Startseite Properties of the one-electron oxidized copper(II) salen-type complexes: relationship between electronic structures and reactivities
Artikel Öffentlich zugänglich

Properties of the one-electron oxidized copper(II) salen-type complexes: relationship between electronic structures and reactivities

  • Yuichi Shimazaki EMAIL logo
Veröffentlicht/Copyright: 18. Januar 2014

Abstract

The Cu(II)-phenoxyl radical formed during the catalytic cycle of galactose oxidase (GO) attracted much attention, and the structures and properties of a number of metal-phenoxyl radical complexes have been studied. Some of the functional model systems of GO reported previously have shown that the Cu complexes oxidize primary alcohols to aldehydes and that the Cu(II)-phenoxyl radical species is formed in the catalytic cycle. Many Cu(II)-phenoxyl radical species have been produced by one-electron oxidation of the Cu(II)-phenolate complexes. On the other hand, one-electron oxidation of a Cu(II)-phenolate complex has the possibility to give different electronic structures, one of which is the Cu(III)-phenolate. From these points of view, this micro review is focused on the one-electron oxidized square-planar Cu(II) complexes of the salen-type ligands. Introduction of substituents into the phenolate moieties and conversion from a 5- to a 6-membered chelate backbone alter the electronic structure of the one-electron oxidized Cu(II) complexes and give rise to a different reactivity of benzyl alcohol oxidation. The relationship between the electronic structure and the reactivity is herein discussed.

Introduction

Organic radical formation has been well established for metalloenzymes such as class I ribonucleotide reductase and prostaglandin endoperoxide synthase [1, 2]. It was discovered that the phenoxyl radical can bind to a metal ion as an open-shell ligand fulfilling the role of an organic radical cofactor in the metalloenzyme, galactose oxidase (GO), which is a single copper oxidase catalyzing two-electron oxidation of a primary alcohol to an aldehyde (Fig. 1) [3, 4]. A Cu(III) species had once been considered to be a possibility as the active form of GO [5, 6], but many physicochemical studies have revealed the formation of a Cu(II)-phenoxyl radical species [7, 8]. A proposed mechanism which has been generally accepted indicates that the oxidized form of GO in the Cu(II)-tyrosine radical state binds a primary alcohol in the fourth equatorial position, giving an aldehyde and the reduced form of GO by subsequent hydrogen atom abstraction and electron transfer via the ketyl radical intermediate [3, 4]. For an improved understanding of the detailed mechanism of GO and the properties of the metal complexes with the coordinated phenoxyl radical, many metal−phenolate complexes have been synthesized and characterized. Several review articles are available for early works on phenoxyl radical complexes [9–14].

Fig. 1 
          Proposed catalytic alcohol oxidation mechanism of GO [3, 4].
Fig. 1

Proposed catalytic alcohol oxidation mechanism of GO [3, 4].

Depending on the relative energies of the redox-active orbitals, a metal complex with non-innocent ligand exists in two limiting descriptions, either a metal-ligand radical (Mn+(L)) or a high-valent metal (M(n+1)+(L)) complex [15–19]. Oxidized metal salen-type complexes are known to exist in either form, and the factors that control the locus of oxidation in these complexes are being actively pursued [14, 20–22]. Therefore, this review deals with our recent studies on the synthesis of the one-electron oxidized Cu(II)-salen-type complexes (Fig. 2) and characterization of their physical properties and reactivities [23–26]. The substituents on the phenolate moieties would change the oxidation locus in the one-electron oxidized complexes as seen for the group 10 metal complexes [27]. Further, the 6-membered chelate backbone would lead to coordination plane distortion, and this structural change would influence the orbital overlap between the Cu ion and the ligand [28]. By a combination of experiments and calculations, the substituents on the phenol ring and the chelate ring size are now shown to alter the electronic structure and reactivity of the one-electron oxidized Cu(II) complexes.

Fig. 2 
          Structures of some Cu(II) salen-type complexes [23–26].
Fig. 2

Structures of some Cu(II) salen-type complexes [23–26].

Syntheses and redox behavior of Cu(II) complexes

The N2O2-donor ligands (Fig. 2) reacted with Cu(II) ion to give the corresponding Cu(II) salen-type complexes as crystals. The salen-type complexes of asymmetric phenolates were also prepared by a step-by-step method by use of the monohydrochloride salt of cyclohexyldiamine as shown in Scheme 1 [14].

Scheme 1 
            Synthetic route to asymmetric salen-type complexes [26].
Scheme 1

Synthetic route to asymmetric salen-type complexes [26].

The cyclic and differential pulse voltammetry (CV and DPV) results of Cu(II)-salen type complexes showed two reversible redox waves, whose potentials are similar to each other within the 0.25 V range as listed in Table 1. Among the symmetric complexes (R1 = R2), those bearing methoxy groups are oxidized at lower potentials than those bearing i-PrS groups. This order matches the electron-donating properties expected from Hammett σ+ constants of these substituents [29]. For the nonsymmetric complexes (R1 ≠ R2), the two redox couples observed can be correlated to the sequential oxidation of the two phenolates: The more electron-rich phenolate is oxidized at a lower potential and the less electron-rich phenolate at a higher potential [26]. These results can be seen as two phenolate-based oxidation processes. On the other hand, the potentials of a 6-membered chelate complex, Cu(1,3-salcn), were only 0.03 V lower than those of the 5-membered chelate Cu(1,2-salcn). These redox waves were found to be one-electron processes from the current analysis for each wave. This result suggests that the one-electron oxidized complexes can be generated by adding an equimolar amount of chemical oxidants such as AgSbF6 [30]. Indeed, the addition of an equimolar amount of AgSbF6 (E = 0.65 V vs. Fc/Fc+ in CH2Cl2) to the salen complexes in CH2Cl2 caused a color change, indicating formation of oxidized species.

Table 1

Electrochemical data for Cu salen-type complexes (E vs. Fc/Fc+) [24–26].

Complex E 1(mV) E 2(mV) E ave(mV) ΔE(mV)
Cu(1,2-salcn) 450 600 530 150
Cu(1,2-MeO2-salcn) 240 400 320 160
Cu(1,2-t-Bu,MeO-salcn) 250 590 340
Cu(1,2-i-PrS2-salcn) 340 440 385 110
Cu(1,2-t-Bu,i-PrS-salcn) 340 590 255
Cu(1,3-salcn) 480 630 550 150

Solid-state characterization of the oxidized Cu(II) complexes

Crystal structures

When the solutions of some of oxidized complexes prepared by AgSbF6 oxidation were kept at low temperature in CH2Cl2/toluene, crystals of the one-electron oxidized complexes could be isolated. The X-ray crystal structure analysis gave the structures of [Cu(1,2-salcn)]SbF6, [Cu(1,2-MeO2-salcn)]SbF6, and [Cu(1,3-salcn)]SbF6 (Fig. 3) [23–25]. All complexes exhibit a pseudo square-planar geometry formed by two phenolato oxygen atoms and two imino nitrogen atoms. The Cu coordination sphere geometry of [Cu(1,2-salcn)]+ is symmetric with contraction in comparison with neutral complex; on average the Cu–O and Cu–N bond lengths decrease by 0.04 and 0.03 Å, respectively, upon oxidation [23]. On the other hand, the coordination geometries in [Cu(1,2-MeO2-salcn)]+ and [Cu(1,3-salcn)]+ are asymmetric, indicating that one salicylideneimino moiety becomes a weaker donor in these complexeswith a longer bond to the Cu atom [24, 25]. The C–O bond length of the weakly donating phenolate moiety is shortened, while such a shortening of the C–O bond in [Cu(1,2-salcn)]+ could not be detected. This quinoid bonding pattern of [Cu(1,2-MeO2-salcn)]+ and [Cu(1,3-salcn)]+ are consistent with radical localization to one side of the molecule, in agreement with the bonding pattern observed for other localized phenoxyl radical complexes, whereas the solid state structure of [Cu(1,2-salcn)]+ suggests that it is a phenolate complex [31]. In addition, the SbF6 counterion was located close to the copper ion in [Cu(1,2-salcn)]+, while a close contact with the quinoid moiety was observed in [Cu(1,2-MeO2-salcn)]+ and [Cu(1,3-salcn)]+. These results support that [Cu(1,2-salcn)]+is assigned to the Cu(III)-phenolate species, while [Cu(1,2-MeO2-salcn)]+ and [Cu(1,3-salcn)]+ are assigned to the Cu(II)-phenoxyl radical species.

Fig. 3 
              Structures of Cu(III)-phenolate complex, [Cu(1,2-salcn)]SbF6, and Cu(II)-phenoxyl radical complexes, [Cu(1,3-salcn)]SbF6 and [Cu(1,2-MeO2-salcn)]SbF6.
Fig. 3

Structures of Cu(III)-phenolate complex, [Cu(1,2-salcn)]SbF6, and Cu(II)-phenoxyl radical complexes, [Cu(1,3-salcn)]SbF6 and [Cu(1,2-MeO2-salcn)]SbF6.

Magnetic properties

The temperature-dependent magnetic susceptibility of Cu(Salcn) complexes displayed a response typical for a simple d9 Cu(II) paramagnet, and the data were fit to a Curie–Weiss law expression (C = 0.405, θ = –0.554 K). The data for a powdered crystalline sample of [Cu(1,2-salcn)]SbF6 show that this sample is essentially diamagnetic (μeff = 0.3 μB) in the temperature range 5–300 K. They were fit with a ca. 4 % impurity mainly due to the neutral Cu(II) complex as estimated from the low-temperature data, which is consistent with a small amount of decomposition in the solid sample. These results suggest that [Cu(1,2-salcn)]SbF6 in the solid state has a diamagnetic electronic ground state (S = 0) with no thermally accessible triplet state at 300 K [23]. The electronic structure of [Cu(1,2-salcn)]SbF6 was also supported by XAFS measurement. The pre-edge of the solid sample of [Cu(1,2-salcn)]SbF6 was found to be more than 1 eV higher than that of neutral Cu(1,2-salcn), clearly indicating that the solid sample of the [Cu(1,2-salcn)]SbF6 has a metal-centered oxidation state.

The methoxy-substituted complexes, [Cu(1,2-MeO2-salcn)]+ and [Cu(1,2-t-Bu,MeO-salcn)]+, are EPR silent, since large zero-field splitting (ZFS) parameters are thought to render the one-electron oxidized species EPR silent at X-band frequencies [25, 26]. On the other hand, the EPR spectra for oxidized complexes [Cu(1,2-i-PrS2-salcn)]+ and [Cu(1,2-t-Bu,i-PrS-salcn)]+ exhibit markedly different behavior. One-electron oxidation does not result in a loss of EPR intensity, or rather, the signals broaden considerably, and the spin integration remains relatively constant at that of their parent neutral forms [26]. Thus, despite the similarities of the EPR spectra of the species before oxidation, significant substituent-dependent differences exist between the spectra of their one-electron oxidized species.

The Q-band spectrum of [Cu(1,3-salcn)]SbF6 at 12 K clearly consists of a two-line pattern centered at g = 2.06, as well as a half-field ΔMS = ±2 transition at g = 4 (Fig. 4a) [24]. This signature is characteristic of a spin triplet system having ZFS parameters smaller than the Q-band energy quantum (~1.1 cm–1). The spectrum could be satisfactorily simulated by using the ZFS parameters |D| = 0.470 ± 0.035 cm–1, E/D = 0.06 ± 0.02, and giso = 2.06. These parameters were additionally used for simulation of the X-band spectrum, and a reasonable agreement with the experimental data has been attained. From these EPR results, the triplet spin state of a Cu(II)-phenoxyl salen complex could be evidenced experimentally. The ZFS parameters reported here are the first measured experimentally for Cu(II)-phenoxyl salen complexes.

Fig. 4 
              (a) Q-band EPR spectrum of a frozen 0.02 M CH2Cl2 solution of [Cu(1,3-salcn)]SbF6: The solid black line is the experimental spectrum, and the dotted red line is a simulation of the triplet system using the parameters given in the text. The asterisk denotes a paramagnetic mononuclear copper (S = 1/2) impurity present in the sample; (b) temperature-dependent magnetic susceptibility of [Cu(1,3-salcn)]SbF6. Red dots are actual measurements, and the black line is a fitting based on the Bleaney–Bower’s equation.
Fig. 4

(a) Q-band EPR spectrum of a frozen 0.02 M CH2Cl2 solution of [Cu(1,3-salcn)]SbF6: The solid black line is the experimental spectrum, and the dotted red line is a simulation of the triplet system using the parameters given in the text. The asterisk denotes a paramagnetic mononuclear copper (S = 1/2) impurity present in the sample; (b) temperature-dependent magnetic susceptibility of [Cu(1,3-salcn)]SbF6. Red dots are actual measurements, and the black line is a fitting based on the Bleaney–Bower’s equation.

The temperature dependent magnetic susceptibility of [Cu(1,3-salcn)]SbF6 in the solid state exhibited a decrease in the susceptibility as the temperature was lowered (Fig. 4b). This fitting is contradictory to the result of EPR experiments. The best-fit parameters are J/kB = 4.1 cm–1 and θ = –3.4 cm–1 using the values from the results of EPR measurements, g = 2.06 and D/kB = 0.47 cm–1. These results indicate that the intramolecular interaction between the Cu and radical unpaired electron spins (J/kB) is ferromagnetic, while the intermolecular interaction (θ) is weakly antiferromagnetic. This intramolecular coupling value matches with the EPR experiments. Thus, complex [Cu(1,3-salcn)]+ can be assigned to a Cu(II)-phenoxyl radical species with a S = 1 ground state exhibiting a weak ferromagnetic interaction between the Cu(II) center and the ligand radical [32, 33].

Characterization of the oxidized complexes in solution

The electronic spectrum of Cu(1,2-salcn) is typical of a Cu(II) d9 complex with an intense CT transition at 26 500 cm–1 (ε = 11 600 M–1 cm–1) and a weak d-d transition at 17 600 cm–1 (ε = 600 M–1 cm–1) . As reported previously, oxidation to [Cu(1,2-salcn)]+ at room temperature results in the appearance of two new bands, an intense band at 18 600 cm–1 (ε = 6500 M–1 cm–1) and a low-energy transition at 5800 cm–1 (ε = 2900 M–1cm–1) (Fig. 5a). Interestingly, the 18 600 cm–1 band exhibits large intensity changes with temperature; from 298 to 190 K, this band approximately doubles in intensity and shifts slightly to higher energy. This change is fully reversible and indicates that [Cu(1,2-salcn)]+ is involved in a temperature-dependent equilibrium in solution. The intensity of the 557-nm absorption band for [Cu(1,2-salcn)]+ is independent of concentration at both 295 and 203 K, which excludes dimerization as a possible reason for the observed temperature-dependent changes in band intensity. The 1H NMR spectrum of [Cu(1,2-salcn)]+ in CD2Cl2 at 298 K displayed broad signals over a wide spectral range, indicating the presence of a substantial amount of a paramagnetic species. Solution susceptibility measurements by the Evans method indicate a χmT value of 0.442 cm3mol–1 K (1.1 ± 0.1 unpaired electrons; spin-only calculation) at room temperature. An interpretation of this bulk susceptibility is the presence of approximately equimolar amounts of a ferromagnetically coupled Cu(II)-phenoxyl radical species and a diamagnetic counterpart Cu(III)-phenolate [23].

Fig. 5 
            UV–vis spectra of complexes Cu(1,2-salcn) and Cu(1,3-salcn) in CH2Cl2 before and after oxidation: (a) complex Cu(1,2-salcn), (b) complex Cu(1,3-salcn). Black line: neutral complexes, red line: one-electron oxidized complexes.
Fig. 5

UV–vis spectra of complexes Cu(1,2-salcn) and Cu(1,3-salcn) in CH2Cl2 before and after oxidation: (a) complex Cu(1,2-salcn), (b) complex Cu(1,3-salcn). Black line: neutral complexes, red line: one-electron oxidized complexes.

On the other hand, the UV–vis–NIR absorption spectrum of [Cu(1,3-salcn)]+ in CH2Cl2 is different from that of [Cu(1,2-salcn)]+ (Fig. 5b). The NIR bands of [Cu(1,3-salcn)]+ are of significantly lower intensity compared to the NIR transitions for the Cu(III) ground-state complex [Cu(1,2-salcn)]+ and fully delocalized ligand radical complexes [23, 34]. From the low intense NIR bands, complex [Cu(1,3-salcn)]+ can be assigned to the class II localization system based on the Robin and Day classification [35]. Further, the resonance Raman spectrum of [Cu(1,3-salcn)]+ in CH2Cl2 showed the typical phenoxyl radical ν7a band at 1492 cm–1 [24]. Therefore, the absorption spectral features of [Cu(1,3-salcn)]+ clearly suggest that complex [Cu(1,3-salcn)]+ is a localized Cu(II)-phenoxyl radical species in solution.

The majority of Cu-salen phenoxyl radical complexes with methoxy and isopropylthio substitution display Class II-like behavior, regardless of the phenolate substitution pattern. Recent reports on the band-shape analysis of a series of nonsymmetric Cu-salen phenoxyl radical complexes indicate that the electron coupling coefficient values of HAB, whose average is at 2100 ± 200 cm–1, remain rather constant despite the different functional groups and the wide range of IVCT νmax values [14, 26]. From the relationship between the redox potential and the IVCT band of the oxidized asymmetrical salcn complexes, the absorption spectral feature of the active form of GO can be explained. The active site of GO has two different phenolate donors, one of which is the Tyr 272 residue having an o-alkylsulfanyl moiety. Assuming the difference of the oxidation potentials between the two phenolates (ca. 600 mV ≈ 4800 cm–1) and the sum of their reorganizational energies, the LLCT band at ca. 13 600 cm–1 is predicted for the active form of GO. This analysis is admittedly very approximate but very well within the envelope of the broad visible band of the enzyme [26].

Reactivity of the oxidized complexes with benzyl alcohol

Some of the salen-type Cu(II) complexes have been reported to have a GO-like catalytic reactivity for oxidation of primary alcohols, while the mechanism of the oxidation is still unclear [36–39]. On the other hand, Stack et al. have reported a detailed reaction mechanism of the one-electron oxidized salen-type complexes including [Cu(1,2-salcn)]+ [40, 41]. However, the electronic structure of [Cu(1,2-salcn)]+ was assigned simply to “a Cu(II)-phenoxyl radical”, which is slightly different from that discussed above. Recently, Thomas et al. have reported that phenoxyl radical complex [Cu(1,2-MeO2-salcn)]+ has a high reactivity for oxidation of benzyl alcohol, but the ionic strength of the solution was quite different due to generation of the oxidized species by electrolysis [25]. Therefore, we studied the detailed reaction mechanism of [Cu(1,3-salcn)]+ and compared it with that of [Cu(1,2-salcn)]+, in order to discuss the relationship between electronic structure and reactivity under the same conditions.

CH2Cl2 solutions of the oxidized complexes [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+are relatively stable at 293 K. Addition of excess benzyl alcohol to the solutionscaused a color change within a few hours. GC-MS and absorption spectral changes revealed that benzaldehyde was formed in 50 % yield relative to the initial amount of the oxidized complexes. This reaction was also observed under a nitrogen atmosphere, which indicates that the presence of air did not affect the benzyl alcohol oxidation. The UV–vis absorption spectral change showed that the spectra of the final products are different from those of Cu(1,2-salcn) and Cu(1,3-salcn). However, addition of one equivalent of Et3N to the solutions of the final products afforded spectra which are similar to those of 1 and 2, indicating that the final product of this reaction can be assigned to the proton adducts of the Cu(II) complexes, [Cu(H1,2-salcn)]+ and [Cu(H1,3-salcn)]+. The overall stoichiometry of the reaction observed is expressed as follows [24]:

In the presence of a large excess of benzyl alcohol, conversion of [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ can be regarded as a first-order process, suggesting that only one molecule of [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+, respectively,is involved in the rate-determining step. The first-order decay constant (kobs) of complex [Cu(1,3-salcn)]+ was not changed by varying the concentration of the complex (kobs = (3.61–4.06) × 10–3 s–1 for 0.1–0.3 mM solution of [Cu(1,3-salcn)]+ at [C6H5CH2OH] = 1.0 M), while the rate constant depended on the benzyl alcohol concentration. A plot of kobs against the concentration of benzyl alcohol exhibited a linear correlation, i.e., kobs = k2[C6H5CH2OH], giving the second-order rate constant k2 = (3.89 ± 0.05) × 103 M–1s–1 at 293 K for the rate-determining process. We could not determine the association constants of complexes [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ with benzyl alcohol, due to the linear correlation. The reactivity characteristics of [Cu(1,3-salcn)]+ are very similar to the reaction of complex [Cu(1,2-salcn)]+ with benzyl alcohol, although the second-order rate constant of [Cu(1,3-salcn)]+ is ca. 4 times larger than that of [Cu(1,2-salcn)]+ (k2 = 1.02 × 10–3 M–1s–1 at 293 K was obtained at [C6H5CH2OH] = 1.0 M). From the kinetic analyses, the primary alcohol oxidation by phenoxyl radical complex [Cu(1,3-salcn)]+ likely occurs by a mechanism similar to that of [Cu(1,2-salcn)]+ but with a faster reaction rate.

In order to determine whether C–H bond cleavage is involved in the rate-determining step, we carried out kinetic analyses using α-benzyl-d2 alcohol. The kinetic isotope effect (KIE), k2(H)/k2(D), for complex [Cu(1,2-salcn)]+ was estimated to be 19 at 298 K [40, 41], and the KIE value of [Cu(1,3-salcn)]+ was estimated to be 13 at 293 K [24], which indicates that the C–H bond scission is included in the rate-determining step. On the other hand, the reaction rate for C6H5CH2OD (k2 = 3.6 × 103 M–1s–1 obtained at [C6H5CH2OD] = 1.0 M) was very similar to that of C6H5CH2OH, indicating that the benzyl alcohol oxidation proceeds without the O–H hydrogen abstraction in the rate-determining step.

The Eyring plots of complexes [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ showed linear fits with similar slopes. The activation parameters of [Cu(1,2-salcn)]+ were determined to be ΔH = 10.7 ± 0.1 kcal mol–1 and ΔS = –35.6 ± 0.3 cal mol–1K–1, and those of [Cu(1,3-salcn)]+ were estimated to be ΔH = 10.2 ± 0.1 kcal mol–1 and ΔS = –34.8 ± 0.5 cal mol–1K–1 (Table 2). The difference of the activation energy (ΔEa) between [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ is estimated to be 0.6 ± 0.2 kcal mol–1 at 293 K. This value is in good agreement with the C–H activation energy difference estimated from the KIE values (Δ(ΔEaKIE) = 0.4 kcal mol–1) [42]. Thus, the activation parameter difference between [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ is relatively small, suggesting a similar reaction mechanism for [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+, as well as similar geometries for the substrate association.

Table 2

Activation parameters for the reaction of [Cu(1,2-salcn)]+ and [Cu(1,3-salcn)]+ with benzyl alcohol.

Complex ΔH/kcal mol–1 ΔS/cal mol–1K–1 E a/kcal mol–1a
[Cu(1,2-salcn)]+ 10.7 ± 0.1 –35.6 ± 0.3 11.3 ± 0.1
[Cu(1,3-salcn)]+ 10.2 ± 0.1 –34.8 ± 0.5 10.8 ± 0.2

The increased reaction rate for [Cu(1,3-salcn)]+ in comparison to [Cu(1,2-salcn)]+ can be attributed to the different electronic structure, i.e., a Cu(II)-radical for [Cu(1,3-salcn)]+ and a Cu(III) ground state for [Cu(1,2-salcn)]+ which is in equilibrium with the Cu(II)-ligand radical in solution at room temperature. The constant k2 for the reaction of [Cu(1,2-salcn)]+ corresponds to the rate constant of a simple bimolecular process between Cu(II)-radical and C6H5CH2OH (k2*) in the rate-determining step (k2 = k2*), while in the reaction of [Cu(1,2-salcn)]+, k2 should involve an equilibrium between Cu(III) and Cu(II)-radical species (Kpre= [Cu(II)-radical]/[Cu(III)]) as expressed by the following equation [42]:

Since complex [Cu(1,2-salcn)]+ has been reported to exist in ca. 1:1 equilibrium between Cu(III)-phenolate and Cu(II)-phenoxyl radical species, Kpre is estimated to be ≈1, and hence k2 for the reaction of [Cu(1,2-salcn)]+ is calculated to be k2k2*/2. If the Cu(III) species does not react with benzyl alcohol, the k2 value for the reaction of [Cu(1,3-salcn)]+ is at least double the value for [Cu(1,2-salcn)]+ on the assumption that k2* is the same for both systems [23].

From these results and discussion, the reaction mechanisms for the benzyl alcohol oxidation are summarized as shown in Scheme 1. The reaction mechanisms of the two [Cu(Salcn)]+ species are very similar. The rate-determining step of these complexes involves the hydrogen atom abstraction from the α-position of benzyl alcohol. Complex [Cu(1,3-salcn)]+ has a slightly faster reaction rate in comparison to [Cu(1,2-salcn)]+ (Scheme 2). The slower reaction rate of [Cu(1,2-salcn)]+ is mainly due to the existence of the equilibrium between the Cu(III)-phenolate and Cu(II)-phenoxyl radical species. In addition, it is plausible that the higher reactivity of the Cu(II)-radical of [Cu(1,3-salcn)]+ toward C6H5CH2OH may be due to the increased conformational flexibility of the 6-membered chelate ring of [Cu(1,3-salcn)]+ in comparison to the 5-membered chelate 1,2-cyclohexanediamine backbone of [Cu(1,2-salcn)]+.

Scheme 2 
            Proposed mechanisms for benzyl alcohol oxidation: (a) simple hydrogen abstraction by a localized phenoxyl radical such as complex [Cu(1,3-salcn)]+; (b) reaction scheme for the equilibrium between less reactive Cu(III)-phenolate and active Cu(II)-phenoxyl radical species in the reaction of complex [Cu(1,2-salcn)]+.
Scheme 2

Proposed mechanisms for benzyl alcohol oxidation: (a) simple hydrogen abstraction by a localized phenoxyl radical such as complex [Cu(1,3-salcn)]+; (b) reaction scheme for the equilibrium between less reactive Cu(III)-phenolate and active Cu(II)-phenoxyl radical species in the reaction of complex [Cu(1,2-salcn)]+.

Summary

This review focused on characterization of one-electron oxidized Cu(II)-salen-type complexes, [Cu(Salcn)]. The tert-butyl substituted 5-membered chelate complex, [Cu(1,2-salcn)]+, has very unique properties. This complex has a Cu(III)-phenolate ground state, while the CH2Cl2 solution at room temperature exhibits the valence tautomerism between Cu(III)-phenolate and Cu(II)-phenoxyl radical. Introduction of a tert-butyl group into the para-position of the phenolate moiety changed its characteristics to relatively localized phenoxyl radical species. However, the characteristics of the salen-type complexes, such as magnetic properties and redox potentials, depend on the sort of the substituents. Further, despite the same donor group in salen-type complexes, the chelate ring size difference gives rise to subtle differences in the electronic structure and oxidation reactivity with benzyl alcohol. One-electron oxidized Cu(1,3-Salcn) exists as a ligand-based radical complex in solution and in the solid state. Both [Cu(1,2-salcn)]+ and [Cu(1,3-Salcn)]+ exhibit a similar reaction mechanism for benzyl alcohol oxidation, but the reaction rate of [Cu(1,3-Salcn)]+ is ca. 4 times faster than that of [Cu(1,2-Salcn)]+ under the same conditions. The oxidation of benzyl alcohol can be described as a second-order process for the complexes without saturation kinetics, and the rate-determining step for both complexes involves hydrogen abstraction. This agreement indicates that the reactivity difference arises from the hydrogen abstraction mechanism, for which the localized Cu(II)-phenoxyl radical is more favorable. From these results, we suggest that the localized phenoxyl radical and possibly the increased ligand flexibility accelerate benzyl alcohol oxidation. In this connection, we have reported that one of the Schiff base phenolate complexes, a salophen (N,N-bis(salicyliden)-1,2-diaminobenzene) ligand complex, has a different electronic structure, which is assigned to an o-diiminobenzene radical species [43].

The radical complexes may show different reactivities and reaction mechanisms depending on subtle differences in the geometrical and electronic structures and will provide an interesting field of investigation in chemistry and biology. Further studies on the correlation between reactivity and electronic structure described herein are under way.


Corresponding author: Yuichi Shimazaki, College of Science, Ibaraki University, Bunkyo, Mito 312-8512, Japan, e-mail:


A collection of invited papers based on presentations at the 33rd International Conference on Solution Chemistry (ICSC-33), Kyoto, Japan, 7–12 July 2013.


The author would like to thank all collaborators, especially Prof. Dr. T. D. P. Stack (Stanford University), Prof. Dr. Tim Storr (Simon Fraser University), Prof. Dr. Fabrice Thomas (Université Joseph Fourier), Prof. Dr. Satoshi Iwatsuki (Konan University), Prof. Dr. Fumito Tani (Kyushu University), and Prof. Dr. Osamu Yamauchi (Nagoya University and Kansai University). This work was supported by Grants-in-Aid for Scientific Research (Nos. 22550055 and 25410060) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan, and by the Cooperative Research Program of Network Joint Research Center for Materials and Devices (Institute for Materials Chemistry and Engineering, Kyushu University).

References

[1] J. Stubbe, W. A. van der Donk. Chem. Rev.98, 705 (1998).10.1021/cr9400875Suche in Google Scholar

[2] P. A. Frey, A. D. Hegeman, G. H. Reed. Chem. Rev.106, 3302 (2006).10.1021/cr050292sSuche in Google Scholar

[3] J. W. Whittaker. Met. Ions Biol. Syst.30, 315 (1994).10.1007/BF02207484Suche in Google Scholar

[4] J. W. Whittaker. Chem. Rev.103, 2347 (2003).10.1021/cr020425zSuche in Google Scholar PubMed

[5] G. R. Dyrkacz, R. D. Libby, G. A. Hamilton. J. Am. Chem. Soc.98, 626 (1976).10.1021/ja00418a060Suche in Google Scholar

[6] G. A. Hamilton, P. K. Adolf, J. de Jersey, G. C. DuBois, G. R. Dyrkacz, R. D. Libby. J. Am. Chem. Soc.100, 1899 (1978).10.1021/ja00474a042Suche in Google Scholar

[7] K. Clark, J. E. Penner-Hahn, M. M. Whittaker, J. W. Whittaker. J. Am. Chem. Soc.112, 6433 (1990).10.1021/ja00173a061Suche in Google Scholar

[8] M. L. McGlashen, D. D. Eads, T. G. Spiro, J. W. Whittaker. J. Phys. Chem.99, 4918 (1995).10.1021/j100014a008Suche in Google Scholar

[9] B. A. Jazdzewski, W. B. Tolman. Coord. Chem. Rev.200–202, 633 (2000).10.1016/S0010-8545(00)00342-8Suche in Google Scholar

[10] P. Chaudhuri, K. Wieghardt. Prog. Inorg. Chem.50, 151 (2001).10.1002/0471227110.ch2Suche in Google Scholar

[11] F. Thomas. Eur. J. Inorg. Chem. 2379 (2007).10.1002/ejic.200601091Suche in Google Scholar

[12] Y. Shimazaki, O. Yamauchi. Indian J. Chem. Sect. A50, 383 (2011).Suche in Google Scholar

[13] Y. Shimazaki. In: Electrochemistry, pp. 51–67, InTech (2013).Suche in Google Scholar

[14] C. T. Lyons, T. D. P. Stack. Coord. Chem. Rev. 257, 528 (2013).10.1016/j.ccr.2012.06.003Suche in Google Scholar

[15] P. J. Chirik, K. Wieghardt. Science327, 794 (2010).10.1126/science.1183281Suche in Google Scholar PubMed

[16] P. L. Holland. Acc. Chem. Res. 41, 905 (2008).10.1021/ar700267bSuche in Google Scholar PubMed PubMed Central

[17] B. de Bruin, D. G. H. Hetterscheid, A. J. J. Koekkoek, H. Grützmacher. Prog. Inorg. Chem. 55, 247 (2007).10.1002/9780470144428.ch5Suche in Google Scholar

[18] P. Chaudhuri, C. N. Verani, E. Bill, E. Bothe, T. Weyhermüller, K. Wieghardt. J. Am. Chem. Soc.123, 2213 (2001).10.1021/ja003831dSuche in Google Scholar

[19] D. Herebian, E. Bothe, E. Bill, T. Weyhermüller, K. Wieghardt. J. Am. Chem. Soc.123, 10012 (2001).10.1021/ja011155pSuche in Google Scholar

[20] Y. Shimazaki, F. Tani, K. Fukui, Y. Naruta, O. Yamauchi. J. Am. Chem. Soc. 125, 10512 (2003).10.1021/ja035806oSuche in Google Scholar

[21] Y. Shimazaki, T. Yajima, F. Tani, S. Karasawa, K. Fukui, Y. Naruta, O. Yamauchi. J. Am. Chem. Soc.129, 2559 (2007).10.1021/ja067022rSuche in Google Scholar

[22] T. Storr, E. C. Wasinger, R. C. Pratt, T. D. P. Stack. Angew. Chem., Int. Ed.46, 5198 (2007).10.1002/anie.200701194Suche in Google Scholar

[23] T. Storr, P. Verma, R. C. Pratt, E. C. Wasinger, Y. Shimazaki, T. D. P. Stack. J. Am. Chem. Soc.130, 15448 (2008).10.1021/ja804339mSuche in Google Scholar

[24] K. Asami, K. Tsukidate, S. Iwatsuki, F. Tani, S. Karasawa, L. Chiang, T. Storr, F. Thomas, Y. Shimazaki. Inorg. Chem.51, 12450 (2012).10.1021/ic3018503Suche in Google Scholar

[25] M. Orio, O. Jarjayes, H. Kanso, C. Philouze, F. Neese, F. Thomas. Angew. Chem., Int. Ed. 49, 4989 (2010).10.1002/anie.201001040Suche in Google Scholar

[26] R. C. Pratt, C. T. Lyons, E. C. Wasinger, T. D. P. Stack. J. Am. Chem. Soc. 134, 7367 (2012).10.1021/ja211247fSuche in Google Scholar

[27] L. Chiang, A. Kochem, O. Jarjayes, T. J. Dunn, H. Vezin, M. Sakaguchi, T. Ogura, M. Orio, Y. Shimazaki, F. Thomas, T. Storr. Chem.—Eur. J. 18, 14117 (2012).10.1002/chem.201201410Suche in Google Scholar

[28] Y. Shimazaki, N. Arai, T. J. Dunn, T. Yajima, F. Tani, C. F. Ramogida, T. Storr. Dalton Trans. 40, 2469 (2011).10.1039/c0dt01574aSuche in Google Scholar

[29] C. Hansch, A. Leo, R. W. Taft. Chem. Rev. 91, 165 (1991).10.1021/cr00002a004Suche in Google Scholar

[30] N. G. Connelly, W. E. Geiger. Chem. Rev. 96, 877 (1996).10.1021/cr940053xSuche in Google Scholar

[31] Y. Shimazaki. Adv. Mater. Phys. Chem.3, 60 (2013).Suche in Google Scholar

[32] O. Kahn. Molecular Magnetism, Wiley-VCH, Weinheim (1993).Suche in Google Scholar

[33] B. Bleany, K. D. Bowers. Proc. R. Soc. LondonA 214, 451 (1952).10.1098/rspa.1952.0181Suche in Google Scholar

[34] Y. Shimazaki, T. D. P. Stack, T. Storr. Inorg. Chem.48, 8383 (2009).10.1021/ic901003qSuche in Google Scholar

[35] M. B. Robin, P. Day. Adv. Inorg. Chem.10, 247 (1967).10.1016/S0065-2792(08)60179-XSuche in Google Scholar

[36] N. Kitajima, K. Whang, Y. Moro-Oka, A. Uchida, Y. Sasada. J. Chem. Soc., Chem. Commun. 1504 (1986).10.1039/c39860001504Suche in Google Scholar

[37] Y. Wang, T. D. P. Stack. J. Am. Chem. Soc. 118, 13097 (1996).10.1021/ja9621354Suche in Google Scholar

[38] K. O. Hodgson, T. D. P. Stack. Science279, 537 (1998).10.1126/science.279.5350.537Suche in Google Scholar PubMed

[39] K. Butsch, T. Günther, A. Klein, K. Stirnat, A. Berkessel, J. Neudörfl. Inorg. Chim. Acta394, 237 (2013).10.1016/j.ica.2012.08.016Suche in Google Scholar

[40] R. C. Pratt, T. D. P. Stack. J. Am. Chem. Soc.125, 8716 (2003).10.1021/ja035837jSuche in Google Scholar

[41] R. C. Pratt, T. D. P. Stack. Inorg. Chem.44, 2367 (2005).10.1021/ic048695iSuche in Google Scholar

[42] J. H. Espenson. In: Chemical Kinetics and Reaction Mechanisms, 2nd ed., McGraw-Hill, New York (2002).Suche in Google Scholar

[43] A. Kochem, O. Jarjayes, B. Baptiste, C. Philoze, H. Vezin, K. Tsukidate, F. Tani, M. Orio, Y. Shimazaki, F. Thomas. Chem.—Eur. J. 18, 1068 (2012).10.1002/chem.201102882Suche in Google Scholar

Published Online: 2014-01-18
Published in Print: 2014-02-01

©2014 IUPAC & De Gruyter Berlin Boston

Artikel in diesem Heft

  1. Masthead
  2. Masthead
  3. Preface
  4. International Union of Pure and Applied Chemistry
  5. Conference paper
  6. Optimization of superamphiphobic layers based on candle soot
  7. Probing “ambivalent” snug-fit sites in the KcsA potassium channel using three-dimensional reference interaction site model (3D-RISM) theory
  8. Perspectives for hybrid ab initio/molecular mechanical simulations of solutions: from complex chemistry to proton-transfer reactions and interfaces
  9. The complex structure of ionic liquids at an atomistic level: from “red-and-greens” to charge templates
  10. Amide I IR probing of core and shell hydrogen-bond structures in reverse micelles
  11. Kinetic studies on cyclopalladation in palladium(II) complexes containing an indole moiety
  12. Properties of the one-electron oxidized copper(II) salen-type complexes: relationship between electronic structures and reactivities
  13. Solvation of a sponge-like geometry
  14. What is “hypermobile” water?: detected in alkali halide, adenosine phosphate, and F-actin solutions by high-resolution microwave dielectric spectroscopy
  15. Reentrant condensation, liquid–liquid phase separation and crystallization in protein solutions induced by multivalent metal ions
  16. Emulsion-templated macroporous polymer/polymer composites with switchable stiffness
  17. Effective interaction between small unilamellar vesicles as probed by coarse-grained molecular dynamics simulations
  18. Enthalpies of solution, limiting solubilities, and partial molar heat capacities of n-alcohols in water and in trehalose crowded media
  19. Protonation of alkanolamines and cyclic amines in water at temperatures from 293.15 to 373.15 K
  20. IUPAC Technical Report
  21. Defining the transfer coefficient in electrochemistry: An assessment (IUPAC Technical Report)
  22. IUPAC Recommendations
  23. Definition of the transfer coefficient in electrochemistry (IUPAC Recommendations 2014)
Heruntergeladen am 30.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/pac-2014-5022/html
Button zum nach oben scrollen