Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Sep 14.
Published in final edited form as: Dalton Trans. 2021 Aug 10;50(34):11889–11898. doi: 10.1039/d1dt01996a

Formation of cobalt-oxygen intermediates by dioxygen activation at a mononuclear nonheme cobalt(II) center

Deesha D Malik a, Anirban Chandra b, Mi Sook Seo a, Yong-Min Lee a, Erik R Farquhar c, Stefan Mebs d, Holger Dau d, Kallol Ray b, Wonwoo Nam a
PMCID: PMC8499697  NIHMSID: NIHMS1734136  PMID: 34373886

Abstract

A mononuclear nonheme cobalt(II) complex, [(TMG3tren)CoII(OTf)](OTf) (1), activates dioxygen in the presence of hydrogen atom donor substrates, such as tetrahydrofuran and cyclohexene, resulting in the generation of a cobalt(II)-alkylperoxide intermediate (2), which then converts to the previously reported cobalt(IV)-oxo complex, [(TMG3tren)CoIV(O)]2+-(Sc(OTf)3)n] (3), in >90% yield upon addition of a redox-inactive metal ion, Sc(OTf)3. Intermediates 2 and 3 represent the cobalt analogues of the proposed iron(II)-alkylperoxide precursor that converts to an iron(IV)-oxo intermediate via O-O bond heterolysis in pterin-dependent nonheme iron oxygenases. In reactivity studies, 2 shows an amphoteric reactivity in electrophilic and nucleophilic reactions, whereas 3 is an electrophilic oxidant. To the best of our knowledge, the present study reports the first example showing the generation of cobalt-oxygen intermediates by activating dioxygen at a cobalt(II) center and the reactivities of the cobalt-oxygen intermediates in oxidation reaction.

Graphical Abstract

graphic file with name nihms-1734136-f0001.jpg

The present study represents the first example showing the generation of Co(II)-alkylperoxide and Co(IV)-oxo intermediates by employing dioxygen as an oxidant, thereby supporting the proposed involvement of such intermediates in nonheme cobalt complex-mediated alkane hydroxylation and dioxygen reduction reactions.

Introduction

One primary goal in biomimetic research is to understand mechanisms of dioxygen (O2) activation at transition metal centers, structures of reactive intermediates generated in the O2-activation reactions, and reactivities of these intermediates in various oxidation reactions.1,2 Very often, enzymatic O2-activation occurs at an iron(II) active site that leads to a variety of two-electron oxidation processes; a co-substrate then provides the remaining two electrons required for the four-electron reduction of O2.39 In pterin-dependent nonheme iron oxygenases, tetrahydrobiopterin (BH4) is used as a co-substrate that delivers two electrons simultaneously to the active site to form iron(II)-alkylperoxide and iron(IV)-oxo species in the proposed reaction mechanism (Scheme 1A).10,11 In biomimetic studies, a number of nonheme iron(IV)-oxo complexes have been synthesized by activating O2 via mechanisms reminiscent of the O2-activation process proposed in biology.1226 Thus, O2-activation at nonheme iron(II) centers is being unveiled in both enzymatic and biomimetic reactions. However, there is a significant gap in our present understanding in cobalt complex-mediated O2-activation reactions. For example, although generation of mono- and dinuclear Co-superoxo and -peroxo complexes by O2-activation at Co(II) centers has been reported in a number of cases,2732 direct spectroscopic evidence for the generation of terminal Co-O intermediates in the O2-activation reactions has never been obtained previously. Moreover, mechanisms of the formation of Co-oxygen intermediates in the O2-activation by cobalt complexes remain elusive. Regarding the Co-O species, some of us recently reported the stabilization of S = 1/2 and S = 3/2 cobalt(IV)-oxo species in solution by employing iodosylbenzene (PhIO) as an artificial oxidant in the presence of a redox‐inactive metal ion (e.g., Sc3+ ion) or proton;3339 however, others have claimed that one of the proposed cobalt(IV)-oxo species is a cobalt(III)-OH species instead.40 Very recently, a terminal Co(III)-oxo complex was synthesized, isolated, characterized structurally and spectroscopically, and investigated in reactivity studies.4144 Herein, we report the generation of CoII-alkylperoxide species (2) in the reaction of a cobalt(II) complex and O2 in the presence of hydrogen atom (H-atom) donors and the conversion of 2 to a CoIV-O-(Sc3+)n species (3) upon addition of Sc3+ ion (Scheme 1B); thus, the Sc3+ ion triggers an O-O bond heterolysis step of 2 that leads to the generation of a high-valent CoIV-O-(Sc3+)n species by stabilizing the CoIV-O core. Then, the reactivities of 2 and 3 are discussed in both nucleophilic and electrophilic oxidation reactions.

Scheme 1.

Scheme 1

Proposed mechanisms for (a) pterin-dependent nonheme iron oxygenases and (b) formation of intermediates 2 and 3 by activating dioxygen by a Co(II) complex.

Results and discussion

The starting [(TMG3tren)CoII(OTf)](OTf) (1) complex (TMG3tren = tris[2-(N-tetramethylguanidyl)ethyl]amine; OTf = CF3SO3), is air-stable in acetone (see Fig. S1 and Tables S1 and S2 for the crystal structure of [(TMG3tren)CoII](BPh4)2 (1-BPh4)).35 Interestingly, when 1 was exposed to air in the presence of a small amount of tetrahydrofuran (THF) in acetone at 25 oC, a green intermediate 2a with absorption bands at 400 (ε = 1900 M–1 cm–1) and 705 nm (ε = 230 M–1 cm–1) was generated within 20 min (Fig. 1a; see also Fig. S2). The formation rate of 2a was found to depend on the concentration of THF added, and the pseudo-first-order fitting of the kinetic data at 400 nm increased linearly with an increase of the THF concentration, affording a second-order rate constant of 3.7(2) × 10−2 M−1 s−1 at 25 °C (Fig. S3). We also found that 2a was formed much slowly when deuterated THF (THF-d8) was used instead of THF, giving a kinetic isotope effect (KIE) value of 30 (Fig. S3). Such a large KIE value indicates that a hydrogen atom abstraction (HAA) from THF (and THF-d8) by 1 is the rate-determining step (r.d.s.) for the formation of 2a (vide infra).

Fig. 1.

Fig. 1

(a) UV-vis spectral changes for the formation of 2a (blue line) upon addition of THF (0.10 M) to an O2-saturated acetone solution of 1 (0.25 mM; black bold line) at 25 °C. Inset shows the EPR spectrum of 2a (1.0 mM) recorded at 5 K. (b) UV-vis spectral change for the conversion from 2a (blue line) to 3 (red line) upon addition of Sc(OTf)3 (0.50 mM; 2.0 equiv.) to an acetone solution of 2a at 0 °C. Inset shows the EPR spectrum of 3 (1.0 mM) recorded at 5 K.

Cyclohexene can also be used as a H-atom donor instead of THF4547 to generate 2b (Scheme 2) with UV-vis absorption features similar to 2a (Fig. S4a), but the absorption peak wavelengths are slightly different from 2a (Fig. S2 and Fig. S4a). Organic product analysis revealed that cyclohex-2-enol and cyclohex-2-enone were formed as major products (Table S3). The formation rate of 2b in the reaction of 1 and cyclohexene in the presence of O2 was determined to be 7.7(5) × 10–5 M–1 s–1 at 25 °C, which is much slower than the formation rate of 2a with THF (i.e., 3.7(2) × 10−2 M−1 s−1 at 25 °C) (Fig. S4b). Furthermore, a KIE value of 4.0 was determined in the reactions of cyclohexene-h10 and cyclohexene-d10 (Fig. S4b), which is significantly lower than that obtained for THF. Although, the C-H bond dissociation energy of THF4851 is higher than that in cyclohexene, the faster rate of formation of 2a with larger KIE in THF can be plausibly attributed to favourable stereo-electronic interactions between the oxygen lone-pairs and the adjacent C-H bond, as noted previously.4853

Scheme 2.

Scheme 2

Proposed structures of intermediates 2a, 2b, and 2c.

Then, what is the nature of 2a and 2b? All our attempts to obtain the crystal structures of 2a and 2b failed. Therefore, we tried to characterize them spectroscopically. The X-band electron paramagnetic resonance (EPR) spectra of 2a (Fig. 1a inset) and 2b (Fig. S5a) are strikingly similar to the axial S = 3/2 EPR signal of 1, with effective g values of g = 4.3 and gǁ = 2.09,35 showing that Co(II) oxidation states and magnetic centers of 2a and 2b are retained. Notably, previous EPR studies of a series of [(TMG3tren)CoIIX]n+ (X = OTf, Cl, and CH3CN, n = 1 or 2) have also shown that the EPR parameters of [(TMG3tren)CoII]2+ are not sensitive to the nature of the fifth ligand.35 Thus, 1H NMR analysis has been performed for 1 and 2a (Fig. 2). Although the paramagnetic property of 1 and 2a makes the peak assignment difficult, the chemical shifts of the TMG3tren ligand peaks in 2a (Fig. 2b) are found to be very similar to those in 1 (Fig. 2a). This supports the similar magnetic properties of 1 and 2a, as suggested by EPR (vide supra). The 1H NMR spectrum of 2a exhibits additional peaks marked with green asterisks (Fig. 2b), which is possibly attributed to the presence of a different fifth ligand in 2a (-OOR in 2a with R = THF-H, instead of -OTf in 1). The X‐ray absorption near edge spectrum (XANES) of 2a reveals an edge energy of 7718.6 eV (Fig. 3a), which is comparable to the 7718.79 eV that is previously reported for 1 and significantly red-shifted relative to the reported value of 7720.04 eV for 3.35 Extended X-ray absorption fine structure (EXAFS) analysis of 2a shows 5 Co-N/O scatters at 2.04 Å, which is similar to the average Co-N distances of 2.05 Å observed in 1 (Fig. 3b; Table S4). Cold-spray ionization time-of-flight mass spectra (CSI-MS) of 2a and 2b, however, exhibits a prominent ion peak at m/z = 664.4, which was assigned to [CoIV(O)(TMG3tren)(OTf)]+ (calcd m/z = 664.3) (Fig. 4 for 2a and Fig. S6 for 2b). Use of isotopically labelled dioxygen (18O2) for the generation of 2a and 2b showed two-mass unit upshift to m/z = 666.4, indicating that the oxygen atom in 2a and 2b was derived from O2 (Fig. 4 inset for 2a and Fig. S6 inset for 2b). It is suggested that 2a and 2b are unstable under the conditions of CSI-MS, and the alkylhydroperoxo O-O bonds of 2a and 2b are cleaved to form the [CoIV(O)(TMG3tren)(OTf)]+ species.

Fig. 2.

Fig. 2

1H NMR spectra of (a) 1 (4.0 mM), (b) 2a (4.0 mM), and (c) 2c (4.0 mM) in acetone-d6 at 298 K.

Fig. 3.

Fig. 3

(a) Normalized Co K-edge X-ray absorption spectra of 1 (black line), 2a (blue line), and 3 (red line). (b) Observed (black line) and simulated (blue line) Fourier-transformed EXAFS spectra of 2a. The inset shows the observed (black line) and simulated (blue line) EXAFS data on a wave-vector scale before calculation of the Fourier transform.

Fig. 4.

Fig. 4

Positive mode CSI-MS spectrum of 2a produced in the O2-activation reaction by 1 upon addition of THF (0.10 M) to an O2-saturated acetone solution of 1 (0.25 mM) at 25 oC. The peaks at m/z = 664.4 and 648.4 correspond to [CoIV(O)(TMG3tren)(OTf)]+ (calcd m/z = 664.3) and [CoII(TMG3tren)(OTf)]+ (calcd m/z = 648.3), respectively. The insets show the observed isotope distribution patterns for [CoIV(16O)(TMG3tren)(OTf)]+ (m/z = 664.4) originated from 2a-16O (left panel) and [CoIV(18O)(TMG3tren)(OTf)]+ (m/z = 666.4) originated from 2a-18O (right panel).

Based on the +2 oxidation state of cobalt in 2a and 2b and their generation from 1 by H-atom abstraction from THF or cyclohexene in the presence of O2, we propose [(TMG3tren)CoII-OOR]+ (R = THF-H for 2a and CyHex-H for 2b) for 2 (Scheme 2). Consistent with this assignment, 2 can also be generated by reacting 1 with alkyl hydroperoxides. For example, upon addition of cumene hydroperoxide (CumOOH) to the solution of 1 in acetone at 25 oC, a green intermediate, denoted as 2c, was formed with UV-vis absorption bands at λmax = 415 (ε = 4600 M–1 cm–1) and 680 nm (ε = 630 M–1 cm–1) (Fig. S7), which is similar to those of 2a and 2b (Fig. S2 and Fig. S4a). EPR (Fig. S5b) and 1H-NMR (Fig. 2c) spectral features were also found to be similar to 2a and 2b. The CSI-MS analysis of 2c, like 2a and 2b, also showed a prominent mass peak at m/z = 664.3 corresponding to [CoIV(O)(TMG3tren)(OTf)]+ (calcd m/z = 664.3) (Fig. S8). In 2c, an additional peak at m/z = 663.3 corresponding to [CoIII(TMG3tren-O)(OTf)]+ (calcd m/z = 663.3) was also observed, which presumably results from further oxidation of the TMG3tren ligand in the presence of residual CumOOH. When Cum18O18OH was used instead of Cum16O16OH, a two-mass unit shift from m/z = 663.3 and 664.3 to m/z = 665.3 and 666.3, respectively, was observed (Fig. S8 inset). 1H-NMR spectrum of 2c, like 2a, also revealed the additional peaks marked with blue asterisks (Fig. 2c). Finally, the infrared spectra of an acetone solution of 2c exhibited a signal at 865 cm−1; this band, which was absent in 1 and red-shifted in CumOOH, is attributed to ν(O-O) of the CoII-OOCum core in 2c (Fig. 5).

Fig. 5.

Fig. 5

Solution IR spectral changes of 2c (60 mM; blue line) upon addition of cumene hydroperoxide (60 mM) in acetone at 25 °C. Black line shows solution IR spectrum of acetone and red line shows cumene hydroperoxide (60 mM).

Interestingly, addition of Sc(OTf)3 to the solutions of the Co(II)-alkylperoxide intermediates, such as 2a, 2b, and 2c, resulted in the instantaneous formation of the previously reported Co(IV)-oxo species, [(TMG3tren)CoIV(O)]2+-(Sc(OTf)3)n (3),35 in a near-stoichiometric yield at 0 oC, as indicated by its characteristic absorption bands centered at 470 and 900 nm (Fig. 1b; Fig. S9) and the rhombic EPR spectrum (E/D = 0.15 ± 0.01) with g = 5.8 and gǁ = 2.58 (Fig. 1b inset). Since the conversion of 2 to 3 was very fast at 0 oC, we could not monitor the rate of reaction even at low temperature using a conventional UV-vis spectrophotometer; therefore, we turned to stopped-flow methods at –40 oC (Fig. 6). Intermediate 2a with absorption bands at 400 and 705 nm was converted to intermediate 3 having absorption bands at 470 and 900 nm with isosbestic points at 440 and 600 nm within 5 s (kobs = 7.1 × 10–1 s–1) (Fig. 6).35 Furthermore, in the reaction of 2c and Sc(OTf)3, cumene alcohol was detected as a sole product (~100%), suggesting heterolytic O-O bond cleavage of the cumyl peroxide group of 2c to yield 3.

Fig. 6.

Fig. 6

Visible spectral changes for the formation of 3 (red line) upon addition of Sc(OTf)3 (2.0 equiv) to a solution of 2a (0.25 mM; blue line) in acetone at –40 °C. Inset shows time course monitored at 470 nm due to the formation of 3.

EXAFS analysis of 3, which was generated by 2a upon addition of Sc3+ ion, can reproduce the previously reported Co-O distance of 1.85 Å for the CoIV-O-(Sc3+)n core (Table S5 and Fig. S10). DFT calculations predict a very short CoIV-O distance of 1.684 Å for the S = 3/2 [(TMG3tren)CoIV(O)]2+ complex (Fig. S11a). Binding of a single Sc3+ ion to the Co(IV)-oxo core leads to an elongation of the DFT-derived CoIV-O distance to 1.764 Å in [(TMG3tren)CoIV(O)(Sc(OTf)3)]2+ (Fig. S11b), with introduction of an additional Sc3+ ion into the CoIV-O-Sc3+ interaction further elongating the calculated CoIV-O distance to 1.854 Å in [(TMG3tren)CoIV(O)(Sc(OTf)(OH)2)2]2+ (Fig. S11c), which is in excellent agreement with experiment (Table S6). Further, the rhombic EPR spectrum of 3 with g = 5.80 and gǁ = 2.58 is unique for cobalt in a +4 oxidation state, and contrasts to the axial S = 3/2 signal observed for all [(TMG3tren)CoII]2+ complexes irrespective of the presence or absence of the axial ligand. Spin quantification based on the EPR signal accounts for 94% of the total cobalt-spin in the solution of 3 (Fig. S12), thereby showing that any Co(III) species, if formed, are present only in a small quantity. We therefore propose a CoIV-O-(Sc3+)2 assignment of 3; EPR, XAS, and DFT data are not consistent with the suggestion of an alternative [(TMG3tren)CoIII(OH)Sc(OTf)3]2+ assignment.39

A proposed mechanism for the O2-activation by 1 is depicted in Scheme 1B. In the absence of any H-atom donors, binding of O2 to Complex 1 is not favoured. However, presence of H-atom donors initiates a preequilibrium binding of O2 to the high-spin Co(II) center in 1 to form a transient cobalt-dioxygen (Co-O2) intermediate (Scheme 1B, reaction a). Subsequent H-atom abstraction from THF or cyclohexene by Co-O2 results in the generation of cobalt(III)-hydroperoxide and a carbon based radical (Scheme 1B, reaction b),4547 which then recombines to yield a cobalt(II)-alkylperoxide species (2) (Scheme 1B, reaction c). The H-atom abstraction by the presumed Co-O2 species is the rate-determining step (r.d.s.), as evident from the measured large KIE values. The rebound step is also important. For example, in the reaction of 1 with THF in the presence of 5,5-Dimethyl-1-pyrroline N-oxide (DMPO) as a radical scavenger (Fig. S13), a trend of decreasing yield of 2a with increasing concentration of DMPO was observed. Furthermore, no generation of 2 is observed in presence of phenols or dihydroanthracene (Fig. S14). In these cases the necessary recombination of cobalt(III)-hydroperoxide and radical for the formation of 2 is presumably hampered by the formation of the C-C coupling products (in phenols) or the second HAT to form anthracene in case of DHA. In the final step, addition of Sc3+ ion to the solution of 2 affords the CoIV-O-(Sc3+)n core via O-O bond heterolysis of the cobalt(II)-alkylperoxide intermediate (Scheme 1B, reaction d), as has been proposed in the chemistry of a nonheme iron(II)-alkylperoxide species, in which the formation of an iron(IV)-oxo species was observed via a heterolytic O-O bond cleavage of Fe(II)-OOH(R) species.54 Similarly, Que, Company and co-workers have proposed O-O bond heterolysis of nonheme iron(III)-hydroperoxide intermediates in the presence of H+ or Sc3+.5557

Detailed reactivities of the cobalt(II)-alkylperoxide species (2) and cobalt(IV)-oxo-Sc3+ (3) species were investigated in both nucleophilic and electrophilic reactions. First, the nucleophilic character of the Co intermediates was examined in aldehyde deformylation reactions. While complex 3 did not show any reactivity with cyclohexane carboxaldehyde (CCA) (Fig. S15), addition of CCA to 2a in acetone at 25 °C resulted in the decay of the characteristic band at 400 nm with a second-order rate constant (k2) of 2.8(2) × 10−1 M−1 s−1 (Fig. 7a and Fig. S16). Product analysis of the reaction solution revealed the formation of cyclohexene, a deformylated product of CCA,5860 in 30(4)% yield based on the amount of 2a used. The electrospray ionization mass spectrum (ESI-MS) of the reaction solution of 2a showed the formation of [CoII(TMG3tren)]2+, which was further confirmed by EPR as CoII (Fig. S17). The reactivity of 2a was further investigated by using substituted benzaldehydes with a series of electron-donating and -withdrawing substituents at the para-position of the phenyl group (para-X-Ph-CHO; X = Me, H, F, and CN) (Table S7 and Fig. S18). A positive slope (ρ) of 1.1 in the Hammett plot was obtained (Fig. 7b), further demonstrating that complex 2a is a nucleophilic oxidant.55,61 Addition of CCA to intermediates 2b and 2c also showed deformylation reactions with second order rate constants of 1.2(1) × 10−1 M−1 s−1 and 1.6(1) × 10−1 M−1 s−1, respectively (Fig. 7a).

Fig. 7.

Fig. 7

(a) Plots of pseudo-first-order rate constants (kobs) against the concentration of CCA to determine the second-order rate constants (k2) for the CCA deformylation reaction by 2a (black circles), 2b (red circles), and 2c (blue circles) in acetone at 25 °C. (b) Hammett plot of log k2 against σp+ for the reactions of 2a with para-X-benzaldehydes (X = Me, H, F, and CN) in acetone at 25 °C.

The electrophilic character of 2a and 3 was also investigated in hydrogen atom transfer (HAT) reactions with substrates containing weak C-H bonds, such as xanthene, DHA, and CHD, and in oxygen atom transfer (OAT) reactions with PPh3. Complex 3 is a competent oxidant in both HAT and OAT reactions, as other high-valent Co(IV)-oxo complexes are electrophilic oxidants.3338 For example, 3 performed OAT to PPh3 at a rate of three orders of magnitude faster than 2a (Fig. 8; Figs. S19S22). Similarly, 3 is approximately one order of magnitude faster than 2a in HAT reactions. In addition, KIE values of 4.6(3) and 2.3(2) were determined in the reactions of xanthene/xanthene-d2 with 2a and 3, respectively (Fig. 8). Further, the rate constants of their reactions with the C-H substrates correlate linearly with the C-H BDEs of the substrates,47 thereby establishing H-atom abstraction as the r.d.s. for both 2a and 3 (Fig. 9; Table S8; Figs. S23 and S24). 2c showed slow reactions towards HAT and OAT reactions (Fig. S25), compared to that of 2a. Product analysis for the HAT and OAT reactions by 2a and 3 was performed with ESI-MS and EPR for the decay product(s) of the cobalt intermediates, and organic product analysis was carried out with GC and HPLC methods (Table S9 and Figs. S26 and S27). From the reactivity studies, we conclude that both 2 and 3 are electrophilic oxidants and further that 2 is an amphoteric oxidant showing reactivity in both electrophilic and nucleophilic reactions. It is worth noting that other metal-(hydro)alkylperoxide complexes have shown reactivities in both electrophilic and nucleophilic reactions.58,6170

Fig. 8.

Fig. 8

(a, b) Plots of pseudo-first-order rate constants (kobs) against the concentration of substrates, (a) xanthene-(h2 and d2) and (b) PPh3, to determine the second-order rate constants (k2) for the oxidation of xanthene and PPh3 by 2a in acetone at 25 °C. (c, d) Plots of pseudo-first-order rate constants (kobs) against the concentration of substrates, (c) xanthene-(h2 and d2) and (d) PPh3, to determine the second-order rate constants (k2) for the oxidation of xanthene and PPh3 by 3 in acetone at 0 °C.

Fig. 9.

Fig. 9

Plots of log k2’ vs C–H BDE values of substrates for the oxidation of xanthene, DHA and CHD by (a) 2a at 25 °C and (b) 3 at 0 °C in acetone. The k2’ values were obtained by dividing the second-order rate constants (k2) by the numbers of equivalent target C-H bonds in the substrates (i.e., 2 for xanthene and 4 for DHA and CHD).

Conclusions

In summary, the previously reported CoIV-O-(Sc3+)n complex, 3,35 is generated in near-stoichiometric yield by activating dioxygen at a Co(II) complex, 1, via a mechanism reminiscent of the dioxygen activation process observed in biology. In particular, 2 represents the cobalt analogue of the proposed iron(II)-alkylperoxide precursor that converts to an iron(IV)-oxo intermediate by O-O bond heterolysis in pterin-dependent nonheme iron oxygenases. A comparative reactivity study demonstrates that complex 2 is amphoteric in nature in contrast to the predominantly electrophilic property of 3. To conclude, the present study represents the first example showing the generation of Co(II)-alkylperoxide and Co(IV)-oxo intermediates by employing dioxygen as an oxidant, thereby supporting the proposed involvement of such intermediates in nonheme cobalt complex-mediated alkane hydroxylation and dioxygen reduction reactions.

Experimental

Materials

All chemicals, such as scandium(III) trifluoromethanesulfonate (ScIII(OTf)3), cyclohexene, xanthene, triphenylphoshine, benzaldehyde, p-methylbenzaldehyde, p-flurobenzaldehyde, p-cyanobenzaldehyde, cumene hydroperoxide, 9,10-dihydroanthracene, 1,4-cyclohexadiene and 2,4 ditert butylphenol, which were of the best available purity, were purchased from Sigma-Aldrich Chemical Co. and Alfa and TCI Chemicals, and used as received unless noted otherwise. Solvents were dried according to the published procedures and distilled under an argon atmosphere prior to use.71 TMG3tren (= 1,1,1-tris{2-(N-tetramethylguanidyl)ethyl}amine) ligand,7274 xanthene-d2,7577 and the Co(OTf)278 (OTf = CF3SO3) were synthesized according to the previously reported methods. To obtain [(TMG3tren)Co](BPh4)2 (1-BPh4) complex, the anion exchange from OTf to BPh4 was performed by adding 10 equiv. of NaBPh4 to a solution of [(TMG3tren)CoII](OTf)2 (1).

Instrumentation

The conclusions section should come in this section at the end of the article, before the acknowledgements.

UV-vis spectra were recorded on a Hewlett Packard 8453 diode array spectrophotometers equipped with a UNISOKU Scientific Instruments Cryostat USP-203A for low-temperature experiments. Cold spray ionization time-of-flight mass (CSI-MS) spectral data were collected on a JMS-T100CS (JEOL) mass spectrometer equipped with the CSI source [conditions: needle voltage = 2.2 kV, orifice 1 current = 50–500 nA, orifice 1 voltage = 0–20 V, ringlens voltage = 10 V, ion source temperature = 5 oC, spray temperature = –40 °C]. CSI-MS spectral data for 2a, 2b, 2c and 3 were obtained by directly infusing the reaction solution into the ion source through pre-cooled tube under high N2 gas pressure. Electrospray ionization mass (ESI-MS) spectra were collected on a Thermo Finnigan (San Jose, CA, USA) LCQTM Advantage MAX quadrupole ion trap instrument, by infusing samples directly into the source at 20 μL/min with a syringe pump. The spray voltage was set at 3.7 kV and the capillary temperature at 80 °C. X-band CW-EPR spectra were recorded at 5 K using X-band Bruker EMX-plus spectrometer equipped with a dual mode cavity (ER 4116DM) [All experimental parameters as follow: microwave frequency = 9.646 GHz, microwave power = 1.0 mW, modulation amplitude = 10 G, gain = 1 × 104, modulation frequency = 100 kHz, time constant = 40.96 ms, conversion time = 81.00 ms]. Low temperature was achieved and controlled with an Oxford Instruments ESR900 liquid He quartz cryostat with an Oxford Instruments ITC503 temperature and gas flow controller. Product analysis was performed with waters 515 high performance liquid chromatography (HPLC), and Agilent Technologies 6890N gas chromatograph (GC) and Thermo Finnigan (Austin, Texas, USA) FOCUS DSQ (dual stage quadrupole) mass spectrometer interfaced with Finnigan FOCUS gas chromatograph (GC-MS).

X-Ray Structural Analysis

[(TMG3tren)CoII](OTf)2 (1) was synthesized according to the literature procedure.35 To improve the quality of the single crystal of 1, the anion exchange from OTf to BPh4 has been performed by introducing NaBPh4 into a solution containing 1. Single crystals of 1-BPh4 suitable for X-ray crystallographic analysis were obtained by slow diffusion of Et2O into a saturated acetone solution of 1-BPh4 (Fig. S1 for the crystal structure). The crystals were taken from the solution by a nylon loop (Hampton Research Co.) on a handmade cooper plate and mounted on a goniometer head in a N2 cryostream. The diffraction data for 1-BPh4 was collected at 170(2) K, on a Bruker SMART AXS diffractometer equipped with a monochromator in the Mo Kα (λ = 0.71073 Å) incident beam. The CCD data were integrated and scaled using the Bruker-SAINT software package, and the structure was solved and refined using SHEXTL V 6.12.79 Hydrogen atoms were located in the calculated positions. CCDC-2090593 contains the supplementary crystallographic data for 1-BPh4. The crystallographic data and selected bond distances and angles for 1-BPh4 are listed in Tables S1 and S2. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif or from the Cambridge Crystallographic Data Centre, 12, Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223–336-033; or email: deposit@ccdc.cam.ac.uk.

Kinetic Measurements

All reactions were run in a 1-cm UV quartz cuvette followed by monitoring UV-visible spectral changes of reaction solutions. Rate constants were determined under pseudo-first-order conditions (i.e., [substrate]/[Intermediate] > 10) by fitting the changes in absorbance for disappearance of peaks at 400 nm due to 2a and 470 nm due to 3. The substrates, such as xanthene, 9,10-dihydroanthracene (DHA), and 1,4-cyclohexadiene (CHD), were used in the C-H bond activation reactions by 2a and 3 in acetone. The kinetic isotope effect (KIE) values for the oxidation of xanthene-(h2 and d2) by 2a and 3 were determined by comparing the k2 values obtained in the C-H and C-D bond activations of xanthene-h2 and xanthene-d2, respectively. The kinetic experiments were run at least in triplicate, and the data reported represent the average of these reactions. The k2′ values were obtained by dividing second-order rate constants (k2) with the number of equivalent target C-H bonds of substrates. Aldehyde deformylation reactions by 2a, 2b, and 2c were also performed to investigate the nucleophilic characters of 2a, 2b, and 2c.

Product Analysis

Products produced in the oxidation of substrates, such as CHD, DHA, and cyclohexane carboxaldehyde (CCA), by 2a and 3 complexes were analyzed by GC. Product yields were determined by comparing the peak areas of sample products in GC chromatograms against standard curves prepared with known authentic references using decane as an internal standard. Products produced in the oxidation of triphenylphosphine by 2a and 3 were analyzed by HPLC. Product yields were determined by comparing the peak areas of sample products in HPLC chromatograms against standard curves prepared with known authentic reference. The inorganic cobalt products obtained in the oxidation of substrates by cobalt intermediates were also analyzed by EPR and ESI-MS spectroscopies.

X-ray Absorption Spectroscopy (SSRL)

XAS data for 1 and 2a were collected at beamline 2–2 of the Stanford Synchrotron Radiation Lightsource (SLAC National Accelerator Lab, Menlo Park, CA, USA), with the storage ring operating at 3.0 GeV and 500 mA. A Si(111) double crystal monochromator was used for energy selection, and was detuned by ~40% for harmonic rejection. Sample temperatures were maintained at approximately 20 K using a He Displex cryostat. Co metal foil spectra were recorded simultaneously using a photodiode for internal energy calibration, with the first inflection point of the reference foil edge set to 7709.0 eV. XAS data were collected as fluorescence spectra using a 13 element solid state germanium detector (Canberra), with the following parameters: 10 eV steps/1 second integration time in the pre-edge region, 0.3 eV steps /2 second integration time in the edge, and 0.05k steps in the EXAFS, with integration time increasing in a k2-weighted fashion from 2 to 8 seconds over the energy range (kmax = 12.5k). The total detector counts were typically 5–25 kHz, well within the linear range of the detector electronics. Samples were monitored for photoreduction during data collection, however no photoreduction was observed for any sample based on the absence of any scan-to-scan red-shift in the absorption edge. Tandem Mossbauer/XAS cups with a sample window of ~ 6mm x 10 mm were used as sample cells.

Averaging and normalization of the XAS data was performed using Athena,80 a graphical implementation of the IFEFFIT package.81 EXAFS analysis of 2a was carried out using Artemis, which incorporates the IFEFFIT fitting engine and FEFF6 for ab initio EXAFS phase and amplitude parameters. DFT calculated structures were used as FEFF6 input to identify significant paths. For a given shell in all simulations, the coordination number n was fixed, while r and σ2 were allowed to float. The amplitude reduction factor S02 was fixed at 0.9, while the edge shift parameter ΔE0 was allowed to float at a single common value for all shells. The fit was evaluated in k3-weighted R-space, and fit quality was judged by the reported R-factor.

XAS measurements and DFT calculations (BESSY)

XAS measurements of 3 were conducted at the KMC3 beamline of the BESSY synchrotron at the Helmholtz-Zentrum Berlin (HZB). The samples were disposed at ca. −60°C into cylindrical Teflon sample holders with thin walls of roughly 100 µm specifically designed for X-ray spectroscopy. Data collection was performed at 20 K in a liquid-helium cryostat in fluorescence detection mode using a 13 element ultra-low energy resolving Silicon drift detector (SDD) from Canberra. Over 25 spectra were averaged in order to improve the signal-to-noise ratio. Averaged spectra were background-corrected and normalized using in-house software. Subsequently, unfiltered k3‐weighted spectra and phase functions from FEFF8.582 were used for least‐squares curve‐fitting of the EXAFS with in-house software and for calculation of Fourier‐transforms representing k‐values between 2 and 14 Å−1. Data were multiplied by a fractional cosine window (10% at low and high k‐side); the amplitude reduction factor S02 was 0.95. The structural models (also used for phase function computations) were obtained by DFT optimization at the UTPSSh/6–311+G(2df,p)83,84 level of theory (an effective core potential was used for Co85,86 applying Gaussian16.87 The COSMO solvation model was used to mimic the acetonitrile solvation.88 Dispersion was implemented by the empirical disperison correction of Grimme.89

Supplementary Material

Supplementary Material
CIF

Acknowledgements

This work was supported by the NRF of Korea through CRI (NRF-2021R1A3B1076539 to W.N.) and Basic Science Research Program (NRF-2020R1I1A1A01074630 to Y.-M.L. and NRF-2019R1I1A1A01055822 to M.S.S.) and the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy - EXC 2008 – 390540038 - UniSysCat to K.R. and H.D. and the Heisenberg-Professorship to K.R.. XAS measurements at SSRL BL 2–2 were made possible by support from U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract Nos. DE-AC02–76SF00515 and DE-SC0012704, as well as the National Institutes of Health (P30-EB-009998).

Footnotes

Conflicts of interest

The authors declare no conflict of interest.

Notes and References

  • 1.Nam W, Acc. Chem. Res, 2007, 40, 465 and articles in the special issue. [DOI] [PubMed] [Google Scholar]
  • 2.Que L Jr., J. Biol. Inorg. Chem, 2017, 22, 171–173 and articles in the special issue. [DOI] [PubMed] [Google Scholar]
  • 3.Groves JT, Proc. Natl. Acad. Sci. U.S.A, 2003, 100, 3569–3574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Costas M, Mehn MP, Jensen MP and Que L Jr., Chem. Rev, 2004, 104, 939–986. [DOI] [PubMed] [Google Scholar]
  • 5.Ray K, Pfaff FF, Wang B and Nam W, J. Am. Chem. Soc, 2014, 136, 13942–13958. [DOI] [PubMed] [Google Scholar]
  • 6.Solomon EI, Goudarzi S and Sutherlin KD, Biochemistry, 2016, 55, 6363–6374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Huang X and Groves JT, Chem. Rev, 2018, 118, 2491–2553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Guengerich FP, ACC Catal, 2018, 8, 10964–10976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Guo M, Corona T, Ray K and Nam W, ACS Cent. Sci, 2019, 5, 13–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Solomon EI, Wong SD, Liu LV, Decker A and Chow MS, Curr. Opin. Chem. Biol, 2009, 13, 99–113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kal S and Que L Jr., J. Biol. Inorg. Chem, 2017, 22, 339–365. [DOI] [PubMed] [Google Scholar]
  • 12.Nam W, Acc. Chem. Res, 2015, 48, 2415–2423. [DOI] [PubMed] [Google Scholar]
  • 13.Sallmann M and Limberg C, Acc. Chem. Res, 2015, 48, 2734–2743. [DOI] [PubMed] [Google Scholar]
  • 14.Sahu S and Goldberg DP, J. Am. Chem. Soc, 2016, 138, 11410–11428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Engelmann X, Monte-Pérez I and Ray K, Angew. Chem., Int. Ed, 2016, 55, 7632–7649. [DOI] [PubMed] [Google Scholar]
  • 16.Hong S, Lee Y-M, Ray K and Nam W, Coord. Chem. Rev, 2017, 334, 25–42. [Google Scholar]
  • 17.MacBeth CE, Golombek AP, Young VG Jr., Yang C, Kuczera K, Hendrich MP and Borovik AS, Science, 2000, 289, 938–942. [DOI] [PubMed] [Google Scholar]
  • 18.Kim SO, Sastri CV, Seo MS, Kim J and Nam W, J. Am. Chem. Soc, 2005, 127, 4178–4179. [DOI] [PubMed] [Google Scholar]
  • 19.Thibon A, England J, Martinho M, Young VG Jr., Frisch JR, Guillot R, Girerd J-J, Münck E, Que L Jr., and Banse F, Angew. Chem., Int. Ed, 2008, 47, 7064–7067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Hong S, Lee Y-M, Shin W, Fukuzumi S and Nam W, J. Am. Chem. Soc, 2009, 131, 13910–13911. [DOI] [PubMed] [Google Scholar]
  • 21.Badiei YM, Siegler MA and Goldberg DP, J. Am. Chem. Soc, 2011, 133, 1274–1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Sallmann M, Siewert I, Fohlmeister L, Limberg C and Knispel C, Angew. Chem., Int. Ed, 2012, 51, 2234–2237. [DOI] [PubMed] [Google Scholar]
  • 23.Nishida Y, Lee Y-M, Nam W and Fukuzumi S, J. Am. Chem. Soc, 2014, 136, 8042–8049. [DOI] [PubMed] [Google Scholar]
  • 24.Hitomi Y, Arakawa K and Kodera M, Chem. Commun, 2014, 50, 7485–7487. [DOI] [PubMed] [Google Scholar]
  • 25.Sankaralingam M, Lee Y-M, Lu X, Vardhaman AK, Nam W and Fukuzumi S, Chem. Commun, 2017, 53, 8348–8351. [DOI] [PubMed] [Google Scholar]
  • 26.Gordon JB, Vilbert AC, DiMucci IM, MacMillan SN, Lancaster KM, Moënne-Loccoz P and Goldberg DP, J. Am. Chem. Soc, 2019, 141, 17533–17547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Egan JW Jr., Haggerty BS, Rheingold AL, Sendlinger SC and Theopold KH, J. Am. Chem. Soc, 1990, 112, 2445–2446. [Google Scholar]
  • 28.Hikichi S, Akita M and Moro-oka Y, Coord. Chem. Rev, 2000, 198, 61–87. [Google Scholar]
  • 29.Nguyen AI, Hadt RG, Solomon EI and Tilley TD, Chem. Sci, 2014, 5, 2874–2878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wang C-C, Chang H-C, Lai Y-C, Fang H, Li C-C, Hsu H-K, Li Z-Y, Lin T-S, Kuo T-S, Neese F, Ye S, Chiang Y-W, Tsai M-L, Liaw W-F and Lee W-Z, J. Am. Chem. Soc, 2016, 138, 14186–14189. [DOI] [PubMed] [Google Scholar]
  • 31.Corcos AR, Villanueva O, Walroth RC, Sharma SK, Bacsa J, Lancaster KM, MacBeth CE and Berry JF, J. Am. Chem. Soc, 2016, 138, 1796–1799. [DOI] [PubMed] [Google Scholar]
  • 32.Corona T, Padamati SK, Acuña-Parés F, Duboc C, Browne WR and Company A, Chem. Commun, 2017, 53, 11782–11785. [DOI] [PubMed] [Google Scholar]
  • 33.Hong S, Pfaff FF, Kwon E, Wang Y, Seo M-S, Bill E, Ray K and Nam W, Angew. Chem., Int. Ed, 2014, 53, 10403–10407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Hong S, Pfaff FF, Kwon E, Wang Y, Seo M-S, Bill E, Ray K and Nam W, Angew. Chem., Int. Ed, 2017, 56, 10630. [DOI] [PubMed] [Google Scholar]
  • 35.Pfaff FF, Kundu S, Risch M, Pandian S, Heims F, Pryjomska-Ray I, Haack P, Metzinger R, Bill E, Dau H, Comba P and Ray K, Angew. Chem., Int. Ed, 2011, 50, 1711–1715. [DOI] [PubMed] [Google Scholar]
  • 36.Wang B, Lee Y-M, Tcho W-Y, Tussupbayev S, Kim S-T, Kim Y, Seo MS, Cho K-B, Dede Y, Keegan BC, Ogura T, Kim SH, Ohta T, Baik M-H, Ray K, Shearer J and Nam W, Nat Commun, 2017, 8, 14839–14849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Nurdin L, Spasyuk DM, Fairburn L, Piers WE and Maron L, J. Am. Chem. Soc, 2018, 140, 16094–16105. [DOI] [PubMed] [Google Scholar]
  • 38.Fukuzumi S, Lee Y-M and Nam W, Bull. Korean Chem. Soc, 2020, 41, 1217–1232. [Google Scholar]
  • 39.Larson VA, Battistella B, Ray K, Lehnert N and Nam W, Nat. Rev. Chem, 2020, 4, 404–419. [DOI] [PubMed] [Google Scholar]
  • 40.Lacy DC, Park YJ, Ziller JW, Yano J and Borovik AS, J. Am. Chem. Soc, 2012, 134, 17526–17535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Goetz MK, Hill EA, Filatov AS and Anderson JS, J. Am. Chem. Soc, 2018, 140, 13176–13180. [DOI] [PubMed] [Google Scholar]
  • 42.Goetz MK and Anderson JS, J. Am. Chem. Soc, 2019, 141, 4051–4062. [DOI] [PubMed] [Google Scholar]
  • 43.Goetz MK and Anderson JS, J. Am. Chem. Soc, 2020, 142, 5439–5441. [DOI] [PubMed] [Google Scholar]
  • 44.Andris E, Navrátil R, Jašík J, Srnec M, Rodríguez M, Costas M and Roithová J, Angew. Chem., Int. Ed, 2019, 58, 9619–9624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lee Y-M, Hong S, Morimoto Y, Shin W, Fukuzumi S and Nam W, J. Am. Chem. Soc, 2010, 132, 10668–10670. [DOI] [PubMed] [Google Scholar]
  • 46.Hong S, Lee Y-M, Sankaralingam M, Vardhaman AK, Park YJ, Cho K-B, Ogura T, Sarangi R, Fukuzumi S and Nam W, J. Am. Chem. Soc, 2016, 138, 8523–8532. [DOI] [PubMed] [Google Scholar]
  • 47.Luo Y-R, Handbook of Bond Dissociation Energies in Organic Compounds, CRC Press, New York, 2003. [Google Scholar]
  • 48.Guo M, Lee Y-M, Gupta R, Seo MS, Ohta T, Wang H-H, Liu H-Y, Dhuri SN, Sarangi R, Fukuzumi S and Nam W, J. Am. Chem. Soc, 2017, 139, 15858–15867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Xue X-S, Ji P, Zhou B and Cheng J-P, Chem. Rev, 2017, 117, 8622–8648. [DOI] [PubMed] [Google Scholar]
  • 50.Finn M, Friedline R, Suleman NK, Wohl CJ and Tanko JM, J. Am. Chem. Soc, 2004, 126, 7578–7584. [DOI] [PubMed] [Google Scholar]
  • 51.Matsumoto T, Ohkubo K, Honda K, Yazawa A, Furutachi H, Fujinami S, Fukuzumi S and Suzuki M, J. Am. Chem. Soc, 2009, 131, 9258–9267. [DOI] [PubMed] [Google Scholar]
  • 52.Malatesta V and Scaiano JC, J. Org. Chem, 1982, 47, 1455–1459. [Google Scholar]
  • 53.Malatesta V and Ingold KU, J. Am. Chem. Soc, 1981, 103, 609–614. [Google Scholar]
  • 54.Bang S, Park S, Lee Y-M, Hong S, Cho K-B and Nam W, Angew. Chem., Int. Ed, 2014, 53, 7843–7847. [DOI] [PubMed] [Google Scholar]
  • 55.Kal S, Draksharapu A, Que L Jr., J. Am. Chem. Soc, 2018, 140, 5798–5804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Serrano-Plana J, Acuña-Parés F, Dantignana V, Oloo WN, Castillo E, Draksharapu A, Whiteoak CJ, Martin-Diaconescu V, Basallote MG, Luis JM, Que L Jr., M. Costas and A. Company, Chem. – Eur. J, 2018, 24, 5331–5340. [DOI] [PubMed] [Google Scholar]
  • 57.Xu S, Draksharapu A, Rasheed W and Que L Jr., J. Am. Chem. Soc, 2019, 141, 16093–16107. [DOI] [PubMed] [Google Scholar]
  • 58.Kim B, Jeong D, Ohta T and Cho J, Commun. Chem, 2019, 2, 81–86. [Google Scholar]
  • 59.Cho J, Sarangi R and Nam W, Acc. Chem. Res, 2012, 45, 1321–1330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Annaraj J, Cho J, Lee Y-M, Kim SY, Latifi R, de Visser SP and Nam W, Angew. Chem., Int. Ed, 2009, 48, 4150–4153. [DOI] [PubMed] [Google Scholar]
  • 61.Cho J, Jeon S, Wilson SA, Liu LV, Kang EA, Braymer JJ, Lim MH, Hedman B, Hodgson KO, Valentine JS, Solomon EI and Nam W, Nature, 2011, 478, 502–505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Shin B, Park Y, Jeong D and Cho J, Chem. Commun, 2020, 56, 9449–9452. [DOI] [PubMed] [Google Scholar]
  • 63.Shin B, Sutherlin KD, Ohta T, Ogura T, Solomon EI and Cho J, Inorg. Chem, 2016, 55, 12391–12399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Kim B, Jeong D and Cho J, Chem. Commun, 2017, 53, 9328–9331. [DOI] [PubMed] [Google Scholar]
  • 65.Hikichi S, Okuda H, Ohzu Y and Akita M, Angew. Chem., Int. Ed, 2009, 48, 188–191. [DOI] [PubMed] [Google Scholar]
  • 66.Sankaralingam M, Lee Y-M, Nam W and Fukuzumi S, Coord. Chem. Rev, 2018, 365, 41–59. [Google Scholar]
  • 67.Bae SH, Li X-X, Seo MS, Lee Y-M, Fukuzumi S and Nam W, J. Am. Chem. Soc, 2019, 141, 7675–7679. [DOI] [PubMed] [Google Scholar]
  • 68.Sankaralingam M, Lee Y-M, Jeon SH, Seo MS, Cho K-B and Nam W, Chem. Commun, 2018, 54, 1209–1212. [DOI] [PubMed] [Google Scholar]
  • 69.Tcho W-Y, Wang B, Lee Y-M, Cho K-B, Shearer J and Nam W, Dalton Trans, 2016, 45, 14511–14515. [DOI] [PubMed] [Google Scholar]
  • 70.Fukuzumi S, Cho K-B, Lee Y-M, Hong S and Nam W, Chem. Soc. Rev, 2020, 49, 8988–9027. [DOI] [PubMed] [Google Scholar]
  • 71.Armarego WLF and Chai CLL, Purification of Laboratory Chemicals, 6th ed., Pergamon Press, Oxford, 2009. [Google Scholar]
  • 72.Wittmann H, Raab V, Schorm A, Plackmeyer J and Sundermeyer J, Eur. J. Inorg. Chem, 2001, 2001, 1937–1948. [Google Scholar]
  • 73.Raab V, Kipke J, Burghaus O and Sundermeyer J, Inorg. Chem, 2001, 40, 6964–6971. [DOI] [PubMed] [Google Scholar]
  • 74.Brar AS and Kaur S, J. Polym. Sci. Pol. Chem, 2005, 43, 5906–5922. [Google Scholar]
  • 75.Goldsmith CR, Jonas RT and Stack TDP, J. Am. Chem. Soc, 2002, 124, 83–96. [DOI] [PubMed] [Google Scholar]
  • 76.Arunkumar C, Lee Y-M, Lee JY, Fukuzumi S and Nam W, Chem. – Eur. J, 2009, 15, 11482–11489. [DOI] [PubMed] [Google Scholar]
  • 77.Dhuri SN, Lee Y-M, Seo MS, Cho J, Narulkar DD, Fukuzumi S and Nam W, Dalton Trans, 2015, 44, 7634–7642. [DOI] [PubMed] [Google Scholar]
  • 78.Inada Y, Nakano Y, Inamo M, Nomura M and Funahashi S, Inorg. Chem, 2000, 39, 4793–4801. [DOI] [PubMed] [Google Scholar]
  • 79.Sheldrick GM, SHELXTL/PC Version 6.12 for Windows XP, Bruker AXS Inc., Madison: Wisconsin, USA, 2001. [Google Scholar]
  • 80.Ravel B and Newville M, J. Synchrotron Rad, 2005, 12, 537–541. [DOI] [PubMed] [Google Scholar]
  • 81.Sheldrick GM, SHELX-97 Program for Crystal Structure Determination, Universität Göttingen, Germany, 1997. [Google Scholar]
  • 82.Ankudinov AL, Ravel B, Rehr JJ and Conradson SD, Phys. Rev. B, 1998, 58, 7565–7576. [Google Scholar]
  • 83.Tao JM, Perdew JP, Staroverov VN and Scuseria GE, Phys. Rev. Lett, 2003, 91, 146401–146404. [DOI] [PubMed] [Google Scholar]
  • 84.Staroverov VN, Scuseria GE, Tao J and Perdew JP, J. Chem. Phys, 2003, 119, 12129–12137. [Google Scholar]
  • 85.Dolg M, Wedig U, Stoll H and Preuss H, J. Chem. Phys, 1987, 86, 866–872. [Google Scholar]
  • 86.Martin JML and Sundermann A, J. Chem. Phys, 2001, 114, 3408–3420. [Google Scholar]
  • 87.Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Petersson GA, Nakatsuji H, Li X, Caricato M, Marenich AV, Bloino J, Janesko BG, Gomperts R, Mennucci B, Hratchian HP, Ortiz JV, Izmaylov AF, Sonnenberg JL, Williams-Young D, Ding F, Lipparini F, Egidi F, Goings J, Peng B, Petrone A, Henderson T, Ranasinghe D, Zakrzewski VG, Gao J, Rega N, Zheng G, Liang W, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Throssell K, Montgomery JA Jr., Peralta JE, Ogliaro F, Bearpark MJ, Heyd JJ, Brothers EN, Kudin KN, Staroverov VN, Keith TA, Kobayashi R, Normand J, Raghavachari K, Rendell AP, Burant JC, Iyengar SS, Tomasi J, Cossi M, Millam JM, Klene M, Adamo C, Cammi R, Ochterski JW, Martin RL, Morokuma K, Farkas O, Foresman JB and Fox DJ, Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford CT, 2016. [Google Scholar]
  • 88.Klamt A and Schüürmann G, J. Chem. Soc., Perkin Trans. 2, 1993, 5, 799–805. [Google Scholar]
  • 89.Grimme S, Anthony J, Ehrlich S and Krieg HA, J. Chem. Phys, 2010, 132, 154104–154119. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Material
CIF

RESOURCES