Abstract

The reactivity of the bis(pentalene)dititanium double-sandwich compound Ti2Pn†2 (1) (Pn† = 1,4-{SiiPr3}2C8H4) with CO2 is investigated in detail using spectroscopic, X-ray crystallographic, and computational studies. When the CO2 reaction is performed at −78 °C, the 1:1 adduct 4 is formed, and low-temperature spectroscopic measurements are consistent with a CO2 molecule bound symmetrically to the two Ti centers in a μ:η2,η2 binding mode, a structure also indicated by theory. Upon warming to room temperature the coordinated CO2 is quantitatively reduced over a period of minutes to give the bis(oxo)-bridged dimer 2 and the dicarbonyl complex 3. In situ NMR studies indicated that this decomposition proceeds in a stepwise process via monooxo (5) and monocarbonyl (7) double-sandwich complexes, which have been independently synthesized and structurally characterized. 5 is thermally unstable with respect to a μ-O dimer in which the Ti–Ti bond has been cleaved and one pentalene ligand binds in an η8 fashion to each of the formally TiIII centers. The molecular structure of 7 shows a “side-on” bound carbonyl ligand. Bonding of the double-sandwich species Ti2Pn2 (Pn = C8H6) to other fragments has been investigated by density functional theory calculations and fragment analysis, providing insight into the CO2 reaction pathway consistent with the experimentally observed intermediates. A key step in the proposed mechanism is disproportionation of a mono(oxo) di-TiIII species to yield di-TiII and di-TiIV products. 1 forms a structurally characterized, thermally stable CS2 adduct 8 that shows symmetrical binding to the Ti2 unit and supports the formulation of 4. The reaction of 1 with COS forms a thermally unstable complex 9 that undergoes scission to give mono(μ-S) mono(CO) species 10. Ph3PS is an effective sulfur transfer agent for 1, enabling the synthesis of mono(μ-S) complex 11 with a double-sandwich structure and bis(μ-S) dimer 12 in which the Ti–Ti bond has been cleaved.
Introduction
In light of increasing global energy demands and environmental concerns over anthropogenic climate change,1 there is considerable interest in the utilization of CO2 as a renewable C1 feedstock for industrially important chemicals.2−4 One of the most promising methods is the use of transition-metal catalysts to bring about selective and energy-efficient CO2 incorporation into organic compounds or reduction to CO, which can then be converted into liquid fuels via Fischer–Tropsch chemistry.5
The difficulty in achieving CO2 reduction lies in the thermodynamic stability of the CO2 molecule and the kinetic barrier to its activation. As a result, highly reducing oxophilic metal complexes are typically employed. Stoichiometric CO2 activation has been achieved using well-defined complexes of the main-group and transition metals6−10 as well as metal-free systems based on frustrated Lewis pairs (FLPs).11−13 Lower-oxidation-state f-element complexes are also being increasingly explored in this context: for example, Evans and co-workers have shown that samarium(II) metallocenes facilitate the reductive coupling of CO2 to form oxalate [O2CCO2]2– 14,15 and the reductive disproportionation of COS to [S2CO]2– and CO.15 More recently, organouranium(III) complexes have attracted interest because of their significant capacity for CO2 activation.16−18 For example, trivalent uranium metallocenes U(CpR)3 reductively activate CO2 (for CpR = η5-C5H4{SiMe3}) and COS (for CpR = η5-C5H4Me) to generate μ-oxo/μ-sulfido products with concomitant release of CO.19,20 Density functional theory (DFT) studies by Maron and co-workers have provided insight into the mechanism of reductive CO2 and COS activation at Sm(II)21 and U(III)22,23 centers, implicating in each case a bimetallic μ-CO2 complex in which carbon dioxide was doubly reduced. The possibility of a concerted pathway involving a monometallic CO2 intermediate with a [CO2]•– radical anion ligand was ruled out on the basis of the observed products. Meyer and co-workers reported the first example of a linearly coordinated [CO2]•– moiety in a U(III) complex supported by a stabilizing macrocyclic ligand, thus illustrating the importance of ligand architecture in reactivity.24 Recent studies by Cloke and co-workers with the mixed-sandwich U(III) complexes (COT′)UCpMe4R{THF}x (COT′ = η8-1,4-{SiMe3}2C8H6; R = Me, Et, iPr, tBu) have shown that the steric environment around the metal center plays a key role in guiding the possible reductive transformation pathways of CO2, i.e., reductive coupling, disproportionation, or deoxygenation.25 Nonetheless, better understanding of the mechanisms of these reductive transformations is required, with a view to tuning the structure of the active complex or the reaction conditions to give more useful product outcomes and assessing the potential for catalytic turnover.26
Divalent titanium sandwich complexes are an underdeveloped series of compounds in this regard, despite the rich and varied reaction chemistry that “titanocenes” have shown with a variety of small molecules.27,28 Alt et al.29 reported the synthesis of the titanium CO2 adduct Cp2Ti(CO2)(PMe3) via simple ligand substitution of Cp2Ti(PMe3)2, and a number of acetylene complexes of Cp2Ti have shown CO2 insertion into the Ti–C bond of the titanacyclopropene ring.30−33 In contrast, reaction of the permethyltitanocene complex Cp*2Ti(η2-Me3SiC2SiMe3) with CO2 yields, after elimination of the silylalkyne, a mixture of the dimeric carbonate (Cp*2Ti)2(CO3) and the carbonyl complex Cp*2Ti(CO)2.34,35 Reductive disproportionation of CO2 has been achieved by Cp2Ti(CO)2 to generate CO and the carbonate-bridged tetranuclear complex [(Cp2Ti)2(CO3)]2.36 Titanium(III) complexes have also shown efficacy in reductive transformations of CO2 to yield oxo,36 carbonate,36−39 and oxalate40 products, and the relative cheapness and low toxicity of this metal makes it attractive for fundamental reactivity studies.
Multiple bonds between metal atoms have fascinated chemists for over 50 years,41 and M–M bonded complexes have also shown ability to facilitate the reductive activation of CO2. Recent examples include systems featuring reduced Mg–Mg,42 Cr–Cr,43 and Pd–Pd cores44 and a highly polar heterobimetallic Zr–Co complex recently reported by Thomas and co-workers.45
Our interest in M–M interactions is focused on the aromatic ligand pentalene (Pn, C8H6), which is related to Cp by the edge-sharing ring fusion of two cyclopentadienyl fragments and shows a unique variety of coordination modes in its organometallic complexes.46 In particular, the ability of Pn to bind two metal centers with η5 hapticity has proven ideal for the synthesis of bimetallic analogues of the ubiquitious bis(Cp)metallocenes, so-called “double-sandwich” complexes.47−53 We have recently reported the synthesis and characterization of the titanium double-sandwich complex Ti2(μ:η5,η5-Pn†)2 (1) using the silylated pentalene ligand 1,4-{SiiPr3}2C8H4 (Pn†). In addition to providing a rare example of a Ti–Ti multiple bond, 1 exhibits a solid-state structure unique among the M2Pn2 complexes in which the pentalene ligands are nonparallel and tilted around the Ti2 core.54 DFT studies on the model system Ti2Pn2 calculated an M–M bond order of 2 and an optimized geometry with C2v molecular symmetry, providing a relatively open structure in which the frontier orbitals are sterically accessible.
The combination of a low-valent group IV metal and a metal–metal double-bonded system led us to explore the reactivity of 1 with small molecules. We recently communicated the room-temperature reaction of 1 with 1 equiv of CO2 to give the bis(oxo)-bridged dimer [(η8-Pn†)Ti(μ-O)]2 (2), in which the Ti–Ti bond has been cleaved, and the dicarbonyl complex Ti2(μ:η5,η5-Pn†)2(CO)2 (3), which was independently synthesized from 1 with 2 equiv of CO.55 Herein we explore the scope of this reductive transformation of CO2, investigating its mechanistic pathway by the identification of key intermediates using spectroscopic and computational techniques. Furthermore, by the use of isoelectronic CS2 and unsymmetrical COS as model molecules, we report the isolation and structural characterization of stable analogues for the proposed reaction intermediates, supporting the theoretical predictions. In the interests of brevity, an in-depth description of the bonding in these molecules and other related compounds is provided in the companion paper.56
Results
Reaction of CO2 with Ti2Pn†2
The reaction of 1 with 1 equiv of either 12CO2 or 13CO2 was carried out in methylcyclohexane-d14 (MeCy-d14) at −78 °C (Scheme 1). This resulted in a color change from deep red to dark green, which accompanied quantitative conversion of 1 to an intermediate species, 4, that was observed by low-temperature spectroscopic measurements but had not been previously detected in a room-temperature reaction. The 1H NMR spectrum of 4 at −30 °C showed four sharp doublets in the aromatic region, consistent with a diamagnetic complex that exhibits C2 molecular symmetry on the NMR time scale, while the 13C{1H} spectrum showed a singlet at 219 ppm with no further labeled 13C signals. In situ IR studies of 4 at −65 °C showed distinct bands at 1678 and 1236 cm–1, which shifted to 1637 and 1217 cm–1, respectively, in the isotopically labeled [13C]4. These IR bands are assigned to asymmetric and symmetric ν(OCO) stretches, respectively, of the CO2 ligand by analogy with the monomeric complex Cp2Ti(CO2)(PMe3),29 which was studied in-depth by Mascetti et al.57 and shows a C-coordinated bonding mode. Compound 4 could not be isolated because of its susceptibility to further reaction.
Scheme 1. Reactivity of 1 with CO2 (R = SiiPr3).
When 4 was allowed to warm from −78 °C to ambient temperature under dynamic vacuum, the red mono(oxo) complex 5 was isolated (Scheme 1). 5 could also be obtained by the slow addition of 1 equiv of N2O to 1 via Toepler line. The 1H NMR spectrum of 5 is again consistent with a diamagnetic complex with C2 symmetry, and elemental analysis and mass spectrometry measurements support the proposed formulation. The molecular structure determined by single-crystal X-ray diffraction (XRD) reveals a “dimetallaepoxide” structure with an oxo ligand symmetrically bridging two formally Ti(III) centers (Figure 1). The Pn† ligands are less tightly bound to the Ti2 core of 5 relative to 1 and show the largest average Ti–C and Ti–centroid distances among all of the double-sandwich derivatives structurally characterized herein (Table 1). Interestingly, the ring slippage parameter (Δ = 0.15) and the twist angle between the Pn† ligands and the dimetal unit (26.6(5)°) also have the largest values among all of these structures (see Figure 2 for definitions of these parameters).
Figure 1.

Displacement ellipsoid plot (30% probability) of 5 with hydrogen atoms and iPr groups omitted for clarity. Primed atoms are generated by symmetry.
Table 1. Selected Distances (Å), Angles (deg), and Values of the Parameters Defined in Figure 2 for 5, 7, 8, 11, and 12 (Ct Denotes the η5 Centroid of a Pn Ring).
| parameter | 5 | 7 | 8 | 11 | 12 |
|---|---|---|---|---|---|
| Ti–Ti | 2.3991(7) | 2.4047(5) | 2.4432(10) | 2.4880(8) | 3.1402(6) |
| Ti–Cta | 2.133(10) | 2.058(9) | 2.104(13) | 2.1088(13) | 1.9470(18) |
| Ct–Ti–Cta | 139.722(12) | 146.41(5) | 138.02(9) | 137.11(5) | 57.09(9) |
| C53–Tiproximala | – | 1.975(9) | 2.238(4) | – | – |
| Ti–O/Sa | 1.8607(15) | 2.352(6) | 2.497(16) | 2.3728(8) | 2.323(2) |
| C53–O/Sa | – | 1.211(11) | 1.657(4) | – | – |
| Δa | 0.150(19) | 0.005(7) | 0.114(4) | 0.135(2) | – |
| twist angle | 26.6(5) | 20.1(8) | 20.8(2) | 20.3(2) | 21.1(5) |
| fold anglea | 0.15(13) | 8.7(4) | 4.6(2) | 3.86(8) | 17.30(12) |
Average values.
Figure 2.

Definition of the ring slippage, twist angle, and fold angle parameters for pentalene structures.
Complex 5 is thermally sensitive, and allowing a C6D6 solution to stand for 12 h at room temperature led to the precipitation of green crystals of 6. X-ray crystallography established the connectivity in 6 as a μ-O dimer in which the Ti–Ti bond has been cleaved and one pentalene ligand binds in an η8 fashion to each of the formally Ti(III) centers (see Scheme 1 and Figure S1 in the Supporting Information). However, the low quality of the X-ray data precludes meaningful discussion of the metric parameters. In the 1H NMR spectrum of 6 in THF-d8 solution, only broad signals were observed, and the effective magnetic moment determined by the Evans method58,59 was 1.73μB per Ti, a value consistent with one unpaired electron per metal center. Elemental analysis and mass spectrometry measurements were consistent with the proposed formulation for 6.
A monocarbonyl complex, 7, was prepared by the addition of 1 equiv of CO to a solution of 1 at −78 °C via Toepler line (Scheme 2). 1H NMR spectroscopy showed that the reaction is quantitative, and red crystals of 7 were isolated from pentane at −35 °C in 62% yield. The solid-state structure of 7 (Figure 3) shows the CO ligand bound “side-on”, asymmetrically bridging the two Ti centers. The proximal Ti atom lies closer to the CO ligand in 7 than in the dicarbonyl complex 3, and an additional interaction between the CO and the distal Ti atom is clearly in effect in the former complex (Ti1–C53a = 2.245(10) Å, Ti1–O1a = 2.294(5) Å), causing the O–C53–Tiproximal angle to deviate considerably from linearity (Table 1).
Scheme 2. Reactivity of 1 with CO (R = SiiPr3).

Figure 3.

Displacement ellipsoid plot (30% probability) of 7 with hydrogen atoms and iPr groups omitted for clarity.
The 1H NMR spectrum of 7 features eight doublet resonances in the aromatic region (see Figure S3 in the Supporting Information), consistent with an unsymmetrical structure. The 1H NMR spectrum of the isotopically labeled product [13C]7 differs from that of [12C]7 in that the Pn† ring signal at 7.69 ppm appears as a triplet (J = 3.2 Hz) with an 1H–13C (HMBC) correlation to the isotopically enriched μ-13CO. Inspection of the 29Si{1H} NMR spectrum, which consists of four singlets for [12C]7, reveals splitting of one signal into a doublet (J = 1.5 Hz) for [13C]7 (see Figure S4 in the Supporting Information). These features are attributed to through-space scalar couplings between the carbonyl carbon and the hydrogen and silicon in the proximal “wing-side” position of the Pn† ligands above and below the Ti2CO plane (Figure 4). The average distance between the CCO and these H and Si atoms in the solid-state structure are 2.807(8) and 3.846(11) Å respectively. The solid-state IR spectrum of 7 exhibits a low ν(CO) stretching frequency of 1655 cm–1, which shifts to 1616 cm–1 upon 13C isotopic substitution, consistent with a side-on-bound carbonyl ligand.
Figure 4.

Through-space coupling between CO and Pn† wing-side substituents.
Kinetic Studies
Experimentally, the reaction of 1 with CO2 at low temperatures (between −90 and −30 °C) led to the quantitative formation of CO2 adduct 4 (vide supra). When solutions of 4 were allowed to reach room temperature (in the absence of a dynamic vacuum) a color change from green to brown was observed, and the coordinated CO2 molecule was reduced quantitatively over a period of minutes to give bis(oxo)-bridged dimer 2 and the dicarbonyl complex 3. 1H NMR spectra recorded during the course of this reaction showed additional transient species, including (as the major components) the mono(oxo) compound 5 and monocarbonyl complex 7, identified by comparison with the spectra of independently synthesized samples.
The rate of decomposition of 4 was such that it could be conveniently followed by 1H NMR spectroscopy. In a typical experiment, a 0.0295 M solution of 1 in MeCy-d14 at −78 °C (in the presence of ferrocene, with which 1 does not react, as an internal standard) was treated with 1 equiv of 13CO2, and the resulting mixture was shaken over 30 min until complete conversion to dark-green 4 was observed. The reaction mixture was then placed in an NMR probe held at 20 °C and allowed to equilibrate thermally. The conversion of reactants to products was measured over time and showed first-order exponential decay of 4 (Figure 5). The rate constant for this process was therefore determined at seven different temperatures in the range 9–40 °C (see Table S3 in the Supporting Information), and the relevant Eyring plot is shown in Figure 6. The derived activation parameters are ΔH⧧ = 19.6 ± 1.0 kcal mol–1, ΔS⧧ = −9 ± 3 e.u., and ΔG298 K⧧ = 22.4 ± 1.5 kcal mol–1.
Figure 5.

Concentration vs time curves for the decomposition of 4 to give oxo products 5 and 2 at 20 °C, as monitored by 1H NMR spectroscopy.
Figure 6.

Eyring plot for the decomposition of 4 in MeCy-d14 (R2 = 0.990).
Reaction of CS2 and COS with Ti2Pn†2
Heteroallenes CS2 and COS are commonly used to model the reactivity of CO2.60,61 Treatment of a pentane solution of 1 with CS2 at −35 °C gave a dark-green solution upon warming to room temperature. Subsequent workup and cooling to −35 °C produced a red microcrystalline solid in 62% yield, which was identified by spectroscopic and analytical methods as 8 (Scheme 3). Crystals suitable for X-ray studies were grown from an Et2O solution, and the molecular structure (Figure 7) reveals a bent CS2 molecule in a μ:η2,η2 binding mode between the two Ti centers.
Scheme 3. Reaction of 1 with CS2 (R = SiiPr3).

Figure 7.

Displacement ellipsoid plots (30% probability) of 8, 11, and 12 with hydrogen atoms and iPr groups omitted for clarity. Primed atoms are generated by symmetry.
The CS2 ligand in 8 exhibits a S–C–S angle of 137.4(3)°, while the C–S bond lengths (av 1.657(4) Å) are longer than the C–S bond length in free CS2 (1.554 Å).62 The double-sandwich structure remains intact, with an elongated Ti–Ti distance (2.4432(10) Å) relative to 1 (2.399(2) Å)54 and a smaller angle between the pentalene rings and the Ti2 unit (av Ct–Ti–Ct = 138.02(9)°), giving a more open sandwich structure (Table 1). The ring slippage parameters (Figure 2) for each metal (ΔTi1 = 0.145, 0.092; ΔTi2 = 0.138, 0.082) indicate differential bonding to each side of the pentalene ligand.
The solid-state structure of 8 provides supporting evidence for the doubly bridging CO2 ligand in 4. However, 8 is stable at room temperature and does not decompose even upon heating to 100 °C, in stark contrast to the CO2 adduct 4, which decomposes above −30 °C to give the μ-oxo complex with concomitant liberation of carbon monoxide. This may be explained by the relative stability of CO compared with CS (ΔfH298 K° = −26.4 and 66.1 kcal mol–1, respectively)63,64 and the driving force for Ti–O bond formation (ΔbondH298 K = 160.7 ± 2.2 kcal mol–1) relative to Ti–S (ΔbondH298 K° = 99.9 ± 0.7 kcal mol–1).65 The notable difference in the reactivity of CS2 versus CO2 with 1 prompted a study with COS.
The reaction of 1 with 1 equiv of COS at −78 °C in MeCy-d14 produced a metastable adduct, 9, that shows an 1H NMR spectrum consistent with the loss of C2 molecular symmetry, as opposed to those of the C2-symmetric species 4 and 8. The 13C{1H} NMR spectrum of 9 shows a singlet at 282 ppm, a value intermediate between those of the CO2 and CS2 analogues (219 and 356 ppm respectively). In situ IR spectroscopy at −65 °C showed a band at 1498 cm–1 which is tentatively assigned to the ν(CO) stretch of coordinated COS. However, the COS adduct 9 proved to be more thermally unstable than the CO2 adduct 4, and the coordinated molecule underwent scission to mono(μ-S) mono(CO) species 10 in 41% conversion as determined by 1H NMR spectroscopy at −40 °C. Intermediate 10 was independently prepared by the reaction of mono(μ-S) complex 11 with 1 equiv of CO; 11 itself was readily synthesized by the reaction of 1 with Ph3PS (Scheme 4) and fully characterized by spectroscopic and analytical methods (vide infra).
Scheme 4. Reactivity of 1 with COS and Ph3PS (R = SiiPr3).
The synthesis of compound 11 was achieved by the reaction of 1 with 1 equiv of Ph3PS in toluene (Scheme 4), and following workup 11 was isolated as red crystals in 52% yield. The proposed formulation of 11 is consistent with 1H, 13C, and 29Si NMR and elemental analysis data and was ultimately confirmed by a single-crystal XRD study. The molecular structure of 11 (Figure 7) shows the sulfide ligand bridging the two Ti centers in a three-membered ring, akin to mono(μ-oxo) complex 5. The Ti–Ti bond is considerably longer in 11 than in 5 (2.4880(8) and 2.3991(7) Å, respectively), and the larger bridging chalcogen atom is bonded at a greater distance from the Ti2 core (av Ti–S = 2.3728(8); cf. av Ti–O = 1.8607(15) Å) with a more acute Ti–E–Ti angle (62.68(4)° and 80.28(8)°, respectively) (see Table 1). However, the smaller values of the Ti–Ct distance, ring slippage (Δ), and twist angle (Figure 2) for 11 reflect the fact that the Pn† ligands are more tightly bound compared with 5, and this may explain the relative stability of this mono(μ-S) complex.
This triangular M–S–M structural motif is common among binuclear complexes with thiolato bridges but is unique in titanocene monosulfide chemistry. For example, [Cp*2Ti]2(μ-S) is a green paramagnetic complex with a Ti···Ti separation of 4.7069(7) Å and a Ti–S–Ti angle of 174.37(4)°.66 The lack of a bonding (or an antiferromagentic coupling) interaction between the two d1 centers in [Cp*2Ti]2(μ-S) gives rise to an electronic triplet state. In contrast, both 11 and 10 are diamagnetic.
The bis(μ-sulfide) complex 12 was independently synthesized by the reaction of 1 with 2 equiv of Ph3PS in toluene (Scheme 4) and isolated in 80% yield. NMR spectroscopy, mass spectrometry, and elemental analysis confirmed the identity and purity of the product. Single crystals suitable for XRD analysis were grown from a saturated Et2O solution at −35 °C, and its molecular structure is depicted in Figure 7, with important structural parameters listed in Table 1. Complex 12 is highly symmetrical; the solid-state structure shows a Ti2S2 unit that makes a regular ring with a perfectly planar diamond core (Ti1–S1–Ti1′–S1′ torsion angle = 0°). However, there is significant asymmetry in the Ti–S bond lengths (2.2238(16) and 2.4227(16) Å). The Ti–S–Ti angle is relatively acute (84.92(6)°), yet the Ti···Ti distance is considerably longer than in the bis(μ-oxo) congener (2.7376(6) Å).55 The relative orientation of the η8-pentalene ligands (as defined by the angles between the two bridgehead C–C vectors) is 21.1(5)°.
Calculated Structures
Theoretical calculations were carried out on model systems with Pn (C8H6) as the ligand, mostly using two different functional/basis set combinations as embedded in two codes (ADF, BP/TZP; Gaussian, B3LYP/SDD). Selected calculated structural parameters are listed in Table 2 (in this article, experimentally observed/isolated compounds are labeled with Arabic numerals and their computational equivalents with the corresponding Roman numerals). In addition to the characterized compounds reported herein, structures were optimized for a possible intermediate and transition state in the transformation of IV.
Table 2. Selected Calculated Structural Parameters (Å, deg) for Optimized Structuresa.
| structure | Ti–Ti | Ti–Ct | Ct–Ti–Ct | Ti–C | Ti–O/S | C–O/S | Ti–O/S–Ti |
|---|---|---|---|---|---|---|---|
| Ti2Pn2 (I)54 | 2.37, 2.31 | 2.00, 2.03 | 153, 158 | – | – | – | |
| PnTiO2TiPn (II) | 2.74, 2.75 | 2.00, 2.02 | 56, 56 | – | 1.87, 1.85 | – | 95, 96 |
| Ti2Pn2(CO)2 (III) | 2.42, 2.42 | 2.05, 2.05 | 144, 144 | 2.08, 2.08 | 1.17, 1.17 | – | |
| 142, 144 | 1.17, 1.18 | ||||||
| Ti2Pn2CO2 (IV) | 2.41, 2.40 | 2.07, 2.10 | 141, 141 | 2.18, 2.14 | 2.27, 2.25 | 1.26, 1.29 | – |
| Ti2Pn2O (V) | 2.38, 2.36 | 2.13, 2.14 | 139, 140 | – | 1.87, 1.85 | – | 79, 79 |
| PnTiOTiPn (S = 1) (VIt) | 3.40, 3.69 | 1.96, 1.99 | 57, 57 | – | 1.86, 1.85 | – | 133, 180 |
| PnTiOTiPn (S = 0) (VIs) | 2.88, 2.80 | 1.96, 1.99 | 57, 57 | – | 1.85, 1.83 | – | 103, 100 |
| Ti2Pn2CO (VII) | 2.38, 2.36 | 2.06, 2.08 | 143, 143 | 2.04, 2.02 | 2.35, 2.26 | 1.21, 1.25 | – |
| Ti2Pn2CS2 (VIII) | 2.43, 2.41 | 2.10, 2.11 | 138, 138 | 2.27. 2.24 | 2.54, 2.58 | 1.67, 1.72 | – |
| Ti2Pn2COS (IX) | 2.41, 2.40 | 2.09, 2.11 (S) | 138, 140 (S) | 2.25, 2.20 (S) | 2.63, 2.69 (S) | 1.68, 1.73 (S) | – |
| 2.08, 2.09 (O) | 140, 141 (O) | 2.19, 2.17 (O) | 2.19, 2.15 (O) | 1.26, 1.29 (O) | |||
| Ti2Pn2S(CO) (X) | 2.48, 2.47 | 2.12, 2.13 | 135, 136 | 2.08, 2.07 | 2.30, 2.31 | 1.16, 1.18 | 61, 61 |
| 2.11, 2.13 | 136, 136 | 2.54, 2.57 | |||||
| Ti2Pn2S (XI) | 2.44, 2.42 | 2.11, 2.12 | 137, 140 | – | 2.37, 2.39 | – | 62, 61 |
| PnTiS2TiPn (XII) | 3.13, 3.19 | 1.98, 1.99 | 57, 57 | – | 2.33, 2.33 | – | 84, 86 |
| Ti2Pn2O(CO) (Int1) | 2.46, 2.43 | 2.18, 2.17 | 135, 137 | 2.08, 2.07 | 1.76, 1.74 | 1.16, 1.17 | – |
| 2.11, 2.13 | 138, 138 | 2.10, 2.07 | |||||
| Ti2Pn2OCO (TS1) | 2.51, 2.50 | 2.11, 2.13 | 136, 136 | 2.16, 2.14 | 1.96, 1.93 | 1.36, 1.40 | – |
| 2.09, 2.10 | 143, 142 | 2.75, 2.63 | 3.16, 3.09 | 1.23, 1.25 |
ADF (BP/TZP) values are given in normal text, and Gaussian (B3LYP/SDD) values are given in italics.
Discussion
Structure of Ti2Pn†2CO2
Geometry optimization of the model system Ti2Pn2CO2 (replacement of the silyl groups by H enhances the molecular symmetry) gave structure IV with C2v symmetry and showing a bent CO2 group bound symmetrically to the two Ti atoms (Figure 8). The calculated asymmetric and symmetric stretching modes for the CO2 group are 1669 and 1214 cm–1, respectively, which are in good agreement with the experimental results. Thus, we conclude that the high-symmetry NMR spectrum is a result of a symmetrical structure rather than a fast exchange process.
Figure 8.

Optimized geometries of IV and VIII.
Carbon dioxide has been observed to bind to dinuclear metal sites with three different bonding modes (Figure 9).10,67−69 The proposed μ:η2,η2- coordination geometry has only recently been structurally authenticated by X-ray crystallography in an N-heterocyclic carbene nickel(0) dimer, [(IPr)Ni]2(μ-CO)(μ:η2,η2-CO2) (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene).70
Figure 9.

Crystallographically characterized bonding modes of CO2 in bimetallic complexes.
Support for the symmetrical binding of CO2 is provided by the authenticated structure of the sulfur analogue. Geometry optimization of Ti2Pn2CS2 also resulted in a structure VIII of C2v symmetry (Figure 8), in good agreement with the experimental stucture of 8. The solution IR spectrum of 8 shows a band at 1104 cm–1 which, compared with the asymmetric stretch of free CS2 (1535 cm–1),71 is consistent with a sulfur-bound CS2 moiety. This is in fair agreement with the stretching vibration for the model system VIII (1079 cm–1) calculated by DFT methods.56 Precedent for this type of CS2 coordination involving both C=S π bonds is limited to a single example of a dinuclear Cu(I) complex recently reported by Limberg and co-workers.72 However, in this complex the mean plane of the CS2 atoms is twisted with respect to the M–C–M mean plane by 49.1°, whereas in 8 these planes are nearly parallel (3.1°).
Structure of Ti2Pn†2CO
The side-on binding of CO in 7 is sufficiently unusual to deserve comment. Compound 7 is, to the best of our knowledge, the first example of a “semibridging” carbonyl ligand between two titanium centers, and heterometallic examples involving titanium carbonyls are very rare. A weak semibridging carbonyl–gold interaction was inferred in the anionic gold–titanium complex [K(15-crown-5)2][Ti(CO)6(AuPEt3)] on the basis of X-ray structural parameters and a difference between the solid-state and solution IR ν(CO) stretching frequencies.73 Semibridging carbonyls have also been observed in the aforementioned Ti/M heterobimetallics Cp2Ti(μ-PEt3)(μ-η1:η2-OC)M(CO)Cp (M = Mo, W),74 Cp2Ti(μ-C{R}CH2)(μ-η1:η2-OC)W(CO)Cp,75,76 and Cp2Ti(μ-CR)(μ-η1:η2-OC)W(CO)Cp (R = p-Tol),77 which show substantial reduction in the C–O bond order compared with typical terminal carbonyl moieties, as evidenced by the C–O bond distance and ν(CO).
The geometric function Ω = exp[d(Mdistal–C)/d(Mdistal–O)] measures the extent of the interactions of the distal metal atom with the C and O ends of the CO ligand.78 This function can be used to differentiate between C- and O-bonded metal carbonyls in “end-on” (Σ-CO) and “side-on” (Π-CO) disposition. The average value for 7 (Ω = 2.69) lies comfortably in the range for a Π-CO (typically 2.2 < Ω < 3.3), and comparable values have been found for Ti/M heterobimetallics with semibridging CO ligands: Cp2Ti(μ-C{p-Tol}CH2)(μ-η1:η2-OC)W(CO)Cp (Ω = 2.65)75,76 and Cp2Ti(μ-C{p-Tol})(μ-η1:η2-OC)W(CO)Cp (Ω = 2.63).77 The C–O bond distance in 7 (av 1.211(11) Å) is lengthened with respect to that in free CO (1.128 Å)79 to an extent that is comparable to the Σ-CO complexes Cp2Ti(THF)(μ-OC)Mo(CO)2Cp (1.201(8) Å)80 and Cp*2Ti(Me)(μ-OC)Mo(CO)2Cp (1.212(5) Å),81 in which the titanocene moiety behaves as a Lewis acid, coordinating to an oxygen atom of the Lewis basic [CpMo(CO)3]− fragment.
Geometry optimization of Ti2Pn2CO resulted in a structure with Cs symmetry (VII) in which the carbonyl ligand is bound side-on to the Ti dimer (Figure 10). A double bond is retained between the two Ti atoms in VII,56 and the calculated Ti–Ti distance (2.38, 2.36 Å) is in reasonable agreement with the X-ray data (2.4047(5) Å). The wavenumber for the CO stretch in VII was calculated as 1644 cm–1, which is in the low end of the range for symmetric bridging carbonyls (1850–1600 cm–1)82 but in good agreement with the experimental value of 1655 cm–1 in 7. The disfavoring of a symmetrically bridged CO is attributable to the absence of a high-lying orbital of suitable symmetry for back-donation in 1.56
Figure 10.

Optimized geometry of VII.
Thermal Transformations of Ti2Pn†2CO2
The fate of 4 upon warming depends on whether the solution is sealed or under dynamic vacuum. Removing CO from the reaction leads to the formation of 5, which undergoes further reaction to give its isomer 6. A possible intermediate, undetected as yet experimentally, in the decomposition of Ti2Pn†2CO2 (4) is Ti2Pn†2(μ-O)CO, in which a CO bond has broken, the detached O bridges the two Ti atoms, and the CO ligand formed is bonded to one of the Ti atoms. Geometry optimization of Ti2Pn2(μ-O)CO gave a local minimum for such a species, Int1 (Figure 11), in which the Ti–Ti distance (2.46, 2.43 Å) is still indicative of metal–metal bonding. The bridging O is placed asymmetrically, further from the Ti to which the CO is coordinated. The search for a transition state between IV and Int1 yielded structure TS1 shown in Figure 11, with geometric parameters given in Table 2; the transition state also possesses Cs symmetry. Both a Ti–C and a Ti–O bond are breaking and the remaining Ti–O bond has shortened, but the new Ti–O bond to the bridging O has not yet formed.
Figure 11.

Optimized geometries of Int1 and TS1.
Loss of CO from Ti2Pn2O(CO) (Int1) leads to a mono(oxo) product. Maintenance of the double-sandwich structure of the Ti2Pn2 fragment leads to a local minimum with C2v symmetry, structure V. Two other structures were found with η8 coordination of Pn to Ti, one with a triplet state, VIt, and the other with a singlet state, VIs (Figure 12). The structures found for the triplet state by the two methods differed in the Ti–O–Ti angle: the ADF (BP/TZP) calculations optimized to a bent angle, while the Gaussian (B3LYP/SDD) calculations gave a linear Ti–O–Ti unit. Similar structures with η8 coordination were found for the singlet state by the two methods (Table 2). The energies of all three structures are close, and which structure is the most stable is method- and temperature-dependent (Table 3): ADF (BP/TZP) has the double-sandwich structure as the most stable, while Gaussian (B3LYP/SDD) estimates the SCF energy of the sandwich structure to be the lowest but on the basis of the free energy at 298 K predicts the triplet η8-coordinated structure to be the most stable. This is in agreement with experiment, as the double-sandwich structure 5 is known to convert to the triplet state 6 at room temperature. The large ring slippage and twist angle found in the solid-state structure of 5 may be construed as a “snapshot” of the Pn† rings as they move across the Ti2 core from a bridging (μ-η5,η5) bonding mode in 5 toward a capping (η8) mode in the structural isomer 6.
Figure 12.

Optimized geometries of the mono(oxo) species V, VIt, and VIs.
Table 3. Ti–O–Ti Angles (deg) and Relative Energies (kcal mol–1) of Structures Found for Ti2Pn2O56,a.
| structure | Ti–O–Ti | ΔE(SCF) | ΔH298 K° | ΔG298 K° |
|---|---|---|---|---|
| V | 79, 79 | 0, 0 | 0, 0 | 0, 2 |
| VIt | 133, 180 | 12, 4 | 13, 3 | 8, 0 |
| VIs | 103, 100 | 12, 19 | 13, 21 | 10, 20 |
ADF (BP/TZP) values are given in normal text, and Gaussian (B3LYP/SDD) values are given in italics.
The free energy profile found for the binding of CO2 to I and the transformation of IV when CO is removed is given in Figure 13 (the reaction pathway of V to form VIt was too complex to model). Binding of CO2 to I is strongly favored, but the activation energy calculated for the decomposition of IV (22, 13 kcal mol–1) is reasonably accessible at room temperature.
Figure 13.

Free energies for binding of CO2 to I and the decomposition of IV under dynamic vacuum. ADF (BP/TZP) values are given in red and Gaussian (B3LYP/SDD) values in black.
The decomposition of 4 to 6 via 5 when CO is removed and the detection of 5 when CO is present strongly suggest that a similar route is followed when the decomposition proceeds further in the presence of CO (see Scheme 1). DFT calculations were used to examine the possible pathways for the formation of 2 and 3 (eqs 1–5 below; in the chemical formulas, the Pn ligands have been omitted; numbers in Roman numerals refer to the optimized, calculated structures in Table 2; numbers at the right are ADF, Gaussian reaction free energies).
| 1 |
| 2 |
| 3 |
| 4 |
| 5 |
Consideration of eqs 1 and 2 suggests that VI is oxophilic and attacks either V or the proposed oxocarbonyl intermediate (Int1) to form the bis(oxo) complex II. The more favorable values found for the reaction shown in eq 2 indicate that the presence of CO helps drive the disproportionation, explaining the lack of further decomposition when CO is removed from the reaction. Binding of a further molecule of CO to form the dicarbonyl III (eq 4) is favorable but not as exothermic as forming the monocarbonyl (eq 3). The overall decomposition is found to be thermodynamically favorable (eq 5). That the reaction appears to be first order suggests that the rate-determining step is the initial rupture of the bound CO2 to form the proposed intermediate, rather than any subsequent disproportionation process, which would result in a second-order or more complex rate expression. Figure 14 shows the energies of intermediates involved in a possible disproportionation route. The experimental activation energy (ΔG298 K⧧ = 22.4 ± 1.5 kcal mol–1) is in good agreement with the ADF (BP/TZP)-calculated activation energy for the decomposition of IV (22 kcal mol–1; vide supra).
Figure 14.
Calculated free energies of intermediates in the disproportionation of Ti2Pn2O(CO) (Int1). Zero represents the energy of Ti2Pn2CO2 (IV). ADF (BP/TZP) values are shown in red and Gaussian (B3LYP/SDD) values in black.
In terms of the final products of the reaction of I with CO2, geometry optimization of Ti2Pn2(CO)2 (III) by both ADF (BP/TZP) and Gaussian (B3LYP/SDD) gave a structure of Cs symmetry only slightly displaced from C2v symmetry (Figure 15), which agrees well with that found experimentally. The calculated wavenumbers for the antisymmetric and symmetric CO stretches in III are 1947, 1899 cm–1 and 1878, 1810 cm–1, respectively, which are higher than the ν(CO) values found experimentally in 3 (1910 and 1987 cm–1). The optimized structure of II had C2h symmetry (Figure 15) with eclipsed pentalene rings and the Ti2O2 plane at an angle of 22° with respect to the pentalene bridging carbons.
Figure 15.

Optimized geometries of II and III.
Preliminary trials on an NMR scale to observe the proposed mono(μ-oxo) mono(CO) intermediate by reaction of [13C]7 with 1 equiv of N2O or 5 with 1 equiv of 13CO were inconclusive. The 13C{1H} NMR spectra were complex in each case, with the dicarbonyl complex [13C]3 identified as the major labeled product after warming to room temperature. 1H NMR spectra confirmed the presence of bis(oxo) 2 in each case, which is a thermodynamic sink in the reaction of 1 with CO2 because of the formation of strong Ti–O bonds.
Reaction of COS with Ti2Pn†2
Support for the proposed intermediate comes from the reaction sequence observed when COS is added to 1. The first formed species, 9, shows a CO stretch at 1498 cm–1, which is in good agreement with the stretching vibration at 1487 cm–1 for model system IX (Figure 16) calculated by DFT methods.56
Figure 16.

Optimized geometries of IX, X, and XI.
The free energy of the reaction of I with COS is intermediate between that of CS2 (−52 kcal mol–1) and CO2 (−37 kcal mol–1) (eq 6).
| 6 |
The formation of X, from IX, is calculated to be slightly favorable (eq 7) with an activation energy of 32, 21 kcal mol–1. The value found for the equivalent reaction to Int1 is 22, 13 kcal mol–1. Loss of CO from X is unfavorable (eq 8) and more so than from Int1 (5, 11 kcal mol–1).
| 7 |
| 8 |
The mono(μ-S) mono(CO) intermediate 10 was too unstable to be isolated in the solid state but was characterized by spectroscopic methods in solution. In situ IR spectroscopy showed a broad band at 2011 cm–1, in the expected region for a terminal CO ligand, that shifted to 1966 cm–1 in the isotopically labeled [13C]11. The calculated structure X (Figure 16) has a CO vibration at 1937 cm–1. The discrepancy here is rather large. The 1H NMR spectrum of 10 consists of four doublets in the aromatic region, indicative of chemically equivalent Pn† rings on the NMR time scale. This can be attributed to a fluxional process involving intramolecular exchange of CO between the two Ti centers that results in an averaged NMR structure of C2 symmetry.
In the presence of excess of COS, green solutions of 10 decompose within hours at room temperature to give a brown solution (Scheme 4). 1H NMR spectroscopy of the brown solution showed near-quantitative conversion to the bis(μ-S)-bridged dimer [(η8-Pn†)Ti(μ-S)]2 (12) as the sole Ti-containing product, in contrast to the decomposition of CO2 adduct 4, which also gives dicarbonyl complex 3. The related oxo species, Ti2Pn2O(CO) (Int1), was implicated in the DFT-modeled pathway of the CO2 reaction but was not observed experimentally.
The free energy of formation of VIII is calculated as −53 kcal mol–1, which is significantly greater than that of IV. The energy to form equivalent decomposition products is unfavorable (eq 9), and we were unable to identify computationally an intermediate equivalent to Int1 where a C–S bond had been ruptured and bound CS formed; instead, the structure reverted to an alternative η2 mode of CS binding.
| 9 |
Conclusions
Complex 1 shows unique reactivity among bis(pentalene)dimetal double-sandwich complexes, including multiple adduct formation with carbon monoxide and reductive activation of carbon dioxide. Mechanistic studies show that 1 binds CO2 at low temperatures, but this adduct 4 is thermally unstable and upon warming to room temperature, reaction of the coordinated CO2 molecule occurs. We have gained significant insight into the mechanism for this reaction using synthetic, spectroscopic, and computational methods to identify key intermediates. Thermally driven cleavage of the bound CO2 occurs to give oxo and carbonyl fragments, and exposure of 4 to dynamic vacuum removes CO, allowing the isolation of a diamagnetic mono(oxo) complex, 5. The latter shows a strained three-membered “dimetallaepoxide” motif in its molecular structure and is unstable at room temperature with respect to the paramagnetic, bridging mono(oxo) complex 6, in which the TiIII centers are better stabilized, electronically and sterically, with η8-bound pentalene ligands.
The reactivity of 1 with the heteroallenes CS2 and COS provides useful insight in comparison with that observed for CO2. The thermally stable CS2 adduct 8 shows symmetrical binding to the Ti2 unit, providing supporting evidence for the doubly bridging CO2 ligand in 4. 1 displays greater, and different, reactivity toward COS, but some parallels can be made with the CO2 reaction. The asymmetric COS molecule undergoes scission at low temperatures, forming the mono(μ-S) mono(CO) species 10, which has sufficient stability for spectroscopic characterization. A mono(μ-O) mono(CO) counterpart was implicated in the CO2 reaction pathway but was not observed experimentally.
Experimental Section
General Procedures
All manipulations were carried out using standard Schlenk techniques under Ar or in an MBraun glovebox under N2 or Ar. All glassware was dried at 160 °C overnight prior to use. Solvents were purified by predrying over sodium wire and then distilled over Na (toluene), K (methylcyclohexane), or Na–K alloy (Et2O and pentane) under a N2 atmosphere. Dried solvents were collected, degassed, and stored over argon in K-mirrored ampules. Deuterated solvents were degassed by three freeze–pump–thaw cycles, dried by refluxing over K for 3 days, vacuum-distilled into ampules, and stored under N2. The gases used were of very high purity—CO (99.999%) and CO2 (99.99%) were supplied by Union Carbide, COS (>97.5%) was supplied by Aldrich, N2O (purity >99.998%) was supplied by Fluka—and all gases were added via Toepler line. Isotopically enriched gases, 13CO (99.7%) and 13CO2 (99%), were supplied by Euriso-top and Cambridge Isotopes, respectively, and added via Toepler pump. Triphenylphosphine sulfide (98%) was purchased from Aldrich and used as received. The compounds Ti2Pn†2 (1),54 Ti2Pn†2O2 (2), and Ti2Pn†2(CO)2 (3)55 were prepared according to published procedures. NMR spectra were recorded on a Varian VNMRS 400 spectrometer (1H, 399.5 MHz; 13C{1H}, 100.25 MHz; 29Si{1H}, 79.4 MHz). The 1H and 13C spectra were referenced internally to the residual protic solvent (1H) or the signals of the solvent (13C). 29Si{1H} NMR spectra were referenced externally relative to SiMe4. IR spectra were recorded between NaCl plates using a PerkinElmer Spectrum One FTIR instrument or a Mettler-Toledo ReactIR system featuring an IR probe inside a gas-tight cell attached to a Toepler pump. Mass spectra were recorded using a VG Autospec Fisons instrument (EI at 70 eV). Elemental analyses for were carried out at the Elemental Analysis Service, London Metropolitan University.
Syntheses
Ti2(μ:η5,η5-Pn†)2CO2 (4)
A J. Young NMR tube was charged with 1 (16 mg, 0.017 mmol) and methylcyclohexane-d14 (0.5 mL), and the solution was cooled to −78 °C. The headspace was evacuated, and CO2 (0.017 mmol) was admitted. The tube was briefly shaken, and a color change from deep red to dark green was observed. NMR spectra were immediately measured, with the probe precooled to −30 °C. Alternatively, using 13CO2 in the method above yielded the labeled product [13C]4. NMR yield: quantitative with respect to 1. 1H NMR (methylcyclohexane-d14, 399.5 MHz, 243 K): δH 5.98 (2H, d, 3JHH = 3.6 Hz, Pn H), 5.83 (2H, d, 3JHH = 2.9 Hz, Pn H), 5.66 (2H, d, 3JHH = 2.8 Hz, Pn H), 5.37 (2H, d, 3JHH = 3.6 Hz, Pn H), 0.69 (6H, dt, 3JHH = 7.4, 14.9 Hz, iPr CH), 0.57 (6H, dt, 3JHH = 7.4, 14.7 Hz, iPr CH), 0.34 (18H, d, 3JHH = 7.3 Hz, iPr CH3), 0.31 (18H, d, 3JHH = 7.5 Hz, iPr CH3), 0.20 (18H, d, 3JHH = 7.4 Hz, iPr CH3), 0.16 (18H, d, 3JHH = 7.4 Hz, iPr CH3). 13C{1H} NMR (methylcyclohexane-d14, 100.5 MHz, 243 K, selected data): δC 219 (s, CO2). 29Si{1H} NMR (methylcyclohexane-d14, 79.4 MHz, 243 K): δSi 2.12, 1.88. IR (methylcyclohexane, −65 °C): 4 1678 (νasym(OCO)), 1236 (νsym(OCO)) cm–1; [13C]4 1637 (νasym(O13CO)), 1217 (νsym(O13CO)) cm–1. EI-MS and microanalysis data were not obtained for 4 because of its susceptibility to further reaction above 0 °C.
Ti2(μ:η5,η5-Pn†)2(μ-O) (5)
Method A: A J. Young NMR tube was charged with 1 (17 mg, 0.018 mmol) and C6D6 (0.6 mL), and the solution was frozen at −78 °C. The headspace was evacuated, and N2O (0.016 mmol) was admitted. The tube was warmed to room temperature, and a color change to bright red was observed as the solution thawed. NMR yield: 68% with respect to 1. Method B: An ampule charged with 1 (78.5 mg, 0.0848 mmol) and methylcyclohexane (5 mL) was cooled to −78 °C, and the headspace was evacuated. CO2 gas (1 atm) was admitted, resulting in a color change from deep red to dark green. The ampule was placed under reduced pressure and then removed from the −78 °C bath. The solvent was slowly removed in vacuo as the reaction mixture warmed to room temperature, giving a crude red residue. Recrystallization from toluene (1 mL) at −35 °C afforded 5 as a red microcrystalline solid. Total yield: 66.3 mg (83% with respect to 1). Subsequent recrystallization from a concentrated toluene/pentane solution at −35 °C afforded X-ray-quality crystals. 1H NMR (C6D6, 399.5 MHz, 303 K, selected data): δH 7.86 (2H, d, 3JHH= 3.3 Hz, Pn H), 6.78 (2H, d, 3JHH = 3.2 Hz, Pn H), 6.45 (2H, d, 3JHH= 3.2 Hz, Pn H), 6.39 (2H, d, 3JHH = 3.2 Hz, Pn H), 1.55 (6H, m, iPr CH), 1.26 (6H, m, iPr CH), 1.13 (18H, d, 3JHH= 7.2 Hz, iPr CH3), 1.10 (18H, d, 3JHH= 7.4 Hz, iPr CH3), 1.01 (18H, d, 3JHH= 7.4 Hz, iPr CH3), 0.96 (18H, d, 3JHH= 7.4 Hz, iPr CH3). 13C{1H} NMR (C6D6, 100.5 MHz, 303 K, selected data): δC 134.2 (Pn C), 133.4 (Pn C), 132.6 (Pn C), 114.9 (Pn C), 111.3 (Pn C), 108.1 (Pn C), 102.2 (Pn C), 99.47 (Pn C), 19.94 (iPr CH), 19.89 (iPr CH), 19.39 (iPr CH3), 19.35 (iPr CH3), 14.11 (iPr CH), 12.97 (iPr CH). 29Si{1H} NMR (C6D6, 303 K): δSi 2.39, 1.01. EI-MS: m/z 941 (100%) [M]+. Anal. Found (Calcd for C52H92OSi4Ti2): C, 66.21 (66.35); H, 9.70 (9.85) %.
[(η8-Pn†)Ti]2(μ-O) (6)
Storage of a C6D6 solution of 5 (6.6 μmol) overnight at room temperature afforded green crystals of 6, which were isolated by decantation, washed with cold pentane (1 mL, −35 °C) and dried in vacuo. Total yield: 2.0 mg (32% with respect to 5). 1H NMR (toluene-d8, 399.5 MHz, 303 K): δH 4.09 (br, Δν1/2 = 138 Hz, Pn H), 4.09 (br, Δν1/2 = 138 Hz, Pn H), 1.77 (br, Δν1/2 = 110 Hz, iPr CH), 1.13 (br, Δν1/2 = 106 Hz, iPr CH3). 13C and 29Si NMR resonances were not observed because of the paramagnetic nature of 6. EI-MS: m/z 941 (100%) [M]+. Anal. Found (Calcd for C52H92OSi4Ti2): C, 66.17 (66.35); H, 9.75 (9.85) %. Magnetic susceptibility (Evans method, THF-d8, 303 K): μeff = 2.45μB per molecule = 1.73μB per Ti.
Ti2(μ:η5,η5-Pn†)2(μ:η2-CO) (7)
An ampule charged with 1 (50 mg, 0.054 mmol) and toluene (0.5 mL) was cooled to −78 °C, and the headspace was evacuated. CO gas (0.9 molar equiv, 0.050 mmol) was admitted slowly to the mixture with vigorous stirring, resulting in a color change from deep red to purple. After the mixture was warmed to room temperature, the volatiles were removed in vacuo to afford a crude purple solid that was recrystallized from pentane (1 mL) at −35 °C. Total yield: 32 mg (62% with respect to 1). Alternatively, exposing a methylcyclohexane-d14 solution of 1 to 0.9 equiv of 13CO via Toepler pump yielded the labeled product [13C]7. 1H NMR (methylcyclohexane-d14, 399.5 MHz, 303 K): δH 7.69 (1H, apparent t, 3JHH = 3.2 Hz, Pn H), 7.10 (1H, d, 3JHH = 3.2 Hz, Pn H), 6.62 (1H, d, 3JHH = 3.5 Hz, Pn H), 6.34 (1H, d, 3JHH = 3.5 Hz, Pn H), 6.30 (2H, d, 3JHH = 3.2 Hz, Pn H), 6.22 (2H, d, 3JHH = 3.6 Hz, Pn H), 5.74 (1H, d, 3JHH = 3.5 Hz, Pn H), 5.62 (1H, d, 3JHH = 3.2 Hz, Pn H), 1.62–1.25 (12H, overlapping m, iPr CH), 1.25–0.83 (72H, overlapping m, iPr CH3). 13C{1H} NMR (methylcyclohexane-d14, 100.5 MHz, 303 K): δC 295.1 (s, CO), 136.8 (Pn C), 132.5 (Pn C), 132.4 (Pn C), 129.1 (Pn C), 127.7 (Pn C), 125.1 (Pn C), 124.9 (Pn C), 115.1 (Pn C), 108.2 (Pn C), 104.7 (Pn C), 103.2 (Pn C), 101.5 (Pn C), 99.75 (Pn C), 95.08 (Pn C), 93.16 (Pn C), 89.87 (Pn C), 20.94 (iPr CH3), 20.84 (iPr CH3), 20.79 (iPr CH3), 20.64 (iPr CH3), 20.62 (iPr CH3), 20.37 (iPr CH3), 20.36 (iPr CH3), 20.13 (iPr CH3), 15.14 (iPr CH), 15.12 (iPr CH), 14.45 (iPr CH), 14.22 (iPr CH). 29Si{1H} NMR (methylcyclohexane-d14, 79.4 MHz, 303 K): δSi 3.40 (s), 2.88 (d, JSiC = 1.2 Hz), 2.40 (s), 2.22 (s). EI-MS: m/z 924 (100%) [M – CO]+, 882 (10%) [M – CO – iPr]+. Anal. Found (Calcd for C53H92OSi4Ti2): C, 66.78 (66.77); H, 9.81 (9.73) %. IR (NaCl): 7 1655 (br, ν(CO)); [13C]7 1616 (br, ν(13CO)) cm–1.
Ti2(μ:η5,η5-Pn†)2(μ:η2,η2-CS2) (8)
To a solution of 1 (58 mg, 0.058 mmol) in pentane (2 mL) at −35 °C was added CS2 (14 μL, 0.23 mmol) via microsyringe. The resultant dark-green mixture was allowed to warm to room temperature and stir for 3 h. Filtration followed by cooling to −35 °C afforded red microcrystals that were isolated by decantation and dried in vacuo. Total yield: 36 mg (62% with respect to 1). Recrystallization from a saturated Et2O solution at −35 °C afforded dark-red crystals suitable for X-ray diffraction studies. Alternatively, recrystallization from a saturated SiMe4 solution at −35 °C afforded red crystals with different unit cell parameters (see Table S1 in the Supporting Information). 1H NMR (toluene-d8, 399.5 MHz, 303 K): δH 7.30 (2H, 3JHH = 3.0 Hz, Pn H), 6.35 (2H, 3JHH = 3.0 Hz, Pn H), 6.25 (2H, 3JHH = 3.5 Hz, Pn H), 5.85 (2H, 3JHH = 3.5 Hz, Pn H), 1.48 (6H, m, iPr CH), 1.42 (6H, m, iPr CH), 1.15 (18H, d, 3JHH = 7.4 Hz, iPr CH3), 1.13 (18H, d, 3JHH = 7.5 Hz, iPr CH3), 1.05 (18H, d, 3JHH = 7.4 Hz, iPr CH3), 1.01 (18H, d, 3JHH = 7.5 Hz, iPr CH3). 13C{1H} NMR (toluene-d8, 100.5 MHz, 303 K): δC 355.5 (CS2), 133.4 (Pn C), 132.4 (Pn C), 127.1 (Pn C), 107.6 (Pn C), 105.5 (Pn C), 95.1 (Pn C), 20.37 (iPr CH3), 20.21 (iPr CH3), 19.82 (iPr CH3), 14.73 (iPr CH), 12.93 (iPr CH). 29Si{1H} NMR (toluene-d8, 79.4 MHz, 303 K): δSi 7.65, 7.39. EI-MS: m/z 1002 (65%) [M]+, 960 (30%) [M – iPr]+. Anal. Found (Calcd for C53H92S2Si4Ti2): C, 63.49 (63.56); H, 9.26 (9.26) %. IR (methylcyclohexane, 26 °C): 1104 (νasym(SCS)) cm–1.
Ti2(μ:η5,η5-Pn†)2COS (9)
A J. Young NMR tube was charged with 1 (20 mg, 0.022 mmol) and methylcyclohexane-d14 (0.6 mL), and the solution was cooled to −78 °C. The headspace was evacuated, and COS (0.048 mmol) was admitted. The tube was briefly shaken, and a color change from deep red to green was observed. NMR spectra were immediately measured, with the probe precooled to −30 °C. NMR yield: 45% with respect to 1. 1H NMR (methylcyclohexane-d14, 399.5 MHz, 243 K, selected data): δH 7.00 (1H, d, 3JHH = 3.0 Hz, Pn H), 6.96 (1H, d, 3JHH = 3.0 Hz, Pn H), 6.72 (1H, d, 3JHH = 3.7 Hz, Pn H), 6.45 (1H, d, 3JHH = 3.0 Hz, Pn H), 6.43 (1H, d, 3JHH = 3.6 Hz, Pn H), 6.31 (1H, d, 3JHH = 3.7 Hz, Pn H), 6.09 (1H, d, 3JHH = 3.7 Hz, Pn H), 1.79–1.44 (12H, overlapping m, iPr CH), 1.34–1.02 (72H, overlapping m, iPr CH3). One Pn H signal was not resolved because of coincidental overlap with signals of 12. 13C{1H} NMR (methylcyclohexane-d14, 100.5 MHz, 243 K, selected data): δC 283 (s, COS). 29Si{1H} NMR (methylcyclohexane-d14, 79.4 MHz, 243 K, selected data): δSi 2.48, 2.07, 1.73, 1.08. IR (NaCl): 1493 (νasym(SCO)) cm–1. IR (methylcyclohexane, −65 °C): 1498 (νasym(SCO)) cm–1. EI-MS and microanalysis data were not obtained for 9 because of its susceptibility to further reaction above −40 °C.
Ti2(μ:η5,η5-Pn†)2S(CO) (10)
Method A: A J. Young NMR tube charged with 1 (16 mg, 0.017 mmol) and toluene-d8 (0.65 mL) at −78 °C was degassed, and COS gas (0.015 mmol) was slowly admitted. The reaction mixture was briefly shaken, resulting in a color change from deep red to green, and the mixture was allowed to warm to room temperature over a period of 2 h. NMR yield: 47% with respect to 1. Method B: An ampule charged with 11 (0.028 mmol) and toluene-d8 (0.7 mL) was cooled to −78 °C, and the headspace was evacuated. 13CO gas (0.025 mmol) was slowly admitted to the stirred mixture, resulting in a color change from red to green. The solution was warmed to room temperature and then decanted into a J. Young tube, and [13C]10 was identified by its NMR and IR spectra. NMR yield: 75% with respect to 1. Alternatively, exposing a solution of 11 to 0.9 molar equiv of CO via Toepler pump yielded primarily the unlabeled product 10. 1H NMR (toluene-d8, 399.5 MHz, 303 K, selected data): δH 7.81 (2H, d, 3JHH = 2.9 Hz, Pn H), 6.09 (br, Δν1/2 = 13 Hz, Pn H), 5.75 (2H, d, 3JHH = 2.9 Hz, Pn H), 5.38 (2H, d, 3JHH = 2.7 Hz, Pn H), 1.62–1.24 (12H, overlapping m, iPr CH), 1.22–0.96 (72H, overlapping m, iPr CH3). 13C{1H} NMR (toluene-d8, 100.5 MHz, 303 K, selected data): δC 247 (br, Δν1/2 = 110 Hz, CO). 29Si{1H} NMR (toluene-d8, 79.4 MHz, 303 K, selected data): δSi 2.56, 1.46. IR (methylcyclohexane, −65 °C): 10 2011 (br, ν(CO)); [13C]10 1966 (br, ν(13CO)) cm–1. EI-MS and microanalysis data were not obtained for 10 because of its susceptibility to further reaction above −40 °C.
Ti2(μ:η5,η5-Pn†)2(μ-S) (11)
A solution of Ph3PS (56 mg, 0.19 mmol) in toluene (3 mL) was added dropwise to a stirred solution of 1 (177 mg, 0.191 mmol) in pentane (3 mL) at −35 °C. The resultant red mixture was allowed to warm to room temperature, stirred for 30 min, and then stripped to dryness. A red residue was obtained, which was placed under high vacuum (ca. 10–6 mbar) and heated at 75 °C for 1 h and then 100 °C for 2 h to ensure complete sublimation of a white crystalline solid, presumably Ph3P. The residue was extracted with pentane (3 × 2 mL), filtered, and allowed to concentrate to ca. 3 mL. Cooling of this solution to −35 °C afforded red crystals of 11 suitable for X-ray diffraction studies. Total yield: 95 mg (52% with respect to 1). 1H NMR (C6D6, 399.5 MHz, 303 K): δH 9.25 (2H, d, 3JHH= 3.4 Hz, Pn H), 7.06 (2H, d, 3JHH = 3.2 Hz, Pn H), 6.10 (2H, d, 3JHH= 3.1 Hz, Pn H), 5.96 (2H, d, 3JHH = 3.4 Hz, Pn H), 2.01 (6H, m, iPr CH), 1.35 (6H, m, iPr CH), 1.17 (18H, d, 3JHH= 7.5 Hz, iPr CH3), 1.13 (18H, d, 3JHH= 7.3 Hz, iPr CH3), 0.98 (18H, d, 3JHH= 7.5 Hz, iPr CH3), 0.60 (18H, d, 3JHH= 7.5 Hz, iPr CH3). 13C{1H} NMR (C6D6, 100.5 MHz, 303 K): δC 141.7 (Pn C), 132.5 (Pn C), 130.0 (Pn C), 123.0 (Pn C), 118.6 (Pn C), 115.3 (Pn C), 113.2 (Pn C), 98.53 (Pn C), 19.50 (iPr CH3), 19.41 (iPr CH3), 19.14 (iPr CH3), 18.89 (iPr CH3), 13.72 (iPr CH), 13.15 (iPr CH). 29Si{1H} NMR (C6D6, 79.4 MHz, 303 K): δSi 2.20, 1.82. EI-MS: m/z 989 (60%) [M + S]+, 956–959 (principal peak 957, 20%) [M]+, 877 (100%) [M – Ti – S]+. Anal. Found (Calcd for C52H92SSi4Ti2): C, 65.33 (65.23); H, 9.61 (9.69) %.
[(η8-Pn†)Ti(μ-S)]2 (12)
Method A: An ampule containing a solution of 1 (141 mg, 0.152 mmol) in benzene (2 mL), frozen at −35 °C, was degassed and fitted with a balloon containing COS gas (1 atm). The solution was allowed to warm to room temperature and, shortly after thawing, appeared green in color. After 30 min of stirring at room temperature, a color change to red-brown was observed. The reaction mixture was frozen and lyophilized, and a spectroscopically pure red-brown solid was isolated. Total yield: 145 mg (96% with respect to 1). Method B: To an ampule containing 1 (100 mg, 0.108 mmol) and Ph3PS (64 mg, 0.216 mmol) was added toluene (2 mL). After 4 h of stirring at room temperature, the red-brown mixture was stripped to dryness, extracted with Et2O (2 × 5 mL), and filtered, and the solvent was removed in vacuo. A brown solid was obtained, which was placed under high vacuum (ca. 10–6 mbar) and heated at 60 °C for 2.5 h, 75 °C for 1 h, and finally 100 °C for 30 min to ensure complete sublimation of a white crystalline solid, presumably Ph3P. The residue was extracted with Et2O (2 × 5 mL), filtered, and allowed to concentrate to ca. 4 mL. Cooling of this solution to −35 °C afforded red crystals suitable for X-ray diffraction studies. Total yield: 86 mg (80% with respect to 1). 1H NMR (C6D6, 399.5 MHz, 303 K): δH 6.59 (4H, d, 3JHH = 3.3 Hz, Pn H), 5.92 (4H, d, 3JHH = 3.2 Hz, Pn H), 1.31 (12H, m, iPr CH), 1.19 (36H, d, 3JHH = 7.3 Hz, iPr CH3), 1.14 (36H, d, 3JHH = 7.3 Hz, iPr CH3). 13C{1H} NMR (C6D6, 100.5 MHz, 303 K): δC 148.9 (Pn C), 130.0 (Pn C), 115.2 (Pn C), 113.9 (Pn C), 19.66 (iPr CH3), 19.50 (iPr CH3), 12.31 (iPr CH). 29Si{1H} NMR (C6D6, 79.4 MHz, 303 K): δSi −0.15. EI-MS: m/z 989 (100%) [M]+, 946 (50%) [M – iPr]+. Anal. Found (Calcd for C52H92S2Si4Ti2): C, 63.20 (63.12); H, 9.44 (9.37) %.
Kinetic Studies
Decomposition rates of 4 at seven temperatures between 9 to 40 °C were measured by the same method: A J. Young NMR tube was charged with 0.50 mL of a stock solution containing 1 (0.0295 M, 0.015 mmol) and ferrocene (0.0935 M) in methylcyclohexane-d14, and the solution was cooled to −78 °C. The headspace was evacuated, and 13CO2 (0.015 mmol) was admitted. The tube was briefly shaken, and a color change from deep red to dark green was observed. After 30 min the tube was removed from the −78 °C bath, placed in a pre-equilibrated VT-NMR probe, and shimmed as thermal equilibrium was attained (10 min). To avoid any complications due to different relaxation times for the resonances, spectra were recorded using one 90° pulse. 1H spectra were acquired periodically (33.6 s repetition time), and the time of each spectrum was recorded along with the data by the internal clock on the spectrometer. A baseline correction was applied, and the concentrations of 4 were determined by integration over the same region of each spectrum relative to the ferrocene resonance. The temperature in the probe was calibrated by measuring the peak separation of a sample of neat methanol.83 Concentration versus time data were analyzed by fitting the data for 4 to a first-order exponential decay. Plots of ln([4]) versus time and the derived rate constants are included in the Supporting Information. The activation parameters were calculated from a linear fit of the Eyring plot, ln(kobs/T) versus 1/T. The error in the determination of the activation parameters was calculated according to Girolami and co-workers.84
Crystallographic Details
Single-crystal XRD data for 7, 8 (SiMe4), 11, and 12 were collected by the U.K. National Crystallography Service (NCS)85 at the University of Southampton on a Rigaku FR-E+ Ultra High Flux diffractometer (λMo Kα) equipped with VariMax VHF optics and a Saturn 724+ CCD area detector. The data were collected at 100 or 150 K using an Oxford Cryosystems Cobra low-temperature device. Data collected by the NCS were processed using CrystalClear-SM Expert 3.1 b18,86 and unit cell parameters were refined against all data. Single-crystal XRD data for 5 and 8 (Et2O) were collected at the University of Sussex on a Enraf-Nonius CAD4 diffractometer with graphite-monochromatized (λMo Kα) radiation or an Agilent Technologies Xcalibur Gemini Ultra diffractometer (λMo Kα or λCu Kα source) equipped with an Eos CCD area detector. The data were collected at 173 K using an Oxford Cryosystems Cobra low-temperature device. Data were processed using KappaCCD software or CrysAlisPro (version 1.171.36.32),87 and unit cell parameters were refined against all data. An empirical absorption correction was carried out using the Multi-Scan program.88 All of the structures were solved using SHELXL-2013,89 DIRDIF-2008,90 or SUPERFLIP91 and refined on Fo2 by full-matrix least-squares refinements using SHELXL-2013.89 Solutions and refinements were performed using the OLEX292 or WinGX93 packages and software packages within. All non-hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were refined using a riding model. Disordered solvent molecules in 8 (Et2O) and 12 could not be modeled properly; therefore, this disorder was treated using the SQUEEZE94 routine in PLATON.95 Disordered solvent molecules in 11 were treated using the solvent mask routine in OLEX2,92 which resulted in improved refinement indices relative to SQUEEZE, so these data have been included.
Computational Details
DFT calculations were carried out using two computational methods. One employed the Amsterdam Density Functional package (version ADF2012.01).96 The Slater-type orbital (STO) basis sets were of triple-ζ quality augmented with one polarization function (ADF basis TZP). Core electrons were frozen (C 1s; Ti 2p) in the model of the electronic configuration for each atom. The local density approximation (LDA) of Vosko, Wilk, and Nusair (VWN)97 was used together with the exchange–correlation corrections of Becke and Perdew (BP86).98,99 The other used Gaussian 09, revision A.02,100 with the B3LYP functional and SDD basis set. In both sets of calculations, tight optimization conditions were used and frequency calculations were used to confirm stationary points.
Acknowledgments
We thank the European Research Council and the University of Sussex for financial support. Dr. S. M. Roe (Sussex), Dr. B. M. Day (Manchester), and NCS Southampton are thanked for their assistance with X-ray crystallography.
Supporting Information Available
Additional NMR spectroscopic, X-ray crystallographic, and kinetic data; a text file of all computed molecule Cartesian coordinates in .xyz format for convenient visualization; crystallographic data in CIF format. The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.5b00315.
The authors declare no competing financial interest.
Supplementary Material
References
- IPCC. Climate Change 2013: The Physical Science Basis; Stocker T. F., Qin D., Plattner G.-K., Tignor M. M. B., Allen S. K., Boschung J., Nauels A., Xia Y., Bex V., Midgley P. M., Eds.; Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, Summary for Policymakers, 2013. [Google Scholar]
- Carbon Dioxide as Chemical Feedstock; Aresta M., Ed.; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
- Sakakura T.; Choi J.-C.; Yasuda H. Chem. Rev. 2007, 107, 2365–2387. [DOI] [PubMed] [Google Scholar]
- Arakawa H.; Aresta M.; Armor J. N.; Barteau M. A.; Beckman E. J.; Bell A. T.; Bercaw J. E.; Creutz C.; Dinjus E.; Dixon D. A.; et al. Chem. Rev. 2001, 101, 953–996. [DOI] [PubMed] [Google Scholar]
- West N. M.; Miller A. J.; Labinger J. A.; Bercaw J. E. Coord. Chem. Rev. 2011, 255, 881–898. [Google Scholar]
- Williams V. A.; Manke D. R.; Wolczanski P. T.; Cundari T. R. Inorg. Chim. Acta 2011, 369, 203–214. [Google Scholar]
- Cokoja M.; Bruckmeier C.; Rieger B.; Herrmann W. A.; Kühn F. E. Angew. Chem., Int. Ed. 2011, 50, 8510–8537. [DOI] [PubMed] [Google Scholar]
- Yin X.; Moss J. R. Coord. Chem. Rev. 1999, 181, 27–59. [Google Scholar]
- Leitner W. Coord. Chem. Rev. 1996, 153, 257–284. [Google Scholar]
- Gibson D. H. Chem. Rev. 1996, 96, 2063–2096. [DOI] [PubMed] [Google Scholar]
- Lu Z.; Wang Y.; Liu J.; Lin Y.-J.; Li Z. H.; Wang H. Organometallics 2013, 32, 6753–6758. [Google Scholar]
- Ashley A.; O’Hare D. Top. Curr. Chem. 2013, 334, 191–217. [DOI] [PubMed] [Google Scholar]
- Mömming C. M.; Otten E.; Kehr G.; Fröhlich R.; Grimme S.; Stephan D. W.; Erker G. Angew. Chem., Int. Ed. 2009, 48, 6643–6646. [DOI] [PubMed] [Google Scholar]
- Evans W. J.; Perotti J. M.; Brady J. C.; Ziller J. W. J. Am. Chem. Soc. 2003, 125, 5204–5212. [DOI] [PubMed] [Google Scholar]
- Evans W. J.; Seibel C. A.; Ziller J. W. Inorg. Chem. 1998, 37, 770–776. [Google Scholar]
- Lam O. P.; Meyer K. Polyhedron 2012, 32, 1–9. [Google Scholar]
- Gardner B. M.; Liddle S. T. Eur. J. Inorg. Chem. 2013, 3753–3770. [Google Scholar]
- La Pierre H. S.; Meyer K. Prog. Inorg. Chem. 2014, 58, 303–416. [Google Scholar]
- Berthet J.-C.; Le Maréchal J.-F.; Nierlich M.; Lance M.; Vigner J.; Ephritikhine M. J. Organomet. Chem. 1991, 408, 335–341. [Google Scholar]
- Brennan J. G.; Andersen R. A.; Zalkin A. Inorg. Chem. 1986, 25, 1761–1765. [Google Scholar]
- Castro L.; Labouille S.; Kindra D. R.; Ziller J. W.; Nief F.; Evans W. J.; Maron L. Chem.—Eur. J. 2012, 18, 7886–7895. [DOI] [PubMed] [Google Scholar]
- Castro L.; Lam O. P.; Bart S. C.; Meyer K.; Maron L. Organometallics 2010, 29, 5504–5510. [Google Scholar]
- Castro L.; Maron L. Chem.—Eur. J. 2012, 18, 6610–6615. [DOI] [PubMed] [Google Scholar]
- Castro-Rodriguez I.; Nakai H.; Zakharov L. N.; Rheingold A. L.; Meyer K. Science 2004, 305, 1757–1759. [DOI] [PubMed] [Google Scholar]
- Tsoureas N.; Castro L.; Kilpatrick A. F. R.; Maron L.; Cloke F. G. N. Chem. Sci. 2014, 5, 3777–3788. [Google Scholar]
- Laitar D. S.; Müller P.; Sadighi J. P. J. Am. Chem. Soc. 2005, 127, 17196–17197. [DOI] [PubMed] [Google Scholar]
- Chirik P. J. Organometallics 2010, 29, 1500–1517. [Google Scholar]
- Chirik P. J.; Bouwkamp M. W. In Comprehensive Organometallic Chemistry III; Mingos D. M. P., Crabtree R. H., Eds.; Elsevier: Oxford, U.K., 2007; pp 243–279. [Google Scholar]
- Alt H. G.; Schwind K.-H.; Rausch M. D. J. Organomet. Chem. 1987, 321, C9–C12. [Google Scholar]
- Demerseman B.; Mah R.; Dixneuf P. H. J. Chem. Soc., Chem. Commun. 1984, 1394–1396. [Google Scholar]
- Burlakov V. V.; Polyakov A. V.; Yanovsky A. I.; Struchkov Y. T.; Shur V. B.; Vol’pin M. E.; Rosenthal U.; Görls H. J. Organomet. Chem. 1994, 476, 197–206. [Google Scholar]
- Lefeber C.; Ohff A.; Tillack A.; Baumann W.; Kempe R.; Burlakov V. V.; Rosenthal U.; Görls H. J. Organomet. Chem. 1995, 501, 179–188. [Google Scholar]
- Burlakov V. V.; Yanovsky A. I.; Struchkov Y. T.; Rosenthal U.; Spannenberg A.; Kempe R.; Ellert O. G.; Shur V. B. J. Organomet. Chem. 1997, 542, 105–112. [Google Scholar]
- Burlakov V. V.; Rosenthal U.; Dolgushin F. M.; Yanovsky A. I.; Struchkov Y. T.; Ellert O. G.; Shur V. B.; Vol’pin M. E. Metalloorg. Khim. 1992, 5, 1231. [Google Scholar]
- Pellny P.-M.; Burlakov V. V.; Baumann W.; Spannenberg A.; Rosenthal U. Z. Anorg. Allg. Chem. 1999, 625, 910–918. [Google Scholar]
- Fachinetti G.; Floriani C.; Chiesi-Villa A.; Guastini C. J. Am. Chem. Soc. 1979, 101, 1767–1775. [Google Scholar]
- Bottomley F.; Lin I. J. B.; Mukaida M. J. Am. Chem. Soc. 1980, 102, 5238–5242. [Google Scholar]
- Burlakov V. V.; Dolgushin F. M.; Yanovsky A. I.; Struchkov Yu. T.; Shur V. B.; Rosenthal U.; Thewalt U. J. Organomet. Chem. 1996, 522, 241–247. [Google Scholar]
- Santamaría D.; Cano J.; Royo P.; Mosquera M. E. G.; Cuenca T.; Frutos L. M.; Castaño O. Angew. Chem., Int. Ed. 2005, 44, 5828–5830. [DOI] [PubMed] [Google Scholar]
- Fröhlich H.-O.; Schreer H. Z. Chem. 1983, 23, 348–349. [Google Scholar]
- Cotton F. A.; Murillo C. A.; Walton R. A. In Multiple Bonds between Metal Atoms, 3rd ed.; Walton R. A., Ed.; Clarendon Press: Oxford, U.K., 2005. [Google Scholar]
- Lalrempuia R.; Stasch A.; Jones C. Chem. Sci. 2013, 4, 4383–4388. [Google Scholar]
- Noor A.; Qayyum S.; Bauer T.; Schwarz S.; Weber B.; Kempe R. Chem. Commun. 2014, 50, 13127–13130. [DOI] [PubMed] [Google Scholar]
- Palit C. M.; Graham D. J.; Chen C.-H.; Foxman B. M.; Ozerov O. V. Chem. Commun. 2014, 50, 12840–12842. [DOI] [PubMed] [Google Scholar]
- Krogman J. P.; Foxman B. M.; Thomas C. M. J. Am. Chem. Soc. 2011, 133, 14582–14585. [DOI] [PubMed] [Google Scholar]
- Summerscales O. T.; Cloke F. G. N. Coord. Chem. Rev. 2006, 250, 1122–1140. [Google Scholar]
- Katz T. J.; Acton N. J. Am. Chem. Soc. 1972, 94, 3281–3283. [Google Scholar]
- Katz T. J.; Acton N.; McGinnis J. J. Am. Chem. Soc. 1972, 94, 6205–6206. [Google Scholar]
- Kuchta M.; Cloke F. G. N. Organometallics 1998, 17, 1934–1936. [Google Scholar]
- Balazs G.; Cloke F. G. N.; Harrison A.; Hitchcock P. B.; Green J.; Summerscales O. T. Chem. Commun. 2007, 873–875. [DOI] [PubMed] [Google Scholar]
- Balazs G.; Cloke F. G. N.; Gagliardi L.; Green J. C.; Harrison A.; Hitchcock P. B.; Shahi A. R. M.; Summerscales O. T. Organometallics 2008, 27, 2013–2020. [Google Scholar]
- Summerscales O. T.; Rivers C. J.; Taylor M. J.; Hitchcock P. B.; Green J. C.; Cloke F. G. N. Organometallics 2012, 31, 8613–8617. [Google Scholar]
- Ashley A. E.; Cooper R. T.; Wildgoose G. G.; Green J. C.; O’Hare D. J. Am. Chem. Soc. 2008, 130, 15662–15677. [DOI] [PubMed] [Google Scholar]
- Kilpatrick A. F. R.; Green J. C.; Cloke F. G. N.; Tsoureas N. Chem. Commun. 2013, 49, 9434–9436. [DOI] [PubMed] [Google Scholar]
- Kilpatrick A. F. R.; Cloke F. G. N. Chem. Commun. 2014, 50, 2769–2771. [DOI] [PubMed] [Google Scholar]
- Kilpatrick A. F. R.; Green J. C.; Cloke F. G. N. Organometallics 2015, 10.1021/acs.organomet.5b00363. [DOI] [Google Scholar]
- Jegat C.; Fouassier M.; Tranquille M.; Mascetti J. Inorg. Chem. 1991, 30, 1529–1536. [Google Scholar]
- Evans J. J. Chem. Soc. 1959, 2003–2005. [Google Scholar]
- Schubert E. M. J. Chem. Educ. 1992, 69, 62. [Google Scholar]
- Werner H. Coord. Chem. Rev. 1982, 43, 165–185. [Google Scholar]
- Pandey K. K. Coord. Chem. Rev. 1997, 140, 37–114. [Google Scholar]
- Herzberg G.Electronic Spectra and Electronic Structure of Polyatomic Molecules; Molecular Spectra and Molecular Structure, Vol. 3; Van Nostrand-Reinhold: New York, 1966. [Google Scholar]
- Eland J. H. D.; Berkowitz J. J. Chem. Phys. 1979, 70, 5151–5156. [Google Scholar]
- Cox J. D.; Wagman D. D.; Medvedev V. A.. CODATA Key Values for Thermodynamics; Hemisphere: New York, 1989. [Google Scholar]
- CRC Handbook of Chemistry and Physics, 80th ed.; Lide D. R., Ed.; CRC Press: Boca Raton, FL, 1999. [Google Scholar]
- Pinkas J.; Císařová I.; Horáček M.; Kubišta J.; Mach K. Organometallics 2011, 30, 1034–1045. [Google Scholar]
- Gibson D. H.; Ye M.; Richardson J. F. J. Am. Chem. Soc. 1992, 114, 9716–9717. [Google Scholar]
- Gibson D. H.; Mehta J. M.; Ye M.; Richardson J. F.; Mashuta M. S. Organometallics 1994, 13, 1070–1072. [Google Scholar]
- Aresta M.; Dibenedetto A. Dalton Trans. 2007, 2975–2992. [DOI] [PubMed] [Google Scholar]
- Lee C. H.; Laitar D. S.; Mueller P.; Sadighi J. P. J. Am. Chem. Soc. 2007, 129, 13802–13803. [DOI] [PubMed] [Google Scholar]
- Nakamoto K.Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part A, 6th ed.; Wiley: Hoboken, NJ, 2009. [Google Scholar]
- Haack P.; Limberg C.; Tietz T.; Metzinger R. Chem. Commun. 2011, 47, 6374–6376. [DOI] [PubMed] [Google Scholar]
- Fischer P. J.; Young V. G. Jr.; Ellis J. E. Chem. Commun. 1997, 1249–1250. [Google Scholar]
- Dick D. G.; Hou Z.; Stephan D. W. Organometallics 1992, 11, 2378–2382. [Google Scholar]
- Barr R. D.; Green M.; Howard J. A. K.; Marder T. B.; Moore I.; Stone F. G. A. J. Chem. Soc., Chem. Commun. 1983, 746–748. [Google Scholar]
- Awang M. R.; Barr R. D.; Green M.; Howard J. A. K.; Marder T. B.; Stone F. G. A. J. Chem. Soc., Dalton Trans. 1985, 2009. [Google Scholar]
- Dawkins G. M.; Green M.; Mead K. A.; Salaün J.-Y.; Stone F. G. A.; Woodward P. J. Chem. Soc., Dalton Trans. 1983, 527–530. [Google Scholar]
- Horwitz C. P.; Shriver D. F. Adv. Organomet. Chem. 1984, 23, 219–305. [Google Scholar]
- NIST. Diatomic Spectral Database. http://www.nist.gov/pml/data/msd-di/index.cfm (accessed Sept 18, 2014).
- Merola J. S.; Gentile R. A.; Ansell G. B.; Modrick M.; Zentz S. Organometallics 1982, 1, 1731–1733. [Google Scholar]
- Hamilton D. M.; Willis W. S.; Stucky G. D. J. Am. Chem. Soc. 1981, 103, 4255–4256. [Google Scholar]
- Brisdon A. K.Inorganic Spectroscopic Methods; Oxford Chemistry Primers, Vol. 62; Oxford University Press: Oxford, U.K., 1998. [Google Scholar]
- Ammann C.; Meier P.; Merbach A. E. J. Magn. Reson. 1982, 46, 319–321. [Google Scholar]
- Morse P. M.; Spencer M. D.; Wilson S. R.; Girolami G. S. Organometallics 1994, 13, 1646–1655. [Google Scholar]
- Coles S. J.; Gale P. A. Chem. Sci. 2012, 3, 683–689. [Google Scholar]
- CrystalClear; Rigaku: Tokyo, 2011.
- CrysalisPro, version 1.171.36.32; Agilent Technologies: Santa Clara, CA, 2011.
- Blessing R. H. Acta Crystallogr., Sect. A 1995, 51, 33–38. [DOI] [PubMed] [Google Scholar]
- Sheldrick G. M. Acta Crystallogr., Sect. A 2008, 64, 112–122. [DOI] [PubMed] [Google Scholar]
- Beurskens P. T.; Beurskens G.; Gelder R.; Smits J. M. M.; LGarcia-Granda S.; Gould R. O.. DIRDIF-2008, 2008.
- Palatinus L.; Chapuis G. J. Appl. Crystallogr. 2007, 40, 786–790. [Google Scholar]
- Dolomanov O. V.; Bourhis L. J.; Gildea R. J.; Howard J. A. K.; Puschmann H. J. Appl. Crystallogr. 2009, 42, 339–341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farrugia L. J. J. Appl. Crystallogr. 1999, 32, 837–838. [Google Scholar]
- van der Sluis P.; Spek A. L. Acta Crystallogr., Sect. A 1990, 46, 194–201. [Google Scholar]
- Spek A. L. J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar]
- Amsterdam Density Functional (ADF) Modeling Suite; Scientific Computing & Modelling NV: Amsterdam, 2006.
- Vosko S. H.; Wilk L.; Nusair M. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar]
- Becke A. Phys. Rev. A 1988, 38, 3098–3100. [DOI] [PubMed] [Google Scholar]
- Perdew J. P. Phys. Rev. B 1986, 33, 8822–8824. [DOI] [PubMed] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas Ö.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09, revision A.02; Gaussian, Inc.: Wallingford, CT, 2009.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



