Abstract
Understanding the details of electronic properties of mixed-valence (MV) states of organometallic molecular wires is essential to gain insights into electron transfer phenomena. Although the field of MV chemistry is matured, there remain issues to be solved, which cannot be accessed by the conventional analytical methods. Here, we describe the synthesis and properties of diruthenium bridging (diethynylbenzene)diyl wires, (μ-p and m-C≡C–C6H4–C≡C){RuCpR(dppe)}21, with the acyl-substituted cyclopentadienyl rings [CpR: Cpe; η5-C5H4COOMe (a-series: ester derivatives), Cpa; η5-C5H4COMe (b-series: acetyl derivatives)], which are installed as IR-tags to estimate electron densities at the metal centers in the MV species. The electrochemical and IR/near IR-spectroelectrochemical studies reveal that the two metal centers in the para-isomers p-1a,b+ interact with each other more strongly than those in the meta-isomers m-1a,b+. Electron-spin resonance study also supports the radicals being delocalized over the Ru-(p-C≡C–C6H4–C≡C)–Ru moieties in p-1a,b+. The spectroelectrochemical IR study shows significant higher-energy shifts of the ν(C=O) vibrations brought about upon 1e-oxidation. Spectral simulation on the basis of the Bloch equations allows us to determine the electron transfer rate constants (ket) between the two metal centers being in the orders of 1012 s–1 (p-1+) and 109 s–1 (m-1+). The shifts of the ν(C=O) bands reveal that the charge densities on the para-isomers p-1a,b+ are widely delocalized over the Ru-(p-C≡C–C6H4–C≡C)–Ru linkages in contrast to the meta-isomers m-1a,b+, where the electron densities are mainly localized on the metal fragments, as supported by the density functional theory and time-dependent density functional theory studies as well as comparison with the reference mononuclear acetylide complexes, C6H5–C≡C–RuCpR(dppe) 2. We have successfully demonstrated that the carbonyl groups (>C=O) in the ancillary Cp ligands also work as IR-tags to report detailed information on the electron densities at the metal centers and the electron distribution over MV complexes as well.
Keywords: mixed-valence species, ruthenium complex, infra-red tag, metal alkynyl complex, acylcyclopentadienyl ligand
Introduction
Mixed-valence (MV) chemistry of inorganic and organometallic complexes has been continuously developed to deepen the fundamental understanding of the electron transfer processes involved since the discovery of the monumental Creutz–Taube ion, [Ru(NH3)5-μ-pyrazine-Ru(NH3)5]5+, in 1969, and the complex is regarded as the first MV coordination complex.1 A central issue of this area is the understanding of the electronic properties of the MV state, that is, the extent of charge delocalization over the molecule and the electron transfer rate between the two redox sites. Robin and Day classified the MV complexes into three categories, that is, a fully delocalized state (class III), a fully localized state (class I), and an intermediate state between them (class II).2 For the determination of the class, various physicochemical methods such as electrochemistry, electron-spin resonance (ESR), Mössbauer, X-ray photoelectron spectroscopy (XPS), and UV–vis–NIR (near IR) absorption spectroscopy have been used. However, discussion of the classification is not always straightforward. For example, the comproportionation constant (KC), which indicates the thermodynamic stability of the MV species, can be readily obtained from the CV study, but the redox potentials are strongly influenced by the environment such as the counter anions and the solvents used for the measurements.3,4 The intervalence charge transfer (IVCT) band in the NIR region observed for the MV species is commonly used for their classification. According to the Marcus–Hush theorem,5 the strength of electronic coupling (Vab) can be determined from the IVCT band parameters. Because, however, inorganic and organometallic MV complexes frequently show multiple and overlapped NIR bands, identification of the IVCT band is frequently ambiguous.6
IR spectroscopy is another powerful method to evaluate the electronic structure of MV complexes because the timescale of IR spectroscopy is short enough (∼10–12 s) to evaluate electron transfer processes of MV complexes. In the seminal work by Kubiak and Ito using the IR tag technique, they reported the electron transfer rates for a series of MV complexes composed of the two μ3-oxotriruthenium clusters bridged by a pyrazine linker on the basis of the detailed analysis of the IR data of the terminal CO ligands.7 The successful IR analysis involving line shape analysis afforded the electron transfer rate constants (ket) of the MV complexes, which were in the order of 1012 s–1. On the other hand, studies on organometallic MV systems by the IR-tag technique are rare. Winter reported dinuclear MV complexes with the vinylruthenium fragments having supporting CO ligands,8,9 and on the basis of the spectroelectrochemical IR study, the localized and delocalized charge states were successfully distinguished. While the methods are relatively simple and provide a variety of information, the studies reported to date are mainly limited to the MV complexes with CO ligands showing intense IR absorptions.10−13
Multinuclear complexes bearing the group 8 metal acetylide fragments, CP(dppe)M–C≡C [M = Fe, Ru; CP = cyclopentadinyl (Cp) or 1,2,3,4,5-pentamethylcyclopentadienyl (Cp*); dppe = 1,2-bis(diphenylphosphino)ethane], have been extensively studied due to the thermodynamic stability of their 1e-oxidized monocationic MV states caused by the electron-rich metal fragments, and such a feature allows studies of a wide variety of one- and two-dimensional multinuclear species.14−19 Deep understanding of electronic structures of MV is important. For example, Low proposed another MV state called “bridge-localized state”. 1e-Oxidation of Cp*(dppe)Ru–C≡C–Ar–C≡C–Ru(dppe)Cp* (Ar = 1,4-benzenediyl, 1,4-naphthalenediyl, and 9,10-anthracenediyl) formed the MV species, where the positive charge is mainly localized on the organic bridging part rather than on the metal centers, and the trend became remarkable as the π-system of the bridging aromatic ring was enlarged.20 Furthermore, details of the electronic structures of such ‘‘bridge-localized’’ states have been investigated on the basis of a theoretical study, which showed that the conformation of the Ru fragments is a dominant factor for the charge localization/delocalization. Detailed density functional theory (DFT) analyses by Kaupp suggested that the parallel and antiparallel conformational isomers with respect to the two Cp*Ru(dppe) fragments exhibit the bridge-localized state, while the twisted conformation causes the metal-localized state (Scheme 1).21
Scheme 1. Schematic Descriptions of Metal-Localized and Bridge-Localized States for CP(dppe)Ru–C≡C–Ar–C≡C–Ru(dppe)CP Complexes Associated with Their Conformational Isomers.
To experimentally clarify the charge distribution, the electron density, that is, the oxidation states, at the ruthenium centers should be determined. We envisaged that new metal fragments with IR-active pendants should provide information on the electron density at the metal centers. Herein, we report the synthesis and characterization of dinuclear ruthenium complexes bridged by the p- and m-(diethynylbenzene)diyl ligands (1) with the Cp ligands having the IR-active acyl groups, CpRRu(dppe) [CpR: Cpe; η5-C5H4COOMe (a-series: ester derivatives), Cpa; η5-C5H4COMe (b-series: acetyl derivatives)], together with the reference mononuclear complexes 2. The electronic structures of the MV species thereof are analyzed on the basis of the ν(C=O) vibrations (Figure 1). Dinuclear MV complexes with (diethynylbenzene)diyl bridges have been studied extensively22,23 and thus would be suitable motifs to verify the usefulness of IR-tags.
Figure 1.
Structures of mono- and di-nuclear complexes with IR-active ruthenium fragments studied in the present work.
Results and Discussion
Synthesis and Characterization
Ruthenium chloride precursors, CpRRu(dppe)Cl [R = COOCH3 (3a), COCH3 (3b)], were synthesized by the reaction of the CpRNa derivatives with RuCl2(PPh3)3, followed by the ligand substitution with dppe in good yields in a manner similar to the synthesis of the Cp derivative.24 Sequential treatment of 3a and 3b with AgBF4 and the appropriate terminal alkyne in the presence of DBU afforded the corresponding mono- and di-nuclear complexes 1 and 2 in 45–75% yields (Scheme 2). It is noted that the conventional synthetic method using Na+/K+ PF6–25 instead of the silver salt left the starting chloride precursors 3 untouched probably due to the stronger Ru–Cl bonds caused by the electron-withdrawing acyl groups.
Scheme 2. Synthetic Routes to 1 and 2.
Spectroscopic data of the complexes are summarized in Table 1. 1H NMR spectra of the mononuclear complexes 2a and 2b contain the methyl signals for the ester and acetyl groups at 3.60 and 1.95 ppm, respectively. 31P{1H} NMR signals for the dppe ligands are observed at 85.0 (2a) and 83.9 (2b) ppm, which are comparable to those for the acetylides bearing the CpRu(dppe) (δp ∼ 87 ppm) and Cp*Ru(dppe) (δp ∼ 81 ppm) groups.22 Symmetric structures of the dinuclear complexes 1 in solution are confirmed by the single sets of the 1H and 31P NMR signals for the CpR (R = COOCH3, COCH3) and dppe ligands. The ν(C≡C) and ν(C=O) vibrations are observed as single bands around 2080 and 1650–1705 cm–1, respectively. In general, diethynylbenzene compounds show single ν(C≡C) vibrations due to the asymmetric coupling of the two C≡C vibrations, while those of the symmetric modes are usually silent.26 The ν(C≡C) vibrational stretching frequencies for the ester and acetyl derivatives 1–2 are virtually identical, suggesting that the electron densities at the Ru centers are similar. Molecular structures of 1–2 determined by X-ray crystallography are shown in Figure 2. The Ru atoms in 1 and 2 adopt the three-legged piano stool geometries and the Ru–C≡ distances ranging between 2.003 and 2.020 Å are comparable to that for CpRu(dppe)–C≡C–C6H5 (4; Ru–C: 2.001 Å, Ru–P: 2.238 Å, Ru–Ccp: 2.257 Å).27,28 However, the slightly longer Ru–P bonds (Δd = 0.009–0.019 Å) and the shorter Ru–Ccp bonds (Δd 0.005–0.018 Å) compared to those in 4 should be due to the decreased and increased back donation from the ruthenium centers, respectively. The Cpcentroid–Ru–Ru–Cpcentroid dihedral angles are 180° (p-1a/b), 117° (m-1a), and 125° (m-1a). The through-space Ru–Ru distances are determined for the para- and meta-isomers to be ∼12.1 and ∼10.7 Å, respectively.
Table 1. Selected Spectroscopic Data for 1–3.
|
1H NMR/δHa |
IR/cm–1e |
||||
|---|---|---|---|---|---|
| complex | C5H4-Rb | CH3 in Rc | 31P NMR/δPa,d | ν(C≡C) | ν(C=O) |
| p-1a (R = COOMe) | 5.44, 4.57 | 3.43 | 84.9 | 2081 | 1703 |
| p-1b (R = COMe) | 5.51, 4.52 | 1.90 | 83.9 | 2081 | 1650 |
| m-1a (R = COOMe) | 5.48, 4.58 | 3.41 | 84.9 | 2075 | 1703 |
| m-1b (R = COMe) | 5.53, 4.53 | 1.92 | 84.1 | 2074 | 1650 |
| 2a (R = COOMe) | 5.47, 4.60 | 3.47 | 85.0 | 2085 | 1703 |
| 2b (R = COMe) | 5.54, 4.54 | 1.95 | 83.9 | 2085 | 1650 |
| 3a (R = COOMe) | 5.24, 4.14 | 3.60 | 79.1 | 1705f | |
| 3b (R = COMe) | 5.30, 4.14 | 1.95 | 78.2 | 1653f | |
In CDCl3.
Multiplet signals, 4H for each signal of 1 and 2H for each signal of 2 and 3
Singlet signals (3H).
1H-decoupled spectra, singlet signals.
In CH2Cl2.
As KBr pellets.
Figure 2.
ORTEP diagrams of mono- (2) and di-nuclear complexes (1) drawn with the thermal ellipsoids at 50% level. Hydrogen atoms and solvents are omitted for clarity.
Electrochemical Study
Redox properties of the Ru complexes were examined by cyclic voltammetry (Figure 3 and Table 2). CV charts of the mononuclear complexes 2 contain single partially reversible redox waves with half-wave potential of around 40 mV (vs Cp2Fe/Cp2Fe+ couple), which are anodically shifted by 120 mV compared to that of Cp(dppe)Ru–C≡C–C6H54, as a result of the decreased electron densities at the metal centers caused by the electron-withdrawing acyl substituents on the Cp rings. The E1/2 values of 2a,b are almost identical, suggesting the comparable electron-withdrawing effects of the ester and acetyl groups. For the dinuclear para-isomers p-1a,b, two fully reversible waves around −190 and 70 mV (vs FeCp2/FeCp2+ couple, 100 mV/s) are observed. The first oxidation processes are shifted cathodically compared to the mononuclear complexes 2 due to the electronic interactions between the two ruthenium centers and also inductive effects due to the attachment of the two electron-rich CpRu(dppe) donors to the bridging ligand. From the separation of the two redox waves (ΔE), the comproportionation constants (KC) are determined to be 1.35 × 104 (p-1a) and 9.00 × 103 (p-1b). In contrast to the para-isomers p-1, the meta-isomers m-1 show two irreversible redox waves in the range of 0–400 mV when scanned at 100 mV/s.
Figure 3.
Cyclic voltammograms of 1–2 ([complex] = ∼1.0 mM in CH2Cl2, [NBu4][PF6] = 0.1 M, W.E. and C.E. Pt, R.E. Ag/Ag+, scanned at 100 mV/s).
Table 2. Selected Electrochemical Data for 1–2 and Related Complexesa.
| compound | E1/21/mV [ipc/ipa] | E1/22/mV [ipc/ipa] | ΔE/mV | KC |
|---|---|---|---|---|
| p-1a | –190 (−195) [1.0] | 55 (50) [1.0] | 245 (245) | 1.35 × 104 |
| p-1b | –160 (−165) [1.0] | 70 (65) [1.0] | 230 (230) | 9.0 × 103 |
| m-1a | 21 (5) [0.4] | (170) [0.4] | (165) | 6.9 × 102 |
| m-1b | 75 (55) [0.5] | 238 (225) [0.3] | 163 (170) | 8.4 × 102 |
| 2a | 42 (25) [0.6] | |||
| 2b | 45 (41) [0.9] | |||
| RuC≡CC6H5 (4)15 | –80 | |||
| (μ-C≡C–p-C6H4–C≡C) Ru2 (5)15 | –320 | –90 | 230 | 9.0 × 103 |
| (μ-C≡C–m-C6H4–C≡C) Ru2* (6)23 | –360 | –200 | 160 | 6.7 × 102 |
Data obtained by DPV measurements are shown in parentheses. Ru = Ru(dppe)Cp, Ru* = Ru(dppe)Cp*. Potentials vs Cp2Fe/Cp2Fe+ couple.
Upon increasing the scan rate to 1000 mV/s, reversibility of the waves was improved. When the range of −500 to 200 mV was swept at a rate of 100 mV/s, partially reversible 1e-oxidation processes were observed, indicating that the monocationic species of the meta-isomers m-1a,b are stable on the timescale of the electrochemical measurements. Then, differential pulse voltammetry (DPV) measurements (see the Supporting Information) was carried out to obtain the KC values of 6.9 × 102 (m-1a) and 8.4 × 102 (m-1b), which are considerably smaller than those of the para-isomers p-1a,b due to the connection through the cross-conjugated meta-linkages. In spite of the small KC values for the meta-isomers, the monocationic species turned out to be populated enough in solutions.
The redox potentials and KC values of the ester and acetyl derivatives p-1 (KC ∼ 1 × 104) and m-1 (KC ∼ 7–8 × 102) are very similar, suggesting that the electronic interactions between the metal centers are akin to each other. The KC values of the dinuclear para-isomers p-1 are comparable to that of the related Cp analogue [p-5, (μ-p-C≡C–C6H4–C≡C){RuCpR(dppe)}2]. Similarly, the KC values of the meta-isomers m-1 are very close to that of the Cp* derivative 6 [μ-m-C≡C–C6H4–C≡C){RuCp*(dppe)}2].28,29 These facts suggest that the thermodynamic stability of the MV species of the dinuclear complexes is rather insensitive to the electron-donating properties of the substituted Cp rings.
Spectroelectrochemical IR Study
Spectroelectrochemical IR studies were performed for CH2Cl2 solutions in the presence of 0.1 M [NBu4][PF6] as an electrolyte by using an OTTLE cell (Figures 4 and 5 and Table 3). We first examined the mononuclear reference complexes 2a,b, which showed single C=O stretching vibrations around 1700 cm–1 (for the ester groups) and around 1650 cm–1 (for the acetyl groups) (Figure 4a,b). In addition, the intense bands observed around 2100 cm–1 were unequivocally assigned to the C≡C stretching vibrations. Upon oxidation of 2, the C≡C and C=O bands were replaced by the bands at around 1930 cm–1 [ν(C≡C)] and 1728 (for 2a) and 1686 cm–1 (for 2b) [ν(C=O)] with the decreased intensities. It is noted that isosbestic points were observed, indicating the occurrence of a two-component system involving direct conversions from 2 to 2+. In addition to the significant shifts of the C≡C stretching bands to lower energies (Δν ∼ 150 cm–1), the C=O bands also shifted to higher energies by 25 cm–1 for 2a (1728 cm–1) and 36 cm–1 for 2b (1686 cm–1) (Figure 5a,b), suggesting that the carbonyl substituents are sensitive to the electron densities at the ruthenium centers.
Figure 4.

IR spectral changes of 1 and 2 observed in an OTTLE cell upon application of bias voltages ([complex] ∼ 10 mM, 0.1 M [NBu4][PF6] in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+). The spectra for the three oxidation states are drawn in red (neutral), blue (monocation), and green (dication). The isosbestic points are indicated with the arrows in the insets.
Figure 5.

Parts of IR spectra [ν(C=O) region] for 1n+ and 2n+ (n = 0–2) ([complex] = ∼10 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+). The spectra for the three oxidation states are drawn in red (neutral), blue (monocation), and green (dication).
Table 3. IR Data for 1n+–2n+ (n = 0–2) Obtained with an OTTLE Cella.
| compound | n | ν(C≡C)/cm–1 | ν(C=O)/cm–1 |
|---|---|---|---|
| 2an+ | 0 | 2085 (s) | 1703 (s) |
| 1 | 2081 (w), 1934 (w) | 1728 (s) | |
| 2bn+ | 0 | 2085 (s) | 1650 (s) |
| 1 | 2083 (w), 1932 (w) | 1686 (s) | |
| p-1an+ | 0 | 2081 (s) | 1703 (s) |
| 1 | 2073 (w), 2009 (s), 1972 (s), 1920 (w) | 1712 (s) | |
| 2 | 2001 (w), 1938 (w) | 1727 (m) | |
| p-1bn+ | 0 | 2081 (s) | 1651 (s) |
| 1 | 2072 (w), 2009 (s), 1969 (s), 1919 (w) | 1667 (s) | |
| 2 | 1995 (w), 1938 (w) | 1685 (m) | |
| m-1an+ | 0 | 2075 (s) | 1703 (s) |
| 1 | 2073 (s), 2001 (w) | 1704 (m), 1726 (m) | |
| 2 | 2072 (w), 1934 (w) | 1728 (m) | |
| m-1bn+ | 0 | 2074 (s) | 1650 (s) |
| 1 | 2071 (w), 2001 (w) | 1651 (m), 1686 (m) | |
| 2 | 1931 (w) | 1686 (m) |
([Complex] = ∼10 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+).
Encouraged by these results, the spectroelectrochemical IR measurements were extended to the dinuclear complexes (1). For the para-isomers p-1a,b (Figures 4c and 5c,d), two-step oxidation processes were observed. Upon application of bias voltages, the weak ν(C≡C) vibrations of the neutral species p-1a,b were replaced by the intense bands around 1970–2020 cm–1 and the weak bands around 2070 and 1920 cm–1, which were assignable to p-1a,b+. The four ν(C≡C) vibrations observed for p-1a,b+ are similar to those reported for the Cp* analogues and are derived from the occurrence of rotamers (Figure 1).30 Further oxidation caused the second process resulting in the replacement of the bands around 2000 cm–1 by the new weak bands around 1940 cm–1, which were ascribed to the dioxidized species p-1a,b2+. For the first step, similar observations were reported for the monocationic dinuclear ruthenium complexes with the Cp* and Cp rings.20,28 The large changes of the intensities may be caused by the coupling of two C≡C bonds. Along with the changes of the ν(C≡C) bands, the C=O vibrations were also shifted in a stepwise manner (Figure 5c). For p-1a, the ν(C=O) bands appeared at 1703 (p-1a), 1712 (p-1a+), and 1727 cm–1 (p-12+). Similarly, for p-1b, the sequential and stepwise shifts of the ν(C=O) band of the acetyl groups was observed: 1651 cm–1 (p-1b) → 1667 cm–1 (p-1b+) → 1685 cm–1 (p-1b2+) (Figure 5d). It is noted that the ν(C=O) vibration energies of the neutral and dicationic species for p-1a,b are very similar to those of the neutral and monocationic species of the mononuclear complexes 2a,b, suggesting that the electron densities at the ruthenium centers in the neutral and cationic species are similar with each other. Interestingly, during the oxidation processes, the intensities of the ν(C≡C) bands drastically changed, while those of the ν(C=O) bands were not affected significantly. This is because the ν(C≡C) vibration is symmetry forbidden, and the degree of charge localization/delocalization strongly affects the symmetry of the molecules, leading to the drastic changes of the intensity. In addition, vibrations of the C≡C bonds in a molecule (such as 2) may couple to show, for example, symmetrically and antisymmetrically coupled vibrations and make analysis difficult. Thus, analysis based on the ν(C=O) bands provides another merit.
Spectroelectrochemical IR measurements of the meta-isomers m-1 were also performed (Figures 4e,f and 5e,f), and the following changes were noted for ν(C≡C) [∼2075 cm–1 (m-1) → ∼2000 cm–1 (very weak: m-1+) → ∼1931, 2000, 2070 cm–1 (m-12+)] and for ν(C=O) [1703 cm–1 (m-1a) → 1726, 1704 cm–1 (m-1a+) → 1728 cm–1 (m-1a2+); 1650 cm–1 (m-1b) → 1651, 1686 cm–1 (m-1b+) → 1686 cm–1 (m-1b2+)]. The ν(C=O) bands of m-1a+ were overlapped with those of m-1a and m-1a2+.
To obtain electron transfer rate constants (ket) of the MV species, simulation with the VIBEX GL program, which is based on the Bloch equation, was performed (Figures 6 and 7).7,31 As shown in Figure 7, upon changing the ket values in the range between 1011 and 1013 s–1, the shapes of the simulated spectra apparently changed, and the ket values were obtained by the fitting analysis of the experimental and calculated spectra. The band shapes are successfully reproduced for p-1a+, p-1b+, and m-1a+, although the shoulder band at 1728 cm–1 should be taken into account for p-1a+.32 Overlapping with the stretching bands of the Ph groups hindered the line fitting analysis of m-1b+.
Figure 6.

Experimental and simulated IR spectra [ν(C=O) region] for 1+ (red dashed lines: the sum of the simulated curves, green dashed lines: individual simulated curves, and blue lines; experimentally obtained curves).
Figure 7.

Band shape analysis of the ν(C=O) stretching vibrations for 1+.
The ket values of p-1a+ and p-1b+ are determined to be 6 × 1012 and 6 × 1012 s–1, respectively. The ket value for m-1a+ was estimated to be at most 1 × 1010 s–1 because in the ket region below 1 × 1010 s–1, the spectral change is rather insensitive to ket, and therefore, ket cannot be determined accurately on the basis of the line shape analysis.
Spectroelectrochemical UV–Vis–NIR Study
To determine electronic coupling Vab on the basis of the NIR spectra of 1+, spectroelectrochemical measurements in an OTTLE cell were performed for CH2Cl2 solutions of 1 containing 0.1 M [NBu4][PF6] as an electrolyte (Figure 8 and Table 4). For the neutral species 1, UV absorption bands were observed around 290–360 nm, which were characteristic π–π* bands admixed with ligand-to-metal charge transfer bands.20 Upon application of the bias voltage to the para-isomers p-1a,b, the UV bands disappeared and new bands assignable to the monocationic species p-1a,b+ appeared in the visible (500 nm, π–π*) and NIR regions (1000–2000 nm) (for the assignments, see below). Upon increment of the bias voltage, the bands due to p-1a,b+ were replaced by the new bands around 750 nm, which were ascribed to the dicationic species p-1a,b2+. On the other hand, the absorption band for the neutral meta-species m-1a at 298 nm was weakened upon 1e-oxidation and the new bands for m-1a+ appeared at 796 and 1326 nm. Further oxidation led to the replacement of the NIR band by that at 1460 nm (for m-1a2+) together with the increase of the intensity of the band at 796 nm, suggesting sequential generation of the mono- and di-cationic species in accord with the IR measurements discussed above. For m-1b, we could not obtain reproducible results probably due to the instability of its cationic species, and thus, in the following section, discussion for the meta-isomers will be focused on m-1a.33
Figure 8.
UV–vis–NIR spectra of (a) p-1an+, (b) p-1bn+, and (c) m-1an+ (n = 0–2) obtained with an OTTLE cell ([p-1a/b] = ∼0.1 mM, [m-1a] = ∼1 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+).
Table 4. UV–Vis–NIR Data for p-1n+ and m-1an+ (n = 0–2) Obtained with an OTTLE Cella.
| compound | absorption maxima/nm (ε × 10–4/cm–1 M–1) |
|---|---|
| p-1a | 356 (4.14) |
| p-1a+ | 530 (3.00), 1242 (2.12), 1542 (2.37), |
| p-1a2+ | 742 (4.33) |
| p-1b | 348 (4.28) |
| p-1b+ | 532 (3.01), 1266 (2.23), 1564 (2.58) |
| p-1b2+ | 750 (4.83) |
| m-1a | 298 (3.11) |
| m-1a+ | 796 (0.197), 1326 (0.127) |
| m-1a2+ | 756 (0.305), 1460 (0.208) |
[p-1a/b] ∼ 1 mM, [m-1a] = ∼10 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+.
IVCT bands of organometallic MV complexes in the NIR region are frequently observed as overlapped multiple bands.30 Therefore, the deconvolution analysis of the bands is essential to obtain Vab values (Figure 9). For p-1a+, the NIR bands were deconvoluted into the three Gaussian curves (bands A–C) as summarized in Table 5. For p-1b+, essentially the same result as p-1a+ was obtained. Similarly, the IVCT bands of m-1a+ were deconvoluted into the two bands (bands A and B) in a manner similar to the Cp* analogue 6+.28,29 The bands for the para-isomers p-1+ (ε ∼ 17,000–20,000 M–1 cm–1) were more intense than those for the meta-isomer m-1a+ (ε ∼ 1000–5000 M–1 cm–1). For para-isomers p-1+, the band widths calculated according to the Hush’s theory [νcalc = (2310 × νmax)1/2] were much larger than those of the bands A–C (Table 5), indicating stronger metal–metal interaction. The insensitivity of the NIR bands for para-isomers p-1a,b+ to the solvent polarity (acetone/CH2Cl2 = 98:2 vs CH2Cl2) (Figure 10) suggests the assignment of p-1+ as a class II/III border or class III species.34 On the other hand, the value for the meta-isomer m-1a+ was comparable to the theoretical value, suggesting that m-1a+ is a class II species.
Figure 9.
NIR spectra of (a) p-1an+, (b) p-1bn+, and (c) m-1an+ obtained with an OTTLE cell and their deconvoluted spectra ([p-1a/b] = ∼1 mM, [m-1a] = ∼10 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+).
Table 5. NIR Spectral Data for p-1+ and m-1a+ and Related Dataa.
| deconvoluted
IVCT bands |
||||||
|---|---|---|---|---|---|---|
| compound | band | νmax/cm–1 | εmax/M–1 cm–1 | ν1/2calcb/cm–1 | ε/M–1 cm–1 | Vab (classIII)c/Vab (classII)d/cm–1 (dRu–Ru) |
| p-1a+ | A | 6290 | 17050 | 1100 | 3812 | 3145/585 |
| B | 7857 | 19521 | 1956 | 4260 | 3929/932 | |
| C | 9851 | 4878 | 2400 | 2355 | 4926/573 | |
| p-1b+ | A | 6200 | 21138 | 1100 | 3784 | 3100/646 |
| B | 7850 | 21156 | 1871 | 4258 | 3925/949 | |
| C | 9851 | 5184 | 2400 | 2355 | 4926/591 | |
| m-1a+ | A | 5500 | 1457 | 3560 | 3564 | 288 (12.1 Å)e, 323 (10.8 Å)f |
| B | 7760 | 5015 | 3500 | 4234 | 623 (12.1 Å)e, 698 (10.8 Å)f | |
[p-1a/b] = ∼1 mM, [m-1a] = ∼10 mM, [NBu4][PF6] = 0.1 M in CH2Cl2, W.E. and C.E. Pt, and R.E. Ag/Ag+.
ν1/2calc = (2310 × νmax)1/2.
Vab(class III) = νmax/2.
Vab(class II) = 2.06 × 10–2 × (νmax εmax ν1/2)/dRu–Ru.
Through bond distances.
Through space distances.
Figure 10.

Normalized NIR spectra of (a) p-1a+ and (a) p-1b+ observed in CH2Cl2 (solid blue lines) and a mixed solvent of acetone/CH2Cl2 (98:2) (dashed red lines).
DFT and TD–DFT Calculations
In order to clarify the electronic structures of the MV species and to assign the NIR bands (Figures 11–13), DFT and TD-DFT calculations on 1n+ and their Cp derivatives, p-5n+ and m-5n+ ([(μ-p/m-C≡C–C6H4–C≡C){RuCp(dppe)}2]n+) (n = 0, 1), were performed. The structural optimization of 1 was started with the antiparallel conformations observed by the X-ray structural analysis. The Kohn–Sham orbitals of the neutral species are summarized in Figure 11. Bond lengths and angles well coincide with those obtained by the X-ray analysis. The highest occupied molecular orbitals (HOMOs) of the para-isomers p-1a,b are delocalized over the Ru–C≡C–C6H4–C≡C–Ru frameworks, whereas the lowest unoccupied molecular orbitals (LUMOs) are localized on the dppe ligands. These results are in accord with the results for the related dinuclear bridging acetylide complexes reported previously.20 Similar features are noted for the meta-isomers m-1a,b. The HOMOs for the para-isomers p-1 are higher in energy by about 0.3 eV compared to those of the meta-isomers m-1, and the result is consistent with the trend of their redox potentials. On the other hand, the energy levels of the LUMOs are less sensitive to the connectivity of the central ring because the dppe ligands mainly constitute them. The HOMO and LUMO features for p/m-1a are very similar to those for p/m-5. The HOMO and LUMO levels of p/m-5 are higher by about 0.2 and 0.4 eV, respectively, than those of the acyl derivatives 1, and the differences should derive from the electron-withdrawing acyl substituents.
Figure 11.

Frontier orbitals of p-1a/b, m-1a,b, and p/m-5 and their energies.
Figure 13.
β-HOSO and β-LUSO orbitals for 1+.
The spin density plots for the para-isomers of the monocationic species p-1+ reveal the full and symmetrical charge-delocalization over the Ru–C≡C–C6H4–C≡C–Ru linkages (Figure 12), suggesting the class III nature. On the other hand, the spin density of the meta-isomers m-1+ is substantially localized on one of the two ruthenium fragments, as is consistent with the charge-localized class II behavior. The spin density distribution is similar to those of p/m-5+, indicating again that the acyl substituents do not strongly influence the charge delocalization.
Figure 12.
Spin density plots of p-1+, m-1+, and p/m-5+. The α and β spin densities are represented in blue and red, respectively.
TD-DFT analysis of 1+ shows that the lowest energy bands arise from the transitions from β-HOSOs (the highest occupied spin orbitals) to β-LUSOs (the lowest unoccupied spin orbitals) (Figure 13 and Table 6). The computed energies well coincide with the experimental data, although the sub-bands (bands B and C) of para-isomers 1+ cannot be reproduced because different rotational isomers may be present.20 For the para-isomers p-1a,b+, both of the β-HOSO and β-LUSO show the fully delocalized Ru–C≡C–C6H4–C≡C–Ru character, which is in accord with the assignment as a class III IVCT transition. On the other hand, the β-HOSOs and β-LUSOs of m-1a,b+ show clear charge-transfer characters from the dπ and pπ orbitals located on one of the Ru–C≡C–C6H4 fragments to those on the other metal fragment. These transition characters can be assignable to typical class II type IVCT bands. On the basis of the TD-DFT results and the assignment of the lowest energy band to the IVCT bands, the Vab values of p-1a+, p-1b+, and m-1a+ estimated from the Marcus–Hush two state model are 3145 (p-1a+), 3100 (p-1b+) and 288 cm–1 (m-1a+).35−37
Table 6. TD–DFT Data of NIR Bands for 1+.
| complex | νmaxexp/cm–1 | νmaxDFT/cm–1 (nm) | oscillator strength | main transition | assignment |
|---|---|---|---|---|---|
| p-1a+ | 6290 | 5435 (1840) | 1.0929 | β-HOSO → β-LUSO (∼100%) | IVCT |
| p-1b+ | 6200 | 5714 (1750) | 1.035 | β-HOSO → β-LUSO (∼100%) | IVCT |
| m-1a+ | 5500 | 4675 (2135) | 0.0523 | β-HOSO → β-LUSO (74%) | IVCT |
| 7760 | 5167 (1913) | 0.0031 | β-HOSO – 3 → β-LUSO (40%) | dπ–dπ transition (ligand field) | |
| β-HOSO – 4 → β-LUSO (25%) | |||||
| m-1b+ | 5000 (2000) | 0.0118 | β-HOSO – 3 → β-LUSO (70%) | dπ–dπ transition (ligand field) | |
| 6897 (1450) | 0.0579 | β-HOSO → β-LUSO (89%) | IVCT |
ESR Study
ESR spectroscopy is another powerful tool to obtain environmental information of an unpaired electron on radical species such as the monocationic species, which was generated by in situ oxidation of the neutral species by the addition of [FeCp2](PF6) in CH2Cl2 and was subjected to the measurements as glasses after immediate freezing at 77 K (Figure 14). While intense and broadened isotropic or partially anisotropic signals with no hyperfine splitting were observed for the para-isomers p-1+,38 very weak signals were found for m-1+. This suggests for p-1+ that the unpaired electrons are delocalized over the bridging ligands on the timescale of ESR spectroscopy (∼10–8 s). The spectra are in contrast to that observed for [Cp*(dppe)Ru–C≡C–C6H5]+7+ with a rhombic g-tensor and an isotropic g value (giso = g1 + g2 + g3) of 2.09.39 The g values of p-1+ (∼2.06) are smaller than that of 7+ but much larger than those of organic free radicals (g = 2.0023). These factors further support the charge-delocalization over the Ru–C≡C–C6H4–C≡C–Ru moieties in p-1+.
Figure 14.

ESR spectra of p-1a+ and p-1b+ recorded as CH2Cl2 glasses at 77 K.
Effects of Substituents on the Cp Ligands
As expected, strong metal–metal interactions are observed for the para-isomers p-1n+ as judged by their KC, Vab, and ket values, which are significantly larger than those of the meta-isomers m-1n+ with the different connectivities at the central benzene ring. Furthermore, the electrochemical and spectroscopic features of the core parts of the Ru complexes with/without the acyl functional groups are virtually the same. In fact, the theoretical DFT and TD-DFT studies support that the charge-delocalization properties are not perturbed by the CpR ligands so much when compared to the nonsubstituted Cp derivatives, although redox potential is sensitive to the electron-withdrawing substituents. Thus, these results support that CpR ligands (R = COOMe, COMe) can be used to evaluate the electronic structure of MV complexes without drastic electronic perturbation.
Charge Distribution Parameter Δρ
The charge distribution parameter, Δρ, which was first introduced by Geiger and Gleiter, provides a degree of charge delocalization between two IR-active centers in mixed valence species (MV) of a compound × [2·X(MV) ⇄ X+(ox) + X– (red)].12 Δρ is defined by eq 1
| 1 |
where Δνox = νox – νoxMV, Δνred = νred – νred. In the present IR system, νox and νred represent vibrational frequencies of the di-oxidized and neutral species, respectively, and νoxMV and νred represent those of the higher-energy and lower-energy bands of the mono-oxidized species, respectively (Figure 15). When Δρ = 0 or 0.50, the complex is a localized class I species or a delocalized class III species, respectively. A compound with an intermediate Δρ value (0 < Δρ < 0.50) is categorized into class II. According to the equation, the p-isomers p-1a,b+ with Δρ = 0.50 turned out to be class III species, whereas the m-isomer m-1a+ with Δρ = 0.06 is characterized as a class II species and the charge is localized on one of the two metal fragments (Table 7). Due to the overlapped bands for m-1b+, its Δρ value cannot be determined accurately, but the value appears to be comparable to that of m-1a+. When the Δρ values are compared to the spin densities at the Ru atoms obtained by the DFT calculations (see above), they are well consistent with the charge distribution estimated by the Δρ values. Thus, the IR-tag serves as a practical indicator of charge delocalization over MV species.
Figure 15.
Schematic representation of IR stretching vibrations of a mixed valence system [2·X(MV) ⇄ X+(ox) + X– (red)].
Table 7. Parameters Derived from IR, UV–Vis–NIR, and DFT Studies.
| compound | Δνox/cm–1 | Δνred/cm–1 | νox – νred | Δρa | charge distribution on the two metal centers (%) | spin densities at the two metal centersb (%) | ΔG*/cm–1 | ket/s–1 |
|---|---|---|---|---|---|---|---|---|
| p-1a+ | 9 | 15 | 24 | 0.50 | 50:50 | 50:50 | 6 × 1012 | |
| p-1b+ | 16 | 18 | 34 | 0.50 | 50:50 | 50:50 | 6 × 1012 | |
| m-1a+ | 1 | 2 | 25 | 0.06 | 94:6 | 100:0 | 1100 | <1.0 × 1010 |
| m-1b+ | 1 | 0 | 36 | c | c | 100:0 | <1.0 × 1010 |
ρ = (Δνox + Δνox)/0.5(νox – νred).
Estimated from the DFT calculations.
Not determined due to overlapped bands.
Comparison of ket and Vab Values
Electron transfer rates of MV complexes are sensitive to the environments. As the rate of solvent reorientation timescale is considered to be in the range from 10–11 to 10–12 s, IVCT bands of class II species should be affected by the polarity of the solvent. Therefore, ket of class II and class III species falls below (<10–11 s) and above (>10–12 s) this range, respectively. As discussed above, the NIR bands of the para-isomers p-1+ showed little dependence on the solvents, and their ket values turned out to be ∼6 × 1012 s–1, supporting their class III nature.
On the other hand, the ket values of the meta-isomers m-1+ are less than 1010 s–1 in accord with the assignment as class II classification. For a weakly coupled class II system, the ket value can also be evaluated by using the following equations in case the electronic coupling Vab is small and the thermal ET process passes through the diabatic regime.40
| 2 |
| 3 |
where ΔG* is the free energy of activation needed for the thermally induced ET transfer, νmax is the energy of the IVCT transition band at the maximum extinction coefficient, h is Planck’s constant, kB is Boltzmann’s constant, and T is the temperature. On the basis of eqs 2 and 3, ΔG* and ket for m-1a+ are determined to be 1100 cm–1 and 1.26 × 1011 s–1, respectively. The ket value is overestimated by an order of magnitude compared to that obtained by the IR methods (<1.0 × 1010 s–1). The deviation may derive from (i) the erroneous estimation of Vab values, which are determined by the deconvolution analysis of overlapped bands, (ii) the ground-state energy curves deviating from the ideal diabatic states, or (iii) the presence of rotamers as discussed above.
Scope and Limitation of the IR-Tag Strategy on the Cp Ligand
As we have discussed above, the application of the IR-tags strategy toward organometallic MV complexes with Cp ligands is useful, especially, for the estimation of the local charges on the metal fragments and of the ket values obtained by the digital simulations. On the other hand, the following limitation was noted in this study. The ket values can be determined only when the values are within the range that the simulation allows (10–10 to 10–12 s). Furthermore, in some case, the C=O vibrational bands can be overlapped with the other bands to hamper the analysis. Nevertheless, the present IR-tag method is promising because it can be facilely applied to MV complexes with various types of metal species, Cp ligands, and supporting ligands.
Conclusions
We have synthesized mono- and di-ruthenium alkynyl complexes with cyclopentadienyl ligands bearing the IR-active acyl groups and examined their electronic structures by electrochemical, spectroelectrochemical UV–vis–NIR and IR, ESR, and theoretical methods. It is clear that the para-isomers p-1 show metal–metal interactions stronger than the meta-isomers m-1. In contrast to the determination process of the Vab value involving ambiguous factors, ket can be unambiguously determined by the line shape analysis of IR spectra of the acylated Cp ligands attached to the metal centers and thus helps us fully understand the whole pictures of the electronic structures of the metal parts in MV states. The ν(C=O) frequencies of the acyl groups are sensitive to the electron densities at the metal centers and, thus, afford information on the extent of charge-delocalization. As described herein, we have successfully extended the IR-tag technique toward metal alkynyl MV complexes with half-sandwich cyclopentadienyl structural motifs, which have been extensively studied for these 3 decades. This study should provide new insights into the electronic structures of metal alkynyl MV complexes.
Experimental Section
Instruments
NMR spectra were recorded on Bruker biospin AVANCE III 400 MHz (1H 400 MHz for 2D NMR measurements) and Bruker biospin ASCEND-500 spectrometers (1H 500 MHz, 31P 202 MHz 13C NMR 126 MHz) in CDCl3. NMR chemical shifts were referenced to the residual protio impurities in the deuterated solvents. ESI-TOF-MS analysis was performed on a Bruker micrOTOF II spectrometer. UV–vis and IR spectra (KBr pellets and CH2Cl2 solution) were recorded on JASCO V670DS and FTIR 4200 spectrometers, respectively. Electrochemical measurements (CV and DPV) were made with a Hokuto Denko HZ-5000 [observed in CH2Cl2; [complex] = ca. 1 × 10–3 M; [NBu4PF6] = 0.1 M; working electrode: Pt, counter electrode: Pt, reference electrode: Ag/AgNO3; and scan rate was 100 mV/s (CV)]. After the measurement, ferrocene (Fc) was added to the mixture and the potentials were calibrated with respect to the Fc/Fc+ redox couple. ESR spectra for the monocationic species were recorded on a JEOL JES-FA100 spectrometer, and samples were prepared by the addition of small amounts of oxidant [tris(4-bromophenyl)aminium hexachloroantimonate] to CH2Cl2 solutions of the complexes in ESR tubes under a N2 atmosphere just before cooling in a liquid nitrogen bath. The spectra were recorded at the liquid nitrogen temperature. X-ray crystal analysis was performed on a Bruker Smart Apex II Ultra diffractometer. CCDC numbers 1861281–1861286 contain the supplementary crystallographic data for p-1a (1861281), p-1b (1861282), m-1a (1861283), m-1b (1861284), 2a (1861285), and 2b (1861286).
Spectroelectrochemistry
All measurements were conducted under a N2 atmosphere. Spectroelectrochemical UV–vis–NIR measurements were performed in a home-built 1 mm quartz cell equipped with working (Pt mesh), counter (Pt wire), and reference electrodes (Ag/AgNO3). A CH2Cl2 solution of complex (0.1 mM for p-1a,b, 1 mM for m-1a,b) was subjected to the measurements in the presence of NBu4PF6 (0.1 M), a supporting electrolyte. The spectroelectrochemical IR measurements were performed with an OTTLE cell, purchased from the University of Reading, which comprised a Pt-mesh working and counter electrodes and a thin silver wire as a pseudo-reference electrode. The optical beam was passed through CaF2 windows of a conventional liquid IR cell. The conditions were the same as those for the spectroelectrochemical UV–vis–NIR measurements except the concentration of the complexes (10 mM for 1).
Materials
Reactions were performed under a N2 atmosphere using the standard Schlenk tube technique unless stated otherwise. Purchased dry THF, CH2Cl2, and hexane were further purified by the Grubbs solvent purification system.41 Dry acetone purchased from Kanto Chemical Co., Inc. was used without further purification. Triethylamine was predried over KOH pellets and distilled from CaH2. DBU was distilled from CaH2. Other regents, silica gel (Kanto chemical Co Inc. Silica Gel 60N), and alumina (Merck Aluminum oxide 90 standardized) were used as received. RuCl2(PPh3)3,42 sodium methoxycarbonylcyclopentadienide,24 and sodium acetylcyclopentadienide24 were synthesized according to the literature procedures.
Synthesis of 3a
A THF (50 mL) solution of RuCl2(PPh3)3 (3.00 g, 3.13 mmol) and sodium methoxycarbonylcyclopentadienide (487 mg, 3.33 mmol) was stirred for 1 h at room temperature. The mixture was concentrated and hexane (80 mL) was added to form red precipitates, which were collected by filtration, washed with hexane, and dried under reduced pressure to afford crude products of 3a. To the obtained solid was added dppe (1.247 g, 3.13 mmol) and toluene (50 mL), and the mixture was refluxed for 4 h. After the mixture was cooled to room temperature, the volatiles were evaporated, and the resulting solid was subjected to silica gel column chromatography (CH2Cl2/Et2O = 9:1) to yield 3a (824 mg, 1.25 mmol, 40%) as a yellow powder. 1H NMR: δ 7.83–7.76 (m, 4H, Ph in dppe), 7.40–7.36 (m, 6H, Ph), 7.33–7.26 (m, 6H, Ph), 7.12 (virtual t, J = 8.7 Hz, 4H, Ph), 5.24 (t, J = 2.0 Hz, 2H, η5-C5H4-R), 4,14 (t, J = 2.0 Hz, 2H, η5-C5H4-R), 3.60 (s, 3H, COOCH3), 2.78–2.61 (m, 2H, CH2), 252–2.35. (m, 2H, CH2). 31P NMR: δ 79.1 (s). 13C{1H} NMR: δ 167.8 (s, COOMe) 140.5 (virtual dt, JC–P = 43.6, 12.9 Hz, ipso-Ph in dppe), 134.2 (t, JC–P = 4.9 Hz, o-Ph in dppe), 134.5–133.8 (m, ipso-Ph in dppe), 131.7 (t, JC–P = 4.8 Hz, o-Ph in dppe), 130.0 (s, p-Ph in dppe), 129.5 (s, p-Ph in dppe-Ph), 128.3 (t, JC–P = 4.4 Hz, m-Ph in dppe), 128.1 (t, JC–P = 4.7 Hz, m-Ph in dppe), 89.1, 79.7, 76.7 (s × 3, η5-C5H4-R), 51.5 (s, COOMe), 27.5 (m, CH2 in dppe). HRESI-TOF MS (m/z): calcd for C33H31P2O2Ru, 623.0846. Found: 623.0845 [M–Cl]+. IR(KBr/cm–1): 1705 ν(C=O). Anal. Calcd for C32.1H31.2Cl1.2O2P2Ru (3a· (CH2Cl2)0.1): C, 58.90; H, 4.80. Found: C, 59.02; H, 4.28.
Synthesis of 3b
Synthesis of 3b was previously reported,24 but we prepared 3b in a manner analogous to the synthesis of 3a using RuCl2(PPh3)3 (3.00 g, 3.13 mmol), sodium acetylcyclopentadienide (429 mg, 3.30 mmol), and dppe(1.247 g, 3.13 mmol). 3b was obtained as a yellow powder (987 mg, 1.54 mmol, 49%). The spectral data were consistent with the value reported previously. 1H NMR: δ 7.74–7.68 (m, 4H, o-Ph in dppe), 7.42–7.38 (m, 6H, Ph), 7.34 (t, J = 7.3 Hz, 2H, Ph), 7.30 (t, J = 7.3 Hz, 4H, Ph), 7.12 (virtual t, J = 8.0 Hz, 4H, o-Ph in dppe), 5.30 (m, 2H, η5-C5H4-R), 4.18–4.10 (m, 2H, η5-C5H4-R), 2.78–2.60 (m, 2H, CH2 in dppe), 2.42–2.23 (m, 2H, CH2 in dppe), 1.95 (s, 3H, COCH3). 31P NMR: δ 78.2 (s).13C{1H} NMR: δ 196.5 (s, COMe), 139.5 (virtual dt, JC–P = 43.8, 12.6 Hz, ipso-Ph in dppe), 133.6 (t, JC–P = 5.0 Hz, o-Ph in dppe), 131.8 (t, JC–P = 5.0 Hz, o-Ph in dppe), 129.9 (s, p-Ph in dppe), 129.7 (s, p-Ph in dppe), 128.2 (t, JC–P = 4.8 Hz, m-Ph in dppe), 127.9 (t, JC–P = 5.0 Hz, m-Ph in dppe), 88.4, 88.0, 76.7 (s × 3, η5-C5H4-R), 28.2 (s, COMe), 26.7 (m, CH2 in dppe). IR(KBr/cm–1): 1653 ν(C=O).
General Procedure for the Synthesis of Alkynylruthenium Complexes
To a mixture of CpRRuCl(dppe) (1 equiv, 0.15–0.5 mmol scale) and AgBF4 (1 equiv) was added CH2Cl2 (5–10 mL), and the resultant mixture was stirred at room temperature for 30 min. The suspension was filtered through a Celite pad to remove precipitated AgCl and the filtrate was added to a solution of the appropriate ethynylbenzene (1 and 0.5 equiv for mono- and di-nuclear complexes, respectively) in CH2Cl2 (5–10 mL). After the mixture was stirred for 1 h, several drops of DBU (<∼50 μL) was added to the reaction mixture, which was further stirred for 1 h. The solution was evaporated and the obtained residue was subjected to silica gel column chromatography (CH2Cl2) to afford the product.
Synthesis of 2a
The title complex was obtained from 3a (100 mg, 0.152 mmol) and 4-ethynyltoluene (17 mg, 0.15 mmol) as a yellow powder (83 mg, 0.11 mmol, 74%). 1H NMR: δ 7.89–7.83 (m, 4H, o-Ph in dppe), 7.42–7.37 (m, 6H, Ph), 7.34–7.21 (m, 10H, Ph), 6.74 (d, 2H, J = 7.9 Hz, C6H5CH3), 6.39 (d, 2H, J = 7.9 Hz, C6H5CH3), 5.49 (t, J = 2.2 Hz, 2H, η5-C5H4-R), 4.60 (t, J = 2.2 Hz, 2H, η5-C5H4-R), 3.48 (s, 3H, C(O)OCH3), 2.80–2.63 (m, 2H, CH2 in dppe), 2.40–2.23 (virtual octet, J = 6.8 Hz, 2H, CH2 in dppe), 2.18 (s, 3H, C6H5CH3). 31P NMR: δ 85.0 (s). 13C{1H} NMR: δ 167.5 (s, COMe), 141.6 (virtual dt, JC–P = 39.2, 9.2 Hz, ipso-Ph in dppe), 136.2 (virtual td, JC–P = 27.4, 4.6 Hz, ipso-Ph in dppe), 134.2 (t, J = 5.1 Hz, o-Ph in dppe), 132.9 (s, C6H5CH3), 131.8 (t, JC–P = 5.1 Hz, o-Ph in dppe), 130.4 (s, C6H5CH3), 129.6 (s, p-Ph in dppe), 129.2 (s, p-Ph in dppe), 128.2 (s, C6H5CH3), 128.1 (t, J = 4.7 Hz, m-Ph in dppe), 127.7 (t, J = 5.0 Hz, m-Ph in dppe), 126.9 (s, C6H5CH3), 111.7 (s, Ru–C≡C), 111.0 (t, J = 25.7 Hz, Ru–C≡C), 89.3, 83.0, 82.4 (s × 3, η5-C5H4-R), 51.2 (s, COOMe), 28.4 (m, CH2 in dppe), 21.2 (s, C6H5CH3). HRESI-TOF MS (m/z): calcd for C42H38P2O2Ru, 738.1397. Found: 738.1397 [M]+. IR(CH2Cl2/cm–1): 2085 ν(C≡C), 1703 ν(C=O). Anal. Calcd for C42H38O2P2Ru: C, 68.38; H, 5.19. Found: C, 68.38; H, 5.24.
Synthesis of 2b
The title complex was obtained from 3b (100 mg, 0.156 mmol) and 4-ethynyltoluene (17 mg, 0.15 mmol) as a yellow powder (63 mg, 0.087 mmol, 58%). 1H NMR: δ 7.77–7.71 (m, 4H, o-Ph in dppe), 7.42–7.38 (m, 6H, Ph), 7.33 (t, J = 7.7 Hz, 2H, p-Ph in dppe), 7.28 (t, J = 7.0 Hz, 4H, Ph), 7.23 (m, 4H, Ph), 6.80 (d, J = 7.8 Hz, 2H, C6H5CH3), 6.47 (d, J = 7.8 Hz, 2H, C6H5CH3), 5.54 (t, J = 2.1 Hz, 2H, η5-C5H4-R), 4.56–4.52 (m, 2H, η5-C5H4-R), 2.80–2.66 (m, 2H, CH2 in dppe), 2.30–2.13 (m, 2H, CH2 in dppe), 2.21 (s, 3H, C6H5CH3), 1.95 (s, 3H, COCH3). 31P NMR: δ 83.9 (s).13C{1H} NMR: δ 196.1 (s, COMe), 140.6 (virtual dt, JC–P = 34.4, 11.1 Hz, ipso-Ph in dppe), 135.7 (virtual td, JC–P = 26.2, 5.4 Hz, ipso-Ph in dppe), 133.5 (t, JC–P = 4.9 Hz, o-Ph in dppe), 133.3 (s, C6H5CH3), 132.0 (t, JC–P = 5.3 Hz, o-Ph in dppe), 130.3 (s, C6H5CH3), 129.4 (s, p-Ph in dppe), 129.3 (s, p-Ph in dppe), 128.2 (s, C6H5CH3), 128.0 (t, JC–P = 4.7 Hz, m-Ph in dppe), 127.5 (t, JC–P = 5.0 Hz, m-Ph in dppe), 126.7 (s, C6H5CH3), 112.7 (t, J = 24.9 Hz, Ru–C≡C), 111.5 (s, Ru–C≡C), 90.3, 89.2, 82.6 (s × 3, η5-C5H4-R), 27.8 (m, CH2 in dppe), 27.4 (s, COMe), 21.1 (s, C6H5CH3). HRESI-TOF MS (m/z): calcd for C42H38P2ORu, 722.1447 [M]+. Found: 722.1447 [M]+. IR(CH2Cl2/cm–1): 2085 ν(C≡C), 1650 ν(C=O). Anal. Calcd for C42.3H38.6Cl0.6OP2Ru (2b· (CH2Cl2)0.3): C, 67.99; H, 5.21. Found: C, 67.77; H, 5.23.
Synthesis of p-1a
The title complex was obtained from 3a (320 mg, 0.486 mmol) and 1,4-diethynylbenzene (30 mg, 0.24 mmol) as a yellow powder (245 mg, 0.179 mmol, 75%). 1H NMR: δ 7.79 (virtual t, J = 8.5 Hz, m, 8H, o-Ph in dppe), 7.37–7.20 (m, 32H, Ph), 6.11 (s, 4H, C6H4), 5.44 (t, J = 2.1 Hz, 4H, η5-C5H4-R), 4.57 (t, J = 2.1 Hz, 4H, η5-C5H4-R), 3.43 (s, 3H, C(O)OCH3), 2.75–2.60 (m, 4H, CH2 in dppe), 2.35–2.18 (m, 4H, CH2 in dppe), 31P NMR: δ 84.9 (s). 13C{1H} NMR: δ 167.4 (s, C(O)OMe), 141.1 (virtual dt, JC–P = 39.0, 9.6 Hz, ipso-Ph in dppe), 136.1 (virtual td, JC–P = 28.0, 4.5 Hz, ipso-Ph in dppe), 134.0 (m, o-Ph in dppe), 131.6 (m, o-Ph in dppe), 129.3 (s, p-Ph in dppe), 129.0 (s, p-Ph in dppe), 127.9 (m, m-Ph in dppe), 127.5 (m, m-Ph in dppe), 112.7 (s, Ru–C≡C), 110.7 (t, J = 26.5 Hz, Ru–C≡C), 89.2 (s, η5-C5H4-R), 83.0 (s, η5-C5H4-R), 51.2 (s, C(O)OMe) 28.4 (m, CH2 in dppe). HRESI-TOF MS (m/z): calcd for C76H66P4O4Ru2, 1370.1998 [M]+. Found 1370.1995 [M]+. IR(CH2Cl2/cm–1): 2085 ν(C≡C), 1703 ν(C=O). Anal. Calcd for C76.3H66.6Cl0.6O4P4Ru2 (p-1a· (CH2Cl2)0.3): C, 65.70; H, 4.81. Found: C, 65.54; H, 4.70.
Synthesis of p-1b
The title complex was obtained from 3b (130 mg, 0.202 mmol) and 1,4-diethynylbenzene (13 mg, 0.10 mmol) as a yellow powder (61 mg, 0.045 mmol, 45%). 1H NMR: δ 7.71 (virtual t, J = 7.8 Hz, 8H, o-Ph in dppe), 7.38–7.25 (m, 24H, Ph), 7.20 (virtual t, J = 8.9 Hz, 8H, Ph), 6.25 (s, 4H, C6H4), 5.51 (t, J = 2.0 Hz, 4H, η5-C5H4-R), 4.52 (s, 4H, η5-C5H4-R), 2.79–2.60 (m, 4H, CH2 in dppe), 2.28–2.11 (m, 4H, CH2 in dppe), 1.91 (s, 6H, COCH3) 31P NMR: δ 83.9 (s). 13C{1H} NMR: δ 196.0 (s, COMe), 140.6 (virtual dt, JC–P = 39.6, 18.6 Hz, ipso-Ph in dppe), 135.7 (virtual dt, JC–P = 52.2, 25.6 Hz, ipso-Ph in dppe), 133.5 (m, o-Ph in dppe), 132.0 (m, o-Ph in dppe), 129.6–129.3 (m, p-Ph in dppe and C6H4), 128.0 (m, p-Ph in dppe), 127.5 (m, m-Ph in dppe), 124.9 (s, C6H4), 113.2 (t, J = 24.3 Hz, Ru–C≡C), 112.4 (s, C≡C), 90.3, 89.2, 82.6 (s × 3, η5-C5H4-R), 27.8 (m, CH2 in dppe), 27.4 (s, COMe) HRESI-TOF MS (m/z): calcd for C76H66P4O2Ru2, 1338.2099 [M]+. Found: 1338.2093 [M]+. IR(CH2Cl2/cm–1): 2081 ν(C≡C), 1650 ν(C=O). Anal. Calcd for C77H68Cl2O2P4Ru2 (p-1b· CH2Cl2): C, 65.02; H, 4.82. Found: C, 65.67; H, 4.70.
Synthesis of m-1a
The title complex was obtained from 3a (160 mg, 0.243 mmol) and 1,3-diethynylbenzene (14 mg, 0.11 mmol) as a yellow powder (121 mg, 0.088 mmol, 80%). 1H NMR: δ 7.80 (virtual t, J = 8.0 Hz, 8H, o-Ph in dppe), 7.32–7.17 (m, 32H, Ph), 6.54 (t, J = 7.7 Hz, 1H, C6H4), 6.20 (t, J = 1.6 Hz, 1H, C6H4), 6.10 (dd, J1 = 7.7 Hz, J2 = 1.6 Hz, 2H, C6H4), 5.48 (t, J = 2.2 Hz, 4H, η5-C5H4-R), 4.58 (t, J = 2.2 Hz, 4H, η5-C5H4-R), 3.41 (s, 6H, COOMe), 2.80–2.66 (virtual octet, J = 5.9 Hz, 4H, CH2 in dppe), 2.39–2.24 (virtual octet, J = 5.3 Hz, 4H, CH2 in dppe). 31P NMR: δ 84.9 (s). 13C{1H} NMR: δ 167.4 (s, COOMe), 141.5 (virtual td, JC–P = 26.9, 8.6 Hz, ipso-Ph in dppe), 136.4 (virtual td, JC–P = 28.7, 4.3 Hz,, ipso-Ph in dppe), 134.0 (t, JC–P = 4.9 Hz, o-Ph in dppe), 132.0 (s, C6H4), 131.9 (t, JC–P = 5.2 Hz, o-C6H5 in dppe), 129.5 (s, p-C6H5 in dppe), 129.1 (s, p-Ph in dppe), 128.9 (s, C6H4), 128.1 (t, JC–P = 4.5 Hz, m-Ph in dppe), 127.6 (t, JC–P = 5.0 Hz, m-Ph in dppe), 126.5 (s, C6H4), 126.4 (s, C6H4), 112.4 (s, C≡C), 109.7 (t, JC–P = 25.6 Hz, Ru–C≡C), 89.3, 83.1, 82.6 (s × 3, η5-C5H4-R), 51.2 (s, COOMe), 28.5 (m, CH2-dppe). HRESI-TOF MS (m/z): calcd for C76H66P4O4Ru2, 1370.1998 [M]+. Found: 1338.2003 [M]+. IR(CH2Cl2/cm–1): 2075 ν(C≡C), 1703 ν(C=O). Anal. Calcd for C76H66O4P4Ru2·0.5(CH2Cl2): C, 65.08; H, 4.78. Found: C, 64.72; H, 4.54.
Synthesis of m-1b
The title complex was obtained from 3b (130 mg, 0.202 mmol) and 1,3-diethynylbenzene (13 mg, 0.10 mmol) as a yellow powder (66 mg, 0.048 mmol, 48%). 1H NMR: δ 7.73–7.67 (m, 8H, o-Ph in dppe), 7.35–7.24 (m, 24H, Ph), 7.19 (virtual t, J = 8.1 Hz, 8H, o-Ph in dppe), 6.62 (t, J = 7.7 Hz, 1H, C6H4), 6.32 (t, J = 1.6 Hz, 1H, C6H4), 6.16 (dd, J = 7.7 Hz, J2 = 1.6 Hz, 2H, C6H4), 5.53 (t, J = 2.1 Hz, 4H, η5-C5H4-R), 4.54–4.52 (m, 4H, η5-C5H4-R), 2.82–2.65 (virtual octet, J = 6.3 Hz, 4H, CH2 in dppe), 2.32–2.15 (virtual octet, J = 6.9 Hz, 4H, CH2 in dppe), 1.91 (s, 6H, COCH3). 31P NMR: δ 84.1(s). 13C{1H} NMR: δ 195.9 (s, −COMe), 140.7 (virtual dt, JC–P = 39.7, 18.1 Hz, ipso-Ph in dppe), 135.8 (virtual dt, JC–P = 27.9, 5.9 Hz, ipso-Ph in dppe), 133.6 (t, JC–P = 4.8 Hz, o-Ph in dppe), 132.0 (t, JC–P = 5.2 Hz, o-Ph in dppe), 131.7 (s, C6H4), 129.5 (s, p-Ph in dppe), 129.4 (s, p-Ph in dppe), 128.9 (s, C6H4), 128.1 (t, JC–P = 4.5 Hz, m-Ph in dppe), 127.6 (t, JC–P = 5.0 Hz, m-Ph in dppe), 126.8 (s, C6H4), 126.7 (s, C6H4), 112.5 (t, J = 24.6 Hz, Ru–C≡C), 112.3 (s, C≡C), 90.4, 89.4, 82.9 (s × 3, η5-C5H4-R), 28.0 (m, CH2-dppe), 27.6 (s, COMe). HRESI-TOF MS (m/z): calcd for. C76H66P4O2Ru2, 1338.2099 [M]+. Found: 1338.2092 [M]+. IR(CH2Cl2/cm–1): 2074 ν(C≡C), 1650 ν(C=O). Anal. Calcd for C76.1H66.2Cl0.2O2P4Ru2 (m-1b· (CH2Cl2)0.1): C, 67.91; H, 4.96. Found: C, 67.56; H, 4.75.
DFT Calculations
DFT calculations were performed by using the Gaussian 09 (D01) and 16 (A01) program packages.43 Optimizations of geometry for the neutral and monocationic complexes were carried out with the (U)B3LYP/LanL2DZ levels of theory with CPCM (CH2Cl2) solvent continuum. Single-point calculations of neutral complexes were performed at the same level of theory. TD-DFT calculations and spin density plots were conducted with UBLYP35/Def2SVP with the CPCM solvent continuum of CH2Cl2 levels of theory.44 We also tried geometry optimization using the UBLYP35/Def2SVP level of theory, but resulted in the charge-localized geometry for the para-isomers, which were not consistent with our experimental results.
Acknowledgments
We sincerely thank Prof. Tadashi Yamaguchi (Waseda Univ.) for his helpful discussion and the generous gift of the VIBEX GL programs. We acknowledge Prof. Masa-aki Haga (Chuo Univ.) for his helpful discussion on the spectroelectrochemical measurements. The computations were performed by using the computer in the Research Center for Computational Science, Okazaki, Japan and by the TSUBAME3.0 supercomputer in the Tokyo Institute of Technology.
Glossary
Abbreviations
- dppe
1,2-bis(diphenylphosphino)ethane
- Cp
cyclopentadinyl
- Cp*
1,2,3,4,5-pentamethylcyclopentadienyl
- W.E
working electrode,
- R.E.
reference electrode,
- C.E.
counter electrode
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsorginorgau.2c00005.
Spectral, electrochemical, and X-ray diffraction data and Cartesian coordinates of optimized geometries (PDF)
Author Contributions
Y.T. and H.T. contributed equally to this work. All the experiments were conducted by H.T. Y.T. contributed to the theoretical study. The manuscript was drafted by Y.T. and edited by M.A. All authors have given approval to the final version of the manuscript.
JSPS KAKENHI grant number 21K05211 ENEOS Tonengeneral Research/Development Encouragement & Scholarship Foundation. The Asahi Glass Foundation. Tokuyama Science Foundation.
The authors declare no competing financial interest.
Supplementary Material
References
- Creutz C.; Taube H. Direct Approach to Measuring the Franck-Condon Barrier to Electron Transfer between Metal Ions. J. Am. Chem. Soc. 1969, 91, 3988–3989. 10.1021/ja01042a072. [DOI] [Google Scholar]
- Robin M. B.; Day P.. Mixed Valence Chemistry-A Survey and Classification. In Advances in Inorganic Chemistry and Radiochemistry; Emeléus H. J., Sharpe A. G., Eds.; Academic Press, 1968; Vol. 10, pp 247–422. [Google Scholar]
- Barrière F.; Camire N.; Geiger W. E.; Mueller-Westerhoff U. T.; Sanders R. Use of Medium Effects to Tune the ΔE1/2 Values of Bimetallic and Oligometallic Compounds. J. Am. Chem. Soc. 2002, 124, 7262. 10.1021/ja020309d. [DOI] [PubMed] [Google Scholar]
- Barrière F.; Geiger W. E. Use of Weakly Coordinating Anions to Develop an Integrated Approach to the Tuning of ΔE1/2 Values by Medium Effects. J. Am. Chem. Soc. 2006, 128, 3980. 10.1021/ja058171x. [DOI] [PubMed] [Google Scholar]
- Hush N. S. Distance Dependence of Electron Transfer Rates. Coord. Chem. Rev. 1985, 64, 135–157. 10.1016/0010-8545(85)80047-3. [DOI] [Google Scholar]
- Demadis K. D.; Hartshorn C. M.; Meyer T. J. The Localized-to-Delocalized Transition in Mixed-Valence Chemistry. Chem. Rev. 2001, 101, 2655–2686. 10.1021/cr990413m. [DOI] [PubMed] [Google Scholar]
- Ito T.; Hamaguchi T.; Nagino H.; Yamaguchi T.; Washington J.; Kubiak C. P. Effects of Rapid Intramolecular Electron Transfer on Vibrational Spectra. Science 1997, 277, 660–663. 10.1126/science.277.5326.660. [DOI] [Google Scholar]
- Mücke P.; Linseis M.; Záliš S.; Winter R. F. Vinyl-Ruthenium Entities as Markers for Intramolecular Electron Transfer Processes. Inorg. Chim. Acta 2011, 374, 36–50. 10.1016/j.ica.2011.03.071. [DOI] [Google Scholar]
- Winter R. F. Half-Wave Potential Splittings ΔE1/2 as a Measure of Electronic Coupling in Mixed-Valent Systems: Triumphs and Defeats. Organometallics 2014, 33, 4517–4536. 10.1021/om500029x. [DOI] [Google Scholar]
- Atwood C. G.; Geiger W. E.; Rheingold A. L. Structural Consequences of Electron-Transfer Reactions. 26. Evidence for Time-Dependent Valence Detrapping in a Mixed-Valent Dimanganese Fulvalenyl Cation. J. Am. Chem. Soc. 1993, 115, 5310–5311. 10.1021/ja00065a059. [DOI] [Google Scholar]
- Stoll M. E.; Lovelace S. R.; Geiger W. E.; Schimanke H.; Hyla-Kryspin I.; Gleiter R. Transannular Effects in Dicobalta–Superphane Complexes on the Mixed-Valence Class II/Class III Interface: Distinguishing between Spin and Charge Delocalization by Electrochemistry, Spectroscopy, and Ab Initio Calculations. J. Am. Chem. Soc. 1999, 121, 9343–9351. 10.1021/ja9914792. [DOI] [Google Scholar]
- Atwood C. G.; Geiger W. E. Investigations of Time-Dependent Delocalization and the Correlation of Carbonyl IR Shifts with ΔE1/2 Values for Hydrocarbon-Linked Class II and Class III Mixed-Valent Complexes. J. Am. Chem. Soc. 2000, 122, 5477–5485. 10.1021/ja994064p. [DOI] [Google Scholar]
- Stoll M. E.; Geiger W. E. In-Situ IR Spectroelectrochemistry Study of One-Electron Oxidations of Bis(Cyclopentadienyl) Molybdenum(II)–Alkyne or −Alkene Compounds. Organometallics 2004, 23, 5818–5823. 10.1021/om040093j. [DOI] [Google Scholar]
- Halet J.-F.; Lapinte C. Charge Delocalization vs Localization in Carbon-Rich Iron Mixed-Valence Complexes: A Subtle Interplay between the Carbon Spacer and the (Dppe)Cp*Fe Organometallic Electrophore. Coord. Chem. Rev. 2013, 257, 1584–1613. 10.1016/j.ccr.2012.09.007. [DOI] [Google Scholar]
- Low P. J. Twists and Turns: Studies of the Complexes and Properties of Bimetallic Complexes Featuring Phenylene Ethynylene and Related Bridging Ligands. Coord. Chem. Rev. 2013, 257, 1507–1532. 10.1016/j.ccr.2012.08.008. [DOI] [Google Scholar]
- Tanaka Y.; Akita M. Organometallic Radicals of Iron and Ruthenium: Similarities and Dissimilarities of Radical Reactivity and Charge Delocalization. Coord. Chem. Rev. 2019, 388, 334–342. 10.1016/j.ccr.2019.02.036. [DOI] [Google Scholar]
- Tanaka Y.; Ozawa T.; Inagaki A.; Akita M. Redox-Active Polyiron Complexes with Tetra(Ethynylphenyl) Ethene and [2,2]Paracyclophane Spacers Containing Ethynylphenyl Units: Extension to Higher Dimensional Molecular Wire. Dalton Trans. 2007, 9, 928–933. 10.1039/b614803d. [DOI] [PubMed] [Google Scholar]
- Mishiba K.; Ono M.; Tanaka Y.; Akita M. A Fully Charge-Delocalized Two-Dimensional Porphyrin System with Two Different Class III States. Chem.—Eur. J. 2017, 23, 2067–2076. 10.1002/chem.201604455. [DOI] [PubMed] [Google Scholar]
- Oyama Y.; Kawano R.; Tanaka Y.; Akita M. Dinuclear Ruthenium Acetylide Complexes with Diethynylated Anthrahydroquinone and Anthraquinone Frameworks: A Multi-Stimuli-Responsive Organometallic Switch. Dalton Trans. 2019, 48, 7432–7441. 10.1039/c9dt01255a. [DOI] [PubMed] [Google Scholar]
- Fox M. A.; Le Guennic B.; Roberts R. L.; Brue D. A.; Yufit D. S.; Howard J. A. K.; Manca G.; Halet J.-F.; Hartl F.; Low P. J. Simultaneous Bridge-Localized and Mixed-Valence Character in Diruthenium Radical Cations Featuring Diethynylaromatic Bridging Ligands. J. Am. Chem. Soc. 2011, 133, 18433–18446. 10.1021/ja207827m. [DOI] [PubMed] [Google Scholar]
- Parthey M.; Gluyas J. B. G.; Fox M. A.; Low P. J.; Kaupp M. Mixed-Valence Ruthenium Complexes Rotating through a Conformational Robin–Day Continuum. Chem.—Eur. J. 2014, 20, 6895–6908. 10.1002/chem.201304947. [DOI] [PubMed] [Google Scholar]
- Fox M. A.; Roberts R. L.; Khairul W. M.; Hartl F.; Low P. J. Spectroscopic Properties and Electronic Structures of 17-Electron Half-Sandwich Ruthenium Acetylide Complexes, [Ru(CCAr)(L2)Cp′]+ (Ar=phenyl, p-Tolyl, 1-Naphthyl, 9-Anthryl; L2=(PPh3)2, Cp′=Cp; L2=dppe; Cp′=Cp*). J. Organomet. Chem. 2007, 692, 3277–3290. 10.1016/j.jorganchem.2007.03.042. [DOI] [Google Scholar]
- Fox M. A.; Farmer J. D.; Roberts R. L.; Humphrey M. G.; Low P. J. Noninnocent Ligand Behavior in Diruthenium Complexes Containing a 1,3-Diethynylbenzene Bridge. Organometallics 2009, 28, 5266–5269. 10.1021/om900200n. [DOI] [Google Scholar]
- Alonso A. G.; Reventós L. B. Cyclopentadienylruthenium Complexes with Sulphur Donor Ligands: II. A Comparative Study of the Reactivity of Ru(η-RC5H4)Cl(L)2 (R = H, CH3, CH3CO; L = CO, Ph2PCH2CH2PPh2/2, P(CH2CH2CN)3) towards Anionic (S-S) Donor Ligands. J. Organomet. Chem. 1988, 338, 249–254. 10.1016/0022-328x(88)80509-6. [DOI] [Google Scholar]
- Bruce M.; Wallis R. Cyclopentadienyl-Ruthenium and-Osmium Chemistry. IX. Some Substituted Η1-Vinylidene and Η1-Acetylide Complexes. Aust. J. Chem. 1979, 32, 1471–1485. 10.1071/ch9791471. [DOI] [Google Scholar]
- The DFT calculation of diethynylbenzene shows almost forbidden symmetric ν(C≡C) vibration (f = 0.0003). The DFT calculation was performed at the B3LYP/6-31G(d) level.
- Bruce M. I.; Humphrey M. G.; Snow M. R.; Tiekink E. R. T. Cyclopentadienyl-Ruthenium and -Osmium Chemistry: XXVII. X-Ray Structures of Ru(C≡CPh)(Dppe)(η-C5H5) and of [Ru(L)(PPh3)2(η-C5H5)]X (L = C(OMe)Et, X = PF6; L = C=CMePh, X = I). J. Organomet. Chem. 1990, 397, 187. 10.1016/0022-328x(86)80364-3. [DOI] [Google Scholar]
- Armitt D. J.; Gaudio M.; Skelton B. W.; White A. H.; Halet J.-F.; Roberts R. L.; Low P. J.; Halet J.-F.; Fox A.; Roberts L.; Hartl F.; Low J. Some Transition Metal Complexes Derived from Mono- and Di-Ethynyl Perfluorobenzenes. Dalton Trans. 2008, 6763–6775. 10.1039/b808798a. [DOI] [PubMed] [Google Scholar]
- Gauthier N.; Tchouar N.; Justaud F.; Argouarch G.; Cifuentes M. P.; Toupet L.; Touchard D.; Halet J.-F.; Rigaut S.; Humphrey M. G.; Costuas K.; Paul F. Bonding and Electron Delocalization in Ruthenium(III) σ-Arylacetylide Radicals [Trans-Cl(Η2-Dppe)2RuC≡C(4-C6H4X)]+ (X = NO2, C(O)H, C(O)Me, F, H, OMe, NMe2): Misleading Aspects of the ESR Anisotropy. Organometallics 2009, 28, 2253–2266. 10.1021/om801138q. [DOI] [Google Scholar]
- Parthey M.; Kaupp M. Quantum-Chemical Insights into Mixed-Valence Systems: Within and beyond the Robin–Day Scheme. Chem. Soc. Rev. 2014, 43, 5067–5088. 10.1039/c3cs60481k. [DOI] [PubMed] [Google Scholar]
- Grevels F.-W.; Kerpen K.; Klotzbücher W. E.; McClung R. E. D.; Russell G.; Viotte M.; Schaffner K. The Very Low Barrier of CO Site Exchange in Tricarbonyl(Η4-1,5-Cyclooctadiene)Iron: Picosecond Kinetics in Solution Investigated by Line Shape Simulation of the ν(CO) IR Bands and Complementary Evidence from the Course of 13CO Incorporation in a Low-Temperature Matrix. J. Am. Chem. Soc. 1998, 120, 10423–10433. 10.1021/ja973897h. [DOI] [Google Scholar]
- We made attempts to fit the spectrum with the assumption that the peak consists of two different peaks due to rotamers, and, however, the attempts were failed.
- The spectral changes upon the oxidation of m-1b are similar to those with m-1a.
- While a couple of rotamers afford the multiple NIR bands, those giving the major NIR bands can be classified into charge-delocalized species.
- For the estimation of Vab of class II compounds, the donor-acceptor distance should be taken into account. Because, however, the charge is delocalized, the exact distance is hardly determined. For details see refs (36) and (37).
- Silverman L. N.; Kanchanawong P.; Treynor T. P.; Boxer S. G. Stark Spectroscopy of Mixed-Valence Systems. Proc. R. Soc. London, Ser. A 2008, 366, 33–45. 10.1098/rsta.2007.2137. [DOI] [PubMed] [Google Scholar]
- Heckmann A.; Lambert C.; Goebel M.; Wortmann R. d. Synthesis and Photophysics of a Neutral Organic Mixed-Valence Compound. Angew. Chem., Int. Ed. 2004, 43, 5851–5856. 10.1002/anie.200460495. [DOI] [PubMed] [Google Scholar]
- ESR spectra of p-1a,b+ observed as CH2Cl2 solutions at room temperature contained the broadened isotropic signals, which suggest the significant contribution of the organic ligand-based radical.
- Paul F.; Ellis B. G.; Bruce M. I.; Toupet L.; Roisnel T.; Costuas K.; Halet J.-F.; Lapinte C. Bonding and Substituent Effects in Electron-Rich Mononuclear Ruthenium σ-Arylacetylides of the Formula [(Η2-Dppe)(Η5-C5Me5)Ru(C:C)-1,4-(C6H4)X][PF6]n (n = 0, 1; X = NO2, CN, F, H, OMe, NH2). Organometallics 2006, 25, 649–665. 10.1021/om050799t. [DOI] [Google Scholar]
- Brunschwig B. S.; Creutz C.; Sutin N. Optical Transitions of Symmetrical Mixed-Valence Systems in the Class II–III Transition Regime. Chem. Soc. Rev. 2002, 31, 168–184. 10.1039/b008034i. [DOI] [PubMed] [Google Scholar]
- Pangborn A. B.; Giardello M. A.; Grubbs R. H.; Rosen R. K.; Timmers F. J. Safe and Convenient Procedure for Solvent Purification. Organometallics 1996, 15, 1518–1520. 10.1021/om9503712. [DOI] [Google Scholar]
- Hallman P. S.; Stephenson T. A.; Wilkinson G.. Tetrakis(Triphenylphosphine)Dichlororuthenium(II) and Tris(Triphenylphosphine)Dichlororuthenium(II). In Inorganic Syntheses; John Wiley & Sons, Ltd, 1970; pp 237–240. [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; 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 J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene C.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford CT, 2016.
- Renz M.; Theilacker K.; Lambert C.; Kaupp M. A Reliable Quantum-Chemical Protocol for the Characterization of Organic Mixed-Valence Compounds. J. Am. Chem. Soc. 2009, 131, 16292–16302. 10.1021/ja9070859. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.











