksharapu, Davide Angelone, Duenpen Unjaroen, Sandeep K. Padamati, Ronald Hage, Marcel Swart, Carole Duboc, Wesley R. Browne. H2O2 Oxidation by FeIII–OOH Intermediates and Its Effect on Catalytic Efficiency. ACS Catalysis 2018, 8 (10) , 9665-9674. https://doi.org/10.1021/acscatal.8b02326
  10. Peng-Cheng Duan, Dennis-Helmut Manz, Sebastian Dechert, Serhiy Demeshko, Franc Meyer. Reductive O2 Binding at a Dihydride Complex Leading to Redox Interconvertible μ-1,2-Peroxo and μ-1,2-Superoxo Dinickel(II) Intermediates. Journal of the American Chemical Society 2018, 140 (14) , 4929-4939. https://doi.org/10.1021/jacs.8b01468
  11. Maria Letizia Merlini, George J. P. Britovsek, Marcel Swart, Paola Belanzoni. Understanding the Catalase-Like Activity of a Bioinspired Manganese(II) Complex with a Pentadentate NSNSN Ligand Framework. A Computational Insight into the Mechanism. ACS Catalysis 2018, 8 (4) , 2944-2958. https://doi.org/10.1021/acscatal.7b03559
  12. Nicole M. Camasso, Allan J. Canty, Alireza Ariafard, and Melanie S. Sanford . Experimental and Computational Studies of High-Valent Nickel and Palladium Complexes. Organometallics 2017, 36 (22) , 4382-4393. https://doi.org/10.1021/acs.organomet.7b00613
  13. Charvi Singhvi, Gunjan Sharma, Rishi Verma, Vinod K. Paidi, Pieter Glatzel, Paul Paciok, Vashishtha B. Patel, Ojus Mohan, Vivek Polshettiwar. Tuning the electronic structure and SMSI by integrating trimetallic sites with defective ceria for the CO 2 reduction reaction. Proceedings of the National Academy of Sciences 2025, 122 (3) https://doi.org/10.1073/pnas.2411406122
  14. Thomas M. Khazanov, Anusree Mukherjee. Harnessing Oxidizing Potential of Nickel for Sustainable Hydrocarbon Functionalization. Molecules 2024, 29 (21) , 5188. https://doi.org/10.3390/molecules29215188
  15. Ayushi Awasthi, Sharath Chandra Mallojjala, Rakesh Kumar, Raju Eerlapally, Jennifer S. Hirschi, Apparao Draksharapu. Altering the Localization of an Unpaired Spin in a Formal Ni(V) Species. Chemistry – A European Journal 2024, 30 (4) https://doi.org/10.1002/chem.202302824
  16. Eduardo Jaimes–Romano, Hugo Valdés, Simon Hernández–Ortega, Rosa Mollfulleda, Marcel Swart, David Morales–Morales. C–S couplings catalyzed by Ni(II) complexes of the type [(NHC)Ni(Cp)(Br)]. Journal of Catalysis 2023, 426 , 247-256. https://doi.org/10.1016/j.jcat.2023.07.001
  17. Philipp Heim, Giuseppe Spedalotto, Marta Lovisari, Robert Gericke, John O'Brien, Erik R. Farquhar, Aidan R. McDonald. Synthesis and Characterization of a Masked Terminal Nickel‐Oxide Complex. Chemistry – A European Journal 2023, 29 (21) https://doi.org/10.1002/chem.202203840
  18. Priya Sahni, Rahat Gupta, Simran Sharma, Amlan K. Pal. The once-elusive Ni(IV) species is now a potent candidate for challenging organic transformations. Coordination Chemistry Reviews 2023, 474 , 214849. https://doi.org/10.1016/j.ccr.2022.214849
  19. Bing-Feng Qian, Yang Gao, Qian-Ya Xu, Ai-Quan Jia, Hua-Tian Shi, Qian-Feng Zhang. Syntheses, characterizations and structures of ruthenium carbene and allenylidene complexes supported by 1,4,7-trimethyl-1,4,7-triazacyclononane (Me3tacn) ligands. Journal of Organometallic Chemistry 2022, 976 , 122429. https://doi.org/10.1016/j.jorganchem.2022.122429
  20. Jayanta Bag, Kuntal Pal. The access of {NiIV(OH)2} intermediate in Ni(II) mediated oxygen atom transfer to coordinated Phosphine: Combined experimental and computational studies. Polyhedron 2022, 224 , 116030. https://doi.org/10.1016/j.poly.2022.116030
  21. Denis V. Chachkov, Oleg V. Mikhailov. Nickel macrocyclic complexes with porphyrazine and some [benzo]substituted, oxo and fluoro ligands: DFT analysis. Journal of Porphyrins and Phthalocyanines 2022, 26 (03) , 222-231. https://doi.org/10.1142/S1088424622500067
  22. Evan P. Jahrman, Jamie L. Weaver, Niranjan Govind, Marko Perestjuk, Gerald T. Seidler. Iron redox analysis of silicate-based minerals and glasses using synchrotron X-ray absorption and laboratory X-ray emission spectroscopy. Journal of Non-Crystalline Solids 2022, 577 , 121326. https://doi.org/10.1016/j.jnoncrysol.2021.121326
  23. Mian Guo, Yong-Min Lee, Shunichi Fukuzumi, Wonwoo Nam. Biomimetic metal-oxidant adducts as active oxidants in oxidation reactions. Coordination Chemistry Reviews 2021, 435 , 213807. https://doi.org/10.1016/j.ccr.2021.213807
  24. Olga V. Safonova, Maarten Nachtegaal. Operando X ‐Ray Spectroscopies on Catalysts in Action. 2021, 339-361. https://doi.org/10.1002/9783527813599.ch19
  25. Soohyung Kim, Ha Young Jeong, Seonghan Kim, Hongsik Kim, Sojeong Lee, Jaeheung Cho, Cheal Kim, Dongwhan Lee. Proton Switch in the Secondary Coordination Sphere to Control Catalytic Events at the Metal Center: Biomimetic Oxo Transfer Chemistry of Nickel Amidate Complex. Chemistry – A European Journal 2021, 27 (14) , 4700-4708. https://doi.org/10.1002/chem.202005183
  26. Bing-Feng Qian, Jun-Ling Wang, Ai-Quan Jia, Hua-Tian Shi, Qian-Feng Zhang. Syntheses, reactivity, structures and photocatalytic properties of mononuclear ruthenium(II) complexes supported by 1,4,7-trimethyl-1,4,7-triazacyclononane (Me3tacn) ligands. Inorganica Chimica Acta 2021, 516 , 120128. https://doi.org/10.1016/j.ica.2020.120128
  27. Neil Heberer, Chi-Herng Hu, Liviu M. Mirica. High-Valent Ni Coordination Compounds. 2021, 348-374. https://doi.org/10.1016/B978-0-08-102688-5.00104-5
  28. Joeri Hessels, Eduard Masferrer‐Rius, Fengshou Yu, Remko J. Detz, Robertus J. M. Klein Gebbink, Joost N. H. Reek. Nickel is a Different Pickle: Trends in Water Oxidation Catalysis for Molecular Nickel Complexes. ChemSusChem 2020, 13 (24) , 6629-6634. https://doi.org/10.1002/cssc.202002164
  29. Patric Zimmermann, Sergey Peredkov, Paula Macarena Abdala, Serena DeBeer, Moniek Tromp, Christoph Müller, Jeroen A. van Bokhoven. Modern X-ray spectroscopy: XAS and XES in the laboratory. Coordination Chemistry Reviews 2020, 423 , 213466. https://doi.org/10.1016/j.ccr.2020.213466
  30. Lorenzo D'Amore, Leonardo Belpassi, Johannes E. M. N. Klein, Marcel Swart. Spin-resolved charge displacement analysis as an intuitive tool for the evaluation of cPCET and HAT scenarios. Chemical Communications 2020, 56 (81) , 12146-12149. https://doi.org/10.1039/D0CC04995F
  31. Cai-Xia Zhang, Duo-Wen Fang, Jun-Ling Wang, Ai-Quan Jia, Qian-Feng Zhang. Syntheses, characterizations, and reactivities of new 1,4,7-trimethyl-l,4,7-triazacyclononane (Me3tacn) molybdenum and tungsten complexes. Inorganica Chimica Acta 2020, 507 , 119599. https://doi.org/10.1016/j.ica.2020.119599
  32. Filip Vlahovic, Maja Gruden, Stepan Stepanovic, Marcel Swart. Density functional approximations for consistent spin and oxidation states of oxoiron complexes. International Journal of Quantum Chemistry 2020, 120 (5) https://doi.org/10.1002/qua.26121
  33. Noel Nebra. High-Valent NiIII and NiIV Species Relevant to C–C and C–Heteroatom Cross-Coupling Reactions: State of the Art. Molecules 2020, 25 (5) , 1141. https://doi.org/10.3390/molecules25051141
  34. Liviu M. Mirica, Sofia M. Smith, Leonel Griego. Organometallic Chemistry of High‐Valent Ni ( III ) and Ni ( IV ) Complexes. 2020, 223-248. https://doi.org/10.1002/9783527813827.ch10
  35. Marcel Swart. Dealing with Spin States in Computational Organometallic Catalysis. 2020, 191-226. https://doi.org/10.1007/3418_2020_49
  36. Muniyandi Sankaralingam, Mani Balamurugan, Mallayan Palaniandavar. Alkane and alkene oxidation reactions catalyzed by nickel(II) complexes: Effect of ligand factors. Coordination Chemistry Reviews 2020, 403 , 213085. https://doi.org/10.1016/j.ccr.2019.213085
  37. Sofia M. Smith, Oriol Planas, Laura Gómez, Nigam P. Rath, Xavi Ribas, Liviu M. Mirica. Aerobic C–C and C–O bond formation reactions mediated by high-valent nickel species. Chemical Science 2019, 10 (44) , 10366-10372. https://doi.org/10.1039/C9SC03758F
  38. Giuseppe Spedalotto, Robert Gericke, Marta Lovisari, Erik R. Farquhar, Brendan Twamley, Aidan R. McDonald. Preparation and Characterisation of a Bis‐μ‐Hydroxo‐Ni III 2 Complex. Chemistry – A European Journal 2019, 25 (51) , 11983-11990. https://doi.org/10.1002/chem.201902812
  39. Xiaojun Yang, Hang Bai, Marie Moriko Qian, Changjiang Wu, Shuo Cao, Zhiguo Yan, Qifeng Tian, Zhiping Du. Enhanced activity of Pd–Cu(Ni) bimetallic catalysts for oxidative carbonylation of phenol. Materials Chemistry and Physics 2019, 234 , 48-54. https://doi.org/10.1016/j.matchemphys.2019.05.049
  40. José G. Moya‐Cancino, Ari‐Pekka Honkanen, Ad M. J. van der Eerden, Herrick Schaink, Lieven Folkertsma, Mahnaz Ghiasi, Alessandro Longo, Florian Meirer, Frank M. F. de Groot, Simo Huotari, Bert M. Weckhuysen. Elucidating the K‐Edge X‐Ray Absorption Near‐Edge Structure of Cobalt Carbide. ChemCatChem 2019, 11 (13) , 3042-3045. https://doi.org/10.1002/cctc.201900434
  41. Miquel Costas. Alkane Oxidation with Biologically Inspired Nonheme Iron Catalysts Based in the Triazacyclononane Ligand Scaffold. 2019, 251-268. https://doi.org/10.1002/9781119379256.ch13
  42. Sarvesh S. Harmalkar, Dattaprasad D. Narulkar, Raymond J. Butcher, Mahesh S. Deshmukh, Anant Kumar Srivastava, Mariappan Mariappan, Prem Lama, Sunder N. Dhuri. Dual-site aqua mononuclear nickel(II) complexes of non-heme tetradentate ligands: Synthesis, characterization and reactivity. Inorganica Chimica Acta 2019, 486 , 425-434. https://doi.org/10.1016/j.ica.2018.10.069
  43. Evan P. Jahrman, William M. Holden, Alexander S. Ditter, Devon R. Mortensen, Gerald T. Seidler, Timothy T. Fister, Stosh A. Kozimor, Louis F. J. Piper, Jatinkumar Rana, Neil C. Hyatt, Martin C. Stennett. An improved laboratory-based x-ray absorption fine structure and x-ray emission spectrometer for analytical applications in materials chemistry research. Review of Scientific Instruments 2019, 90 (2) https://doi.org/10.1063/1.5049383
  44. Maja Gruden, Wesley R. Browne, Marcel Swart, Carole Duboc. Computational Versus Experimental Spectroscopy for Transition Metals. 2019, 161-183. https://doi.org/10.1007/978-3-030-11714-6_6
  45. Evan P. Jahrman, Lisa A. Pellerin, Alexander S. Ditter, Liam R. Bradshaw, Timothy T. Fister, Bryant J. Polzin, Steven E. Trask, Alison R. Dunlop, Gerald T. Seidler. Laboratory-Based X-ray Absorption Spectroscopy on a Working Pouch Cell Battery at Industrially-Relevant Charging Rates. Journal of The Electrochemical Society 2019, 166 (12) , A2549-A2555. https://doi.org/10.1149/2.0721912jes
  46. Bhawana Pandey, Kallol Ray, Gopalan Rajaraman. Structure, Bonding, Reactivity and Spectral Features of Putative Ni III =O Species: A Theoretical Perspective. Zeitschrift für anorganische und allgemeine Chemie 2018, 644 (14) , 790-800. https://doi.org/10.1002/zaac.201800122
  47. Hitoshi Abe, Giuliana Aquilanti, Roberto Boada, Bruce Bunker, Pieter Glatzel, Maarten Nachtegaal, Sakura Pascarelli. Improving the quality of XAFS data. Journal of Synchrotron Radiation 2018, 25 (4) , 972-980. https://doi.org/10.1107/S1600577518006021
  48. Muniyandi Sankaralingam, Yong-Min Lee, Wonwoo Nam, Shunichi Fukuzumi. Amphoteric reactivity of metal–oxygen complexes in oxidation reactions. Coordination Chemistry Reviews 2018, 365 , 41-59. https://doi.org/10.1016/j.ccr.2018.03.003
  49. Masahiro Kouno, Nobuto Yoshinari, Naoto Kuwamura, Kohei Yamagami, Akira Sekiyama, Mitsutaka Okumura, Takumi Konno. Valence Interconversion of Octahedral Nickel(II/III/IV) Centers. Angewandte Chemie 2017, 129 (44) , 13950-13954. https://doi.org/10.1002/ange.201708169
  50. Masahiro Kouno, Nobuto Yoshinari, Naoto Kuwamura, Kohei Yamagami, Akira Sekiyama, Mitsutaka Okumura, Takumi Konno. Valence Interconversion of Octahedral Nickel(II/III/IV) Centers. Angewandte Chemie International Edition 2017, 56 (44) , 13762-13766. https://doi.org/10.1002/anie.201708169
  51. Susanta Hazra, Nuno M. R. Martins, Maxim L. Kuznetsov, M. Fátima C. Guedes da Silva, Armando J. L. Pombeiro. Flexibility and lability of a phenyl ligand in hetero-organometallic 3d metal–Sn( iv ) compounds and their catalytic activity in Baeyer–Villiger oxidation of cyclohexanone. Dalton Transactions 2017, 46 (39) , 13364-13375. https://doi.org/10.1039/C7DT02534C

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2017, 139, 25, 8718–8724
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.7b04158
Published June 5, 2017

Copyright © 2017 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Article Views

8114

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Formation of 3 from 1 and NaOCl

    Figure 1

    Figure 1. Spin-density plot (S12g/TZ2P) for (antiferromagnetically coupled) 1 with spin-up spin-density shown in blue (around Nia, left) and spin-down spin-density in red (around Nib, right).

    Figure 2

    Figure 2. UV–vis absorption spectrum of 1 (3.5 mM) in methanol before (black) and after (red 6 s, blue 48 s, green 463 s) addition of 11 equiv of NaOCl(aq) at 293 K. Inset: Expansion of the NIR region. The data indicate that for 3 ε612 nm > 715 M–1·cm–1.

    Figure 3

    Figure 3. (Top) Ni K XANES before (blue) and after (yellow) addition of 4 equiv of NaOCl to 1 (30 mM) in acetonitrile. (Bottom) Ni K edge XANES for structures 1, 3, and 3a, simulated using FEFF9.0 (18) using coordinates available from DFT-optimized structures (see the SI).

    Scheme 2

    Scheme 2. Formation of 3 from 1 and NaOCl and Subsequent Decay to 4 in Acetonitrile

    Scheme 3

    Scheme 3. Structures (3, 3a, 3b) Consistent with ESI Mass Spectral Data and Calculated Driving Forces for Their Formation from 1

    Figure 4

    Figure 4. Raman spectra (λexc 532 nm) of 1 (3.5 mM) in acetonitrile (a) and with (b) 4.5 equiv of NaOBr, (c) 4.5 equiv of Na16OCl, and (d) 4.5 equiv of Na18OCl. *Solvent band. #Raman band from quartz.

    Figure 5

    Figure 5. Top: Calculated spectra for 3 with various degrees of 18O substitution (16O3 (blue), 16O218O (red), 16O18O2 (green), 18O3 (purple)). Bottom: Resonance Raman spectra of 3 generated from 1 (4 mM) in methanol by addition of 4 equiv of NaOCl/H2O with (blue) 100% 16O, (red) 50% 18O, (green) 34% 18O, and (purple) 26% 18O.

  • References


    This article references 27 other publications.

    1. 1
      (a) Costas, M.; Mehn, M. P.; Jensen, M. P.; Que, L., Jr. Chem. Rev. 2004, 104, 939 986 DOI: 10.1021/cr020628n
      (b) Solomon, E. I.; Brunold, T. C.; Davis, M. I.; Kemsley, J. N.; Lee, S.-K.; Lehnert, N.; Neese, F.; Skulan, A. J.; Yang, Y.-S.; Zhou, J. Chem. Rev. 2000, 100, 235 350 DOI: 10.1021/cr9900275
      (c) Punniyamurthy, T.; Velusamy, S.; Iqbal, J. Chem. Rev. 2005, 105, 2329 2363 DOI: 10.1021/cr050523v
    2. 2
      (a) Decker, A.; Solomon, E. I. Curr. Opin. Chem. Biol. 2005, 9, 152 163 DOI: 10.1016/j.cbpa.2005.02.012
      (b) Bertini, I.; Gray, H. B.; Stiefel, E. I.; Valentine, J. S. Biological.Inorganic Chemistry. Structure & Reactivity; University Science Books: Sausalito, CA, 2007.
      (c) Meunier, B. Chem. Rev. 1992, 92, 1411 1456 DOI: 10.1021/cr00014a008
      (d) Nam, W. Acc. Chem. Res. 2015, 48, 2415 2423 DOI: 10.1021/acs.accounts.5b00218
      (e) Oloo, W. N.; Que, L. Acc. Chem. Res. 2015, 48, 2612 2621 DOI: 10.1021/acs.accounts.5b00053
    3. 3
      Shearer, J. Acc. Chem. Res. 2014, 47, 2332 2341 DOI: 10.1021/ar500060s
    4. 4
      (a) Cho, J.; Kang, Y.; Liu, L. V.; Sarangi, R.; Solomon, E. I.; Nam, W. Chem. Sci. 2013, 4, 1502 1508 DOI: 10.1039/c3sc22173c
      (b) Honda, K.; Cho, J.; Matsumoto, T.; Roh, J.; Furutachi, H.; Tosha, T.; Kubo, M.; Fujinami, S.; Ogura, T.; Kitagawa, T.; Suzuki, M. Angew. Chem., Int. Ed. 2009, 48, 3304 3307 DOI: 10.1002/anie.200900222
      (c) Tano, T.; Doi, Y.; Inosako, M.; Kunishita, A.; Kubo, M.; Ishimaru, H.; Ogura, T.; Sugimoto, H.; Itoh, S. Bull. Chem. Soc. Jpn. 2010, 83, 530 538 DOI: 10.1246/bcsj.20090346
      (d) Kunishita, A.; Doi, Y.; Kubo, M.; Ogura, T.; Sugimoto, H.; Itoh, S. Inorg. Chem. 2009, 48, 4997 5004 DOI: 10.1021/ic900059m
      (e) Morimoto, Y.; Bunno, S.; Fujieda, N.; Sugimoto, H.; Itoh, S. J. Am. Chem. Soc. 2015, 137, 5867 5870 DOI: 10.1021/jacs.5b01814
    5. 5
      (a) Corona, T.; Pfaff, F. F.; Acua-Pares, F.; Draksharapu, A.; Whiteoak, C. J.; Martin-Diaconescu, V.; Lloret-Fillol, J.; Browne, W. R.; Ray, K.; Company, A. Chem. - Eur. J. 2015, 21, 15029 15038 DOI: 10.1002/chem.201501841
      (b) Pfaff, F. F.; Heims, F.; Kundu, S.; Mebs, S.; Ray, K. Chem. Commun. 2012, 48, 3730 3732 DOI: 10.1039/c2cc30716b
    6. 6
      (a) Draksharapu, A.; Codolá, Z.; Gómez, L.; Lloret-Fillol, J.; Browne, W. R.; Costas, M. Inorg. Chem. 2015, 54, 10656 10666 DOI: 10.1021/acs.inorgchem.5b01463
      (b) Pirovano, P.; Farquhar, E. R.; Swart, M.; McDonald, A. R. J. Am. Chem. Soc. 2016, 138, 14362 14370 DOI: 10.1021/jacs.6b08406
      (c) Corona, T.; Draksharapu, A.; Padamati, S. K.; Gamba, I.; Martin-Diaconescu, V.; Acuña-Parés, F.; Browne, W. R.; Company, A. J. Am. Chem. Soc. 2016, 138, 12987 12996 DOI: 10.1021/jacs.6b07544
    7. 7
      Corona, T.; Company, A. Chem. - Eur. J. 2016, 22, 13422 13429 DOI: 10.1002/chem.201602414
    8. 8
      (a) Camasso, N. M.; Sanford, M. S. Science 2015, 347, 1 7 DOI: 10.1126/science.aaa4526
      (b) Riordan, C. G. Science 2015, 347, 1203 DOI: 10.1126/science.aaa7553
    9. 9
      (a) Gray, H. B.; Hare, C. R. Inorg. Chem. 1962, 1, 363 368 DOI: 10.1021/ic50002a034
      (b) O’Halloran, K. P.; Zhao, C.; Ando, N. S.; Schultz, A. J.; Koetzle, T. F.; Piccoli, P. M. B.; Hedman, B.; Hodgson, K. O.; Bobyr, E.; Kirk, M. L.; Knottenbelt, S.; Depperman, E. C.; Stein, B.; Anderson, T. M.; Cao, R.; Geletii, Y. V.; Hardcastle, K. I.; Musaev, D. G.; Neiwert, W. A.; Fang, X.; Morokuma, K.; Wu, S.; Kögerler, P.; Hill, C. L. Inorg. Chem. 2012, 51, 7025 7031 DOI: 10.1021/ic2008914
    10. 10
      Wieghardt, K.; Schmidt, W.; Herrmann, W.; Kueppers, H. J. Inorg. Chem. 1983, 22, 2953 2956 DOI: 10.1021/ic00162a037
    11. 11
      Bossek, U.; Nühlen, D.; Bill, E.; Glaser, T.; Krebs, C.; Weyhermüller, T.; Wieghardt, K.; Lengen, M.; Trautwein, A. X. Inorg. Chem. 1997, 36, 2834 2843 DOI: 10.1021/ic970119h
    12. 12
      (a) Klamt, A.; Schüürmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 799 805 DOI: 10.1039/P29930000799
      (b) Pye, C. C.; Ziegler, T. Theor. Chem. Acc. 1999, 101, 396 408 DOI: 10.1007/s002140050457
      (c) Swart, M.; Rösler, E.; Bickelhaupt, F. M. Eur. J. Inorg. Chem. 2007, 2007, 3646 3654 DOI: 10.1002/ejic.200700228
    13. 13
      van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597 4610 DOI: 10.1063/1.466059
    14. 14
      (a) Swart, M. Chem. Phys. Lett. 2013, 580, 166 171 DOI: 10.1016/j.cplett.2013.06.045
      (b) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098 3100 DOI: 10.1103/PhysRevA.38.3098
      (c) Perdew, J. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33, 8822 8824 DOI: 10.1103/PhysRevB.33.8822
    15. 15

      At S12g/TZ2P//BP86-D3/TDZP.

    16. 16
      Tak, H.; Lee, H.; Kang, J.; Cho, J. Inorg. Chem. Front. 2016, 3, 157 163 DOI: 10.1039/C5QI00206K
    17. 17
      (a) Sun, Y.-K; Kim, M. G.; Kang, S. H.; Amine, K. J. Mater. Chem. 2003, 13, 319 322 DOI: 10.1039/b209379k
      (b) Gu, W.; Wang, H.; Wang, K. Dalton Trans. 2014, 43, 6406 6413 DOI: 10.1039/c4dt00308j
    18. 18
      Rehr, J. J.; Kas, J. J.; Vila, F. D.; Prange, M. P.; Jorissen, K. Phys. Chem. Chem. Phys. 2010, 12, 5503 5513 DOI: 10.1039/b926434e
    19. 19

      In the absence of Ni(II) or 1, NaOCl is stable in CH3CN.

    20. 20

      This complex is reminiscent of the (μ-η22-disulfido)dinickel(II) complexes with 6-methyl-TPA ligands reported by Itoh et al.; see:

      Inosako, M.; Kunishita, A.; Kubo, M.; Ogura, T.; Sugimoto, T.; Itoh, S. Dalton Trans. 2009, 43, 9410 9147 DOI: 10.1039/b910237j
    21. 21

      The formation of species such as (O═Ni(IV)–O–Ni(IV)═O) is highly unlikely, as it is calculated to lie 45 to 58 kcal mol–1 (depending on functional used, e.g., BP86-D3 and S12g) higher in energy than even the peroxy species 3a. Furthermore, DFT calculations indicate a high tendency for a species to converge to the structure of the 3a species. In addition ligand hydroxylation can be discounted on the basis of the recovery of the Ni(II) complexes, e.g., 1 and 4, once 3 has reacted with solvent or substrate.

    22. 22
      Wieghardt, K.; Bossek, U.; Nuber, B.; Weiss, J.; Bonvoisin, J.; Corbella, M.; Vitols, S. E.; Girerd, J. J. J. Am. Chem. Soc. 1988, 110, 7398 7411 DOI: 10.1021/ja00230a021
    23. 23
      (a) Hage, R.; Krijnen, B.; Warnaar, J. B.; Hartl, F.; Stufkens, D. J.; Snoeck, T. L. Inorg. Chem. 1995, 34, 4973 4978 DOI: 10.1021/ic00124a010
      (b) Angelone, D.; Abdolahzadeh, S.; de Boer, J. W.; Browne, W. R. Eur. J. Inorg. Chem. 2015, 21, 3532 3542 DOI: 10.1002/ejic.201500195
    24. 24
      Mandimutsira, B. S.; Yamarik, J. L.; Brunold, T. C.; Gu, W.; Cramer, S. P.; Riordan, C. G. J. Am. Chem. Soc. 2001, 123, 9194 9195 DOI: 10.1021/ja016209+

      PhTttBu = phenyltris((tert-butylthio)methyl)borate.

    25. 25
      Itoh, S.; Bandoh, H.; Nakagawa, M.; Nagatomo, S.; Kitagawa, T.; Karlin, K. D.; Fukuzumi, S. J. Am. Chem. Soc. 2001, 123, 11168 11178 DOI: 10.1021/ja0104094
    26. 26
      Kieber-Emmons, M. T.; Schenker, R.; Yap, G. P. A.; Brunold, T. C.; Riordan, C. G. Angew. Chem. 2004, 116, 6884 6886 DOI: 10.1002/ange.200460747

      tmc = 1,4,8,11-tetramethyl-1,4,8,11-tetraazadodecane.

    27. 27
      Rettenmeier, C. A.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2015, 54, 4880 4884 DOI: 10.1002/anie.201500141

      iso-pmbox = bis(oxazolinylmethylidene)pyrrolidine.

  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b04158.

    • Animated gifs for selected vibrational modes between 517 and 893 cm−1 of 3 (ZIP)

    • Animated gifs for selected vibrational modes between 536 and 892 cm−1 of 3a (ZIP)

    • Animated gifs for vibrational modes at 653, 689, 700, 742 cm−1 of 3b (ZIP)

    • Details of synthesis and charcaterization of complexes 1, 2, and 4, UV–vis absorption, (resonance) Raman, NMR, EPR spectroscopy, ESI-MS, XANES, XES, and computational data (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.