Róbert Gyepesb,
Jiří Pinkasa,
Ivana Císařovác,
Jiří Kubištaa,
Michal Horáčeka and
Karel Mach*a
aJ. Heyrovský Institute of Physical Chemistry, Academy of Sciences of the Czech Republic, v.v.i., Dolejškova 3, 182 23 Prague 8, Czech Republic. E-mail: karel.mach@jh-inst.cas.cz
bJ. Selye University, Department of Chemistry, Faculty of Education, Bratislavská cesta 3322, 945 01 Komárno, Slovak Republic
cDepartment of Inorganic Chemistry, Faculty of Science, Charles University in Prague, Hlavova 2030, 128 40 Prague 2, Czech Republic
First published on 26th September 2016
A thermally robust triple-decker complex [bis(η5-pentamethylcyclopentadienyltitanium)-μ-(η4:η4-1,2,4,5-tetrakis(trimethylsilyl)cyclohexa-1-4-diene-3,6-diyl)] (3) was obtained in 4% yield by thermolysing Cp*TiMe3 in the presence bis(trimethylsilyl)acetylene (BTMSA). The solid-state structure of centrosymmetric 3 features rather long all C–C bonds in the nearly planar bridging ligand (1.4720(14)–1.4896(15) Å) and a short distance of its least-square plane to the titanium atoms (1.7381(5) Å). Computational results revealed the bonding of the central ligand to be accomplished through back-bonding of its two CC bonds and through the simultaneous generation of two σ-Ti–C(H) bonds. Based on CASSCF and CASPT2 results, the molecule acquires several electronic configurations simultaneously, which hinders its representation by one single Lewis structure. Apart from being coordinated to the central ligand, the metal atoms are involved in a direct Ti–Ti bonding by the formation of one σ- and two δ-bonds between them. The bond order of this Ti–Ti overlap shows only a slight decrease upon electronic excitation. The presence of ionic contribution to the bonding of the central ligand is manifested by the charge −1.4e summed on the carbon atoms of the bridging ring. Based on computational results, the spin multiplicity of the ground state is singlet, while the first low lying excitation state is triplet. This is in agreement with the absence of EPR signal in either toluene solution and glass, and with slight downfield shifts of broadened 1H NMR signals of SiMe3 and Cp* methyl groups observed with increasing temperature.
In efforts to extend this chemistry area we payed attention to the eventual exploitation of pentamethylcyclopentadienyltitanium intermediates, which apparently take part in the thermolysis of pentamethylcyclopentadienyltitanium trimethyl [(η5-C5Me5)TiMe3] (1).4 The thermolysis of 1 at temperatures 96–120 °C was reported by Mena et al.5 to yield a titanium-methylidyne tetramer [{Cp*Ti(CH)}4] (2) based on the Ti4C4 cubane-like structure. Methane elimination during the reaction was proven to originate from titanium-methyl groups, hence the methylidene and methylidyne intermediates (1a–1c) depicted schematically in Scheme 1 were suggested to take part in the reaction pathway to compound 2.5
![]() | ||
Scheme 1 Thermolysis of 1 as suggested by Mena et al.5 |
The present work was initiated by our attempts to obtain new cyclopentadienyl-titanium derivatives by trapping the above intermediates with some internal acetylenes. These attempts resulted in the isolation of [bis(η5-pentamethylcyclopentadienyltitanium)-μ-(η4:η4-1,2,4,5-tetrakis(trimethylsilyl)cyclohexa-1-4-diene-3,6-diyl)] (3) only for the case of bis(trimethylsilyl)acetylene (BTMSA). It has to be noted that apart from [(CpV)2-μ-(η6:η6-benzene)] and [(CpV)2-μ-(η6:η6-mesitylene)],6 compound 3 is the only structurally-defined triple-decker complex in the first row of transition metals (Ti–Ni). The syntheses, molecular and electronic structure of 3 are the subject of the present communication.
Compound 3 was reproducibly obtained as green crystals in very low yield of 4% related to 1. The thermally robust compound melts at 280 °C (nitrogen atmosphere), even though undergoing simultaneous decomposition. The EI-MS spectra of 3 display the molecular ion m/z 732 as a base peak. Its fragmentation is rather poor, showing mainly the elimination of one [Cp*Ti] moiety (m/z 549) and the [Cp*Ti − 2H]+ (m/z 181) and [SiMe3]+ fragment ions. Infrared spectra of 3 in KBr pellet are dominated by very intense absorption bands of the SiMe3 substituents at 1247 cm−1 and 835 cm−1 leaving other absorption bands of the central bridging ligand and Cp* ligands of rather low intensity. The absorption band of aromatic C–H valence vibration reported for the free arene at 3076 cm−1 (ref. 8) was not observed, likely being shifted to lower wavenumbers into the methyl groups ν(C–H) region. 1H NMR spectra in toluene-d8 showed two broad resonances at 2.3 ppm (ν1/2 = 43 Hz) and 1.3 ppm (ν1/2 = 54 Hz) which were tentatively assigned to C5Me5 and SiMe3 groups on the basis of their integral intensities (10:
12; i.e. two Cp* rings versus four SiMe3 groups). Their chemical shifts were not far from the values for Cp* and SiMe3 groups in related diamagnetic complexes e.g., Cp*2Ti(η2-BTMSA) (C6D6): δH 1.718 ppm (Cp*) and 0.016 ppm (SiMe3),9 and [Cp*Ti(η2-BTMSA)(μ-Cl)]2 (C6D6): δH 2.078 ppm (Cp*) and 0.070 ppm (SiMe3).3a The signal of the SiMe3 groups is also significantly shifted downfield compared to that in 1,2,4,5-tetrakis(trimethylsilyl)benzene (δH 0.39 ppm)8 and the corresponding dianion in [Li(dme)]2[1,2,4,5-(SiMe3)4C6H2] (dme = dimethoxyethane) (δH 0.21 ppm).10 The 1H NMR resonances acquired at temperatures in the range 25–65 °C (see ESI†) showed a slight downfield shift with increasing temperature (0.04 ppm and 0.07 ppm per 20 °C, respectively), whereas their half-widths remained with no change. These observations imply a partial mixing of thermally populated triplet state with a ground singlet state resembling the behaviour of the green dimeric titanocene, [(μ-η5:η5-C5H4C5H4)(μ-H)2{Ti(η5-C5H5)}2].11 The two central ring hydrogen atoms were not detected at all, supposed to undergo a considerably larger broadening of their signal. The EPR spectra of the same toluene-d8 solutions exhibited no signals both at 22 °C as well as in glass at −160 °C as the d1–d1 triplet state mentioned above should relax rapidly. Essential information on the molecular structure of 3 was obtained from its single crystal X-ray diffraction study and its electronic structure was examined by computational methods (see below).
a Symmetry transformations used to generate equivalent atoms: −x + 2, −y, −z + 1.b Cg(1) is centroid and Pl(1) is least-squares plane of the C(1–5) cyclopentadienyl ring.c Cg(2) is the centroid and PL(2) is the least-squares plane of the C(11–13 and 11′–13′) bridging ring.d Dihedral angle between Pl(1) and PL(2).e Dihedral angle between Pl(2) and the Si(1), C(11), C(12), and Si(2) least-squares plane. | |||
---|---|---|---|
Bond lengths (Å) | |||
Ti–Cg(1)b | 2.1076(6) | Ti–Cg(2)c | 1.7381(5) |
Ti–PL(1)b | 2.1068(2) | Ti–PL(2)c | 1.7381(2) |
Ti–C(1–5) | 2.3946(11)–2.4500(11) | C–C(1–5) | 1.4136(18)–1.4240(17) |
Ti–C(13) | 2.2229(11) | Ti–C(13′) | 2.2621(11) |
Ti–C(11) | 2.2809(11) | Ti–C(12′) | 2.2851(10) |
Ti–C(12) | 2.3155(11) | Ti–C(11′) | 2.3184(10) |
C(11)–C(12) | 1.4896(15) | C(12)–C(13) | 1.4720(14) |
C(11)–C(13′) | 1.4735(15) | Si(1)–C(11) | 1.8889(11) |
Si(2)–C(12) | 1.8871(11) | Ti(1)–Ti(1′) | 3.4762(4) |
![]() |
|||
Bond angles (deg) | |||
Cg(1)–Ti–Cg(2)a,b | 176.85(2) | C(11)–C(12)–C(13) | 117.06(9) |
C(12)–C(11)–C(13′) | 117.12(9) | C(12)–C(13)–C(11′) | 125.52(9) |
![]() |
|||
Dihedral angles (deg) | |||
φd | 4.62(9) | ψ(1)e | 7.05(7) |
The overall molecule distortion is characterised by the angle Cg(1)–Ti–Cg(2) (Cg(1) is the centre of gravity of the C(1–5) cyclopentadienyl ring and Cg(2) is the centre of gravity of the C(11–13 and 11′–13′) ring) 176.85(2)° and the dihedral angle subtended by the cyclopentadienyl and the bridging ring least-squares planes amounting to 4.62(9)°.
The least-squares plane of Cp* (C1–C5) shows ring carbon atom deviations in the range −0.0076(7)–0.0092(7) Å, while all methyl carbon atoms (C6–C10) are displaced from this plane opposite to the neighbouring titanium atom. The displacement magnitude falls in the range 0.2629(22)–0.3427(21) Å except for C(10) which is displaced only by 0.0855(21) Å. This can be accounted for the absence of steric congestion between C(10) and the arene carbon atom C(13) bearing the sterically non-demanding C–H bond. The bridging ring (C11–C13, C11′–C13′) is nearly planar; its carbon atoms deviating from the least-squares plane by ±0.0242(6) Å for C(13) and ±0.0230(6) Å for C(11) and C(12). The Si(1) and Si(2) atoms are, however, displaced from this plane by −0.2713(20) Å and 0.2606(20) Å, respectively. The bridging ligand ring angle at the C(H) carbon atoms C(12)–C(13)–C(11′) exceeds the angle at carbon atoms C(11) and C(12) which bear the SiMe3 groups (125.52(9)° versus average 117.09(9)°). These angles are differing only slightly from those found for the free crystalline 1,2,4,5-tetrakis(trimethylsilyl)benzene,12 being thus apparently caused by steric demands of bulky SiMe3 substituents.13 In contrast to nearly unaffected bond angles, all bond lengths in the bridging ring (C–Cav. 1.478 Å) are substantially elongated (Table 2). Their lengths are markedly larger than in vanadium triple-decker complexes [(CpV)2-μ-(η6:η6-C6H6)] and [(CpV)2-μ-(η6:η6-1,3,5-C6H3Me3)] with C–Cav. 1.443(5) and 1.439(8) Å, respectively,6 and their elongation with respect to those in free arene12 by av. 0.072 Å throws doubts on eventual aromaticity of the bridging ring.
H | H˙− a | H2− b | 3 | |
---|---|---|---|---|
a The cation is [K([18]crown-6)(THF)2]+.b Cations [Li(1,2-dimethoxyethane)]+ are centred over opposite faces of 1,2,4,5-tetrakis(trimethylsilyl)benzenide. | ||||
Bond lengths | ||||
C11–C12 | 1.414(2) | 1.462(4) | 1.558(10), 1.549(11) | 1.4896(15) |
C12–C13 | 1.404(2) | 1.401(5), 1.405(5) | 1.392(11)–1.416(11) | 1.4720(14) |
Si–Cring | 1.870(3)–1.897(2) | 1.845(4)–1.854(4) | 1.807(8)–1.839(8) | 1.8871(11)–1.8889(11) |
![]() |
||||
Bond angles | ||||
C12–C11–C13′ | 117.1(2) | 115.5(3)–116.4(3) | 113.5(6)–115.3(6) | 117.12(9) |
C12–C13–C11′ | 125.9(2)–129.1(2) | 127.5(4), 128.3(4) | 128.9(7)–130.2(6) | 125.52(9) |
Comparison of geometry parameters of 3 with those of relevant compounds containing Ti–Cp* and Ti–arene bonds will be useful in further bonding considerations. Thus, the Ti–Cg(1) distance of 2.1076(6) Å is virtually identical with analogous Ti–Cg distances of 2.0995(9) Å and 2.0970(8) Å for [Cp*2Ti(η2-MeCCPh)] complex, where the titanocene (TiII) binds the alkyne by back-bonding and where the Cp* – alkyne steric hindrance is low.14 A shorter Ti–Cg distance of 2.042(5) Å was found for the half-sandwich (TiII) complex [Cp*Ti(η2-BTMSA)(μ-Cl)]2 (Chart 1, A), and 1.983(2) Å for the sandwich titanocene [Cp*2Ti] (co-crystal of [Cp*2Ti] with [Cp*2TiCl]) as a reference.15
Dibenzenetitanium [Ti0(η6-C6H6)2] (Chart 2, C) showed a virtually identical distance between its titanium atom and the coplanar benzene ring least-squares planes (1.736 Å)16 as compound 3. A group of (η6-arene)Ti(II)bis(tetrahaloaluminate) complexes afforded average Ti–Carene distances 2.49–2.50 Å for arene = benzene (Chart 2, D),17a mesitylene,17b durene17c or hexamethylbenzene,17d,e whereas the same parameter is distinctly shorter for 3 (av. 2.2808(11) Å). Relevant to the discussion of arene–titanium bonding in 3 is also the crystal structure of 1,2-cyclohexane-tethered bis(amidinate)titanium(II)(η6-toluene) (Chart 2, E) which contained the puckered toluene ring, having the two halves of its ring intersected with a dihedral angle of 20°. The ring geometry suggests the presence of cyclohexadiene dianion with enhanced sp3 hybridisation on intersecting carbon atoms and their σ-bonding to titanium(IV) as a resonance structure contributing to a Ti(II) centre back-bonding the arene ring.18
An example of a dititanium complex containing two σ-/π-bonded tripyrrole auxiliary ligands and a bridging π-faced toluene ligand (Chart 2, F) is also closely related to the present case. The two titanium moieties are different because one potassium cation [K(dimethoxyethane)2]+ is coordinated to one of the two auxiliary ligands, making the corresponding titanium atom formally Ti(I). DFT calculations indicated that the two titanium d-electrons are back-donated to two benzene π*-orbitals and the Ti(I) atom binds the toluene ligand more strongly than the Ti(II) one. A considerable difference in titanium–toluene binding strength is reflected in Ti–toluene centroid distances 1.758 and 1.830 Å.19 An even shorter Ti–Cg(2) distance of 1.7381(2) Å for 3 is to be understood on the basis of DFT calculations (see below).
A tentative clue to the nature of bonding in 3 could be drawn also from comparison of its crystallographic geometric parameters with those for 1,2,4,5-tetrakis(trimethylsilyl)benzene (H),12 its radical-anion (H˙−),20 and dianion (H2−)10 (Table 2).
A considerable elongation of all C–C bonds in 3 with respect to those of 1,2,4,5-tetrakis(trimethylsilyl)benzene (H) (cf. 1.4720(14)–1.4896(15) Å versus 1.403(2)–1.414(2) Å) (Table 2) is suggestive of the back-bonding mechanism resulting in elongation of π-coordinated unsaturated bonds.14 On the other side, these species do not differ in their ring angle and Si–Carene bond length values. The anionic H˙− and H2− species show remarkably enlarged differences in bond lengths between carbon atoms bearing the trimethylsilyl groups and the other ones, in larger differences in ring angles, and in a pronounced shortening of the arene carbon–Si atom bond lengths. All these indicate that ionic bonding should not dominate in 3, however, it has to be noted that ionic bonding of bridging arenes is quite common for f-elements. For instance, the lanthanocene and cerocene dinuclear benzene-bridged complexes [(Cpx2LnII)2(μ-η6:η6-C6H6)]− form ion pairs with a chelated potassium cation, whereas the negative charge of the lanthanide complex anion is concentrated on bridging benzene ligand – monoanion (C6H6)− for Cpx = C5H3(1,3-tBu2) and dianion (C6H6)2− for Cpx = C5H4SiMe3.21 A number of dilanthanocene22 and diuranocene23 – π-faced arene complexes were prepared and their crystal structures were used as probes for bonding models involving in addition to mono- and di-anionic arene species even π-faced tetraanionic arenes (6C, 10π aromatic system).24 The latter tetraanions should be characterised by the largest sum of C–C bond lengths in the tetraanionic arene ligand,25 which value is virtually equal to that for 3 (∑(C–C) = 8.8702 Å). In spite of the above surprising fit of ∑(C–C) parameters one can hardly accept that two equivalent, formally Ti(I) atoms could generate [Cp*Ti]2+ ions to bind to the central [arene]4− ion. Hence, computational methods have been employed to elucidate the nature of the Ti-π-faced bridging ligand bonding in 3.
The multiplicity of 3 was determined by comparing the total energies obtained for singlet, triplet and quintet multiplicity computed on the solid-state geometry. Since the lowest energy was obtained for the singlet multiplicity and since changes in multiplicity from singlet to quintet showed a steady increase in energy (Table 3), higher multiplicities were not considered. Using DFT results, the electronic properties and the bonding scheme of 3 were devised from the final model assuming the molecule a diradicaloid singlet.
Multiplicity | ΔE (kJ mol−1) |
---|---|
a Energies were obtained using the M06 functional and the 6-31+G(d,p) basis set on all atoms. | |
Singlet | 0.00 |
Triplet | 20.89 |
Quintet | 128.59 |
According to DFT results, the coordination of the Cp* ligands is accomplished in the usual manner, however, the coordination of central bridging ligand to the two titanium atoms poses some peculiarities. There are two π systems localised between C(SiMe3) atoms, and the coordination effected by their overlap with the metals bears analogy with the coordination of alkene ligands in mononuclear bent titanocene alkene/alkyne complexes. The bonding in the latter examples employs a π-donation to the metal and a d-electron back donation to the corresponding alkene/alkyne π*-orbital.14 The dinuclear 3 coordinates analogously through its two olefin moieties, even with one notable distinction. With mononuclear complexes, the π/π* orbital overlap is effected by ligand orbitals directed towards the empty space of the bent titanocene shell, while for dinuclear 3, the π/π* systems oriented perpendicular to the ligand plane are utilized. This is an appropriate orientation enabling both π/π* systems to concurrently overlap with both metals (Fig. 2).
In analogy with the case of mononuclear complexes, where the back donation to the alkyne ligand results in the carbon–carbon multiple bond elongation, the solid state structure of 3 displayed both its double bonds elongated to 1.4896(15) Å. The π/π* coordination of the central ligand is confined to the olefin carbon atoms, whereas its C–H carbon atoms utilize a σ-like overlap with the metals. Despite of both bonds being of completely different genesis, DFT results reported the canonical orbitals connected with the π/π* and the σ-coordination to be near-degenerate in energy. In addition, Mayer bond orders computed for the six Ti–C bonds were reported to be nearly equal (0.396 to the olefin atoms; 0.323 for the σ-coordinated ones), suggesting a remarkable similarity of the six carbon atoms. The sum of NBO charges on the six skeletal ligand atoms amounted to −1.41e that implies a remarkable contribution of ionic bonding in 3.
Combined with the planarity of the ligand, these results did not exclude the possibility of a benzene-like ligand embedded between two metal atoms. Should this be true, the ligand planarity would inherently come from its aromaticity. However, the aromatic character of a system having originally four π electrons remained still puzzling. On the other side the possibility of non-aromatic character of the ligand was suggested by the missing fully bonding combination of the six skeletal carbon atoms among the canonical orbitals, which contradicted the model of a benzene-like moiety. To resolve these inconclusive observations, the extent of aromaticity of the central ligand has been investigated by Nuclear Independent Chemical Shift (NICS).30
The first NICS computed exactly at the centroid of the central ring (denoted further as NICS(0)) yielded NICS(0) = −46.1517 ppm and NICS(0)zz, = −48.6315 ppm, which neither provided any directly useful information, since their absolute value exceeded the range expected for any aromatic or even antiaromatic organic system by a wide margin. Such low values indicated the presence of a diatropic ring current and as such would suggest significant ligand aromaticity. However, since NICS values and NICS(0) in particular have been questioned as unequivocal and reliable measure of ring aromaticity31 we have also computed the NICS(1) value (NICS taken 1.000 Å from the centroid and normal to the ring plane), which reached NICS(1) = −67.3144 ppm and NICS(1)zz = −73.5728 ppm. Neither these values were useful in assessing ligand aromaticity, and their exceedingly high magnitude was explained by the contribution of some local currents contributed from the metals to the devised NICS values.
In order to eliminate the contribution from its two neighbouring metal atoms, the NICS computation was repeated for the central ligand only, which was treated as dianion during this step. Considering the dianion of the central ligand only, both NICS values have fallen in the expected range: NICS(0) = 20.3732 ppm, NICS(0)zz = 12.0100 ppm and NICS(1) = 13.5098 ppm, NICS(1)zz = 9.1035 ppm. Based on these values, the central ligand is an antiaromatic system. Its multiple bonds thus remain localized and the planarity of this particular ligand in 3 must be the consequence of its interaction with the metals.
To find the reason of the near-degeneracy in energy of the two canonical orbitals incorporating the σ-coordination of the C–H carbons and the π-coordination of the acetylenic atoms and also to provide an explanation for the similar Ti–C bond orders found for all the six carbon atoms, the mixing between the respective canonical orbitals has been examined. Since the central ligand is non-aromatic on its own, the respective mixing was expected to be mediated by the metals. This was verified by NBO second-order Perturbation Theory, which has reported a CC bond delocalization occurring into the Ti–C σ-antibond with energy 51.30 kJ mol−1. This significant delocalization energy thus confirmed the anticipated mixing between the two orbitals and has also underlined the importance of the metallic contribution in its occurrence. This mixing is well discernible also among the canonical orbitals in the particular orbital that incorporates the Ti–C σ-overlap. The presence of such delocalization explains the preference for a planar ligand skeleton – the mixing is most efficient when the carbon atoms are lying in plane (Fig. 3); such an arrangement however cannot occur without the involvement of the metals, which provide their d-orbitals (mostly dx2−y2) for this purpose.
In search for the reason of high thermal stability of complex 3, neither the covalent nor the ionic interactions between the ligands and the metals provided a fully convincing explanation for its occurrence. Seeking for additional stabilizing contributions, the presence of a direct metal–metal interaction has been uncovered. This interaction is generated from the overlap between two originally nonbonding metallic d orbitals (Fig. 4), and although the separation between the respective centres disfavours a markedly beneficial overlap, the Ti–Ti Mayer bond order computed by DFT still acquired the value 0.333. This value suggested the importance of the direct Ti–Ti bond in the molecule, which could contribute to the overall stability of 3.
![]() | ||
Fig. 4 Alpha canonical orbitals no. 198 (HOMO; left) and 199 (LUMO; right) of 3. Both orbitals are drawn with 4% probability and are dominantly of d-character (96%). |
However, the presence of this particular metal–metal bond has raised suspicion that other d orbitals could also take part in its formation. Since these orbitals were for principal reasons treated as unoccupied by DFT, results obtained by DFT could possibly describe the metal–metal bonding underestimated in its magnitude. The probable involvement of other d orbitals was also supported by technical details during DFT studies, like the SCF procedure yielding notoriously unstable wavefunctions, and also by NBO, which reported a system apparently in excited state even on a stabilized ground-state wavefunction. These observations were interpreted as warning signs of treating a molecule that is mixing several electronic configurations simultaneously. To cope with such an electronic system, we have extended our computational studies by Complete Active Space SCF (CASSCF) followed by Second Order Perturbation Theory.
The active space in CASSCF was constructed from the canonical orbitals that incorporated a large contribution either from the metals, or from the central ligand, or from both. Any orbitals involved in coordinating the Cp* ligands only have been kept inactive. To account for possible additional low-lying excitation(s), computations were carried out as state-averaged CASSCF with three roots.
In agreement with DFT, CASSCF confirmed the coordination of the central ligand through concurrent ligand π-coordination and back donation to its π* orbitals and reported also the σ-coordination from the two C–H carbon atoms. However, it has uncovered some additional nuances in describing the metal–metal bond. Comparably with DFT, a direct Ti–Ti σ-interaction has been reported (cf. Fig. 4), however, state-averaged CASSCF yielded markedly non-integer occupation number (o.n.) not only for its bonding combination (o.n. 0.84) but also for its corresponding antibonding counterpart (o.n. 0.52). These results justified the necessity to apply a multiconfigurational approach in order to describe the Ti–Ti bond in 3.
To further increase the quality of results obtained by multiconfigurational computations, a CASPT2 (CASSCF followed by Second Order Perturbation Theory) computation has been carried out. Using the same CASSCF active space as described above, the successive CASPT2 approach yielded the first excitation energy 1.675 eV, in fairly good agreement with the experimental value 1.503 eV (λmax = 825 nm). Bond analysis has revealed, that in addition to the σ-coordination described above, two δ-type orbitals (with their antibonding counterparts) are also present (Fig. 5) in generating a direct bond between the metals.
Even though there are three metallic orbitals involved in the generation of the direct Ti–Ti bond, due to the large intermetallic separation (d(Ti–Ti) 3.4762(4) Å) in the molecule, the overlap between the δ-orbitals is impaired. The interaction between the metals can thus be considered as some intermediate between a regular chemical bond and antiferromagnetically coupled electrons separated to two centres.
The occupancy numbers of orbitals (Table 4) obtained by CASSCF suggested that all d-orbitals (and their combinations) are partially populated in ground state. Although upon excitation some major changes in occupancies of individual orbitals could be observed, the overall change in the Ti–Ti bond order remained still modest. This insignificant change of the Ti–Ti bond order was uncovered by Natural Bond Order analysis based on the LoProp computed density, yielding the values: root 1 = 1.864; root 2 = 1.719; root 3 = 1.710. On behalf of these results, the thermal stability of 3 can be explained by the presence of a direct Ti–Ti covalent bond, whose bond order decreased only slightly even upon excitation of the parent molecule.
CASSCF root | Orb. 99 (3dδg) | Orb. 100 (3dσg) | Orb. 101 (3dδg) | Orb. 265 (3dδu) | Orb. 266 (3dσu) | Orb. 267 (3dδu) |
---|---|---|---|---|---|---|
1 (ground state) | 0.23 | 1.05 | 0.21 | 0.09 | 0.95 | 0.08 |
2 | 0.25 | 0.73 | 0.89 | 0.27 | 0.30 | 0.09 |
3 | 0.92 | 0.74 | 0.25 | 0.10 | 0.29 | 0.23 |
By using the multiconfigurational methods, we have rechecked the correct multiplicity of 3. In accordance with DFT results, the lowest total energy was reported for the singlet ground state, since the CASPT2 root 1 energy was 109.33 kJ mol−1 higher for the analogous triplet state. However, the first excited triplet state energy was already falling below the corresponding value of the singlet state. Although not studied further in detail, the observed downfield shift of NMR signals of 3 (see above) can possibly be attributed to the presence of thermally populated triplet state molecules in solution. The intersystem crossing between the singlet and triplet states (and vice versa) is possibly achieved through spin–orbital coupling, supported by the presence of two metals in the complex. CASPT2 energies of the first three roots for both singlet and triplet states are given in ESI.†
The stirred mixture changed the initial red color rapidly to green and then to brown. After 2 h a dark brown solution was evaporated in vacuum, and a dark residue was repeatedly extracted with hexane. The concentrated extract was cooled to −28 °C precipitating brown crystalline solid. The latter was fractionally crystallised to give [C6H2-1,2,4,5-(SiMe3)4] (0.22 g, 60%) and brown compound [Cp*Ti(η2-BTMSA)(μ-Cl)]2 (Chart 1, A) (0.09 g, 11%). The both compounds were identified by 1H NMR and IR (KBr) spectra, the latter also by X-ray single crystal diffraction. The presence of 3 was not detected in either of the fractions.
CASSCF and CASPT2 computations were carried out by MOLCAS 8.0 (ref. 39) using the Dolg pseudopotential as built in MOLCAS for all atoms. The State-Averaged CASSCF employed three roots all with equal weights. The active space included a total of 13 orbitals and 10 electrons and was constructed from canonical orbitals incorporating a large contribution either from the metals, from the central ligand or from both. Orbitals involved in coordinating only the Cp* ligands were kept inactive. This selection of orbitals allowed for a smooth convergence in CASSCF with no large orbital rotations reported during convergence. Upon successful termination of CASSCF, orbitals in the active space were checked for possible rotations out from the active space. To account also for dynamic correlation energy, the CASSCF step was followed by Multi-State CASPT2 computation.
Attempts to perform geometry optimization prior to devising any properties from the wavefunction were also attempted. The DFT optimization has however removed the planarity of the central ring (although not too significantly), which was interpreted as the result of using only one electron configuration for the description of a multiconfigurational molecule. The attempt to optimize the geometry using CASSCF proved to be more successful and after computing the required gradients the program has reported a structure at stationary point with no requirement for further geometry optimization.
Footnote |
† Electronic supplementary information (ESI) available: CIF file for the structure 3. 1H NMR spectra of solution of 3 in toluene-d8 measured at variable temperatures, experimental data for 1,2,4,5-tetra(trimethylsilyl)benzene and crystallographic data and data collection, structure refinement details and parameters for 3. CCDC 1446324 (3). For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra14940e |
This journal is © The Royal Society of Chemistry 2016 |