Abstract

A high-spin, mononuclear TiII complex, [(TptBu,Me)TiCl] [TptBu,Me– = hydridotris(3-tert-butyl-5-methylpyrazol-1-yl)borate], confined to a tetrahedral ligand-field environment, has been prepared by reduction of the precursor [(TptBu,Me)TiCl2] with KC8. Complex [(TptBu,Me)TiCl] has a 3A2 ground state (assuming C3v symmetry based on structural studies), established via a combination of high-frequency and -field electron paramagnetic resonance (HFEPR) spectroscopy, solution and solid-state magnetic studies, Ti K-edge X-ray absorption spectroscopy (XAS), and both density functional theory and ab initio (complete-active-space self-consistent-field, CASSCF) calculations. The formally and physically defined TiII complex readily binds tetrahydrofuran (THF) to form the paramagnetic adduct [(TptBu,Me)TiCl(THF)], which is impervious to N2 binding. However, in the absence of THF, the TiII complex captures N2 to produce the diamagnetic complex [(TptBu,Me)TiCl]2(η1,η1;μ2-N2), with a linear Ti=N=N=Ti topology, established by single-crystal X-ray diffraction. The N2 complex was characterized using XAS as well as IR and Raman spectroscopies, thus establishing this complex to possess two TiIII centers covalently bridged by an N22– unit. A π acid such as CNAd (Ad = 1-adamantyl) coordinates to [(TptBu,Me)TiCl] without inducing spin pairing of the d electrons, thereby forming a unique high-spin and five-coordinate TiII complex, namely, [(TptBu,Me)TiCl(CNAd)]. The reducing power of the coordinatively unsaturated TiII-containing [(ΤptBu,Me)TiCl] species, quantified by electrochemistry, provides access to a family of mononuclear TiIV complexes of the type [(TptBu,Me)Ti=E(Cl)] (with E2– = NSiMe3, N2CPh2, O, and NH) by virtue of atom- or group-transfer reactions using various small molecules such as N3SiMe3, N2CPh2, N2O, and the bicyclic amine 2,3:5,6-dibenzo-7-azabicyclo[2.2.1]hepta-2,5-diene.
Short abstract
Tetrahedral TiII complexes with high-spin d2 electronic configurations have hitherto eluded isolation. We show that treatment of mer-[TiCl3(THF)3] with [Tl(TptBu,Me)] and KC8 affords such a complex: [(TptBu,Me)TiCl]. The electronic structure of [(TptBu,Me)TiCl] was explored by SQUID magnetometry, electrochemistry, X-ray absorption spectroscopy, and high-frequency and -field electron paramagnetic resonance as well as theoretical studies. Tetrahedral [(TptBu,Me)TiCl] activates various small molecules including N2, N2O, a strained amine, THF, and an isonitrile, benefiting from the low coordination number and reducing power of the TiII ion.
Introduction
Divalent Ti compounds have been widely used in C–C bond coupling reactions, such as the Kulinkovich or McMurry type reactions,1 and recently in catalytic C–N bond coupling reactions by Tonks.2 However, the coordination chemistry of isolable TiII is largely restricted to octahedral systems of the types [TiX2(L)] {X– = Cl, CH3, BH4, and OPh and L = 2 dmpe [1,2-bis(dimethylphosphino)ethane];3 X– = Cl and L = 2 TMEDA (tetramethylethylenediamine), 4 pyridines;4 X22– = porphyrin,5 L = 2 THF (tetrahydrofuran), 2 phosphine (-oxide)} and [TiΤp2] [Tp– = hydridotris(pyrazolyl)borate].6 Other strategies to stabilize the large and highly reducing TiII ion are to coordinatively saturate it with soft or π acids, such as N2,7 CO,8 isonitriles,9 bipyridine,10 1,10-phenanthroline,11 alkynes,12 olefins,13 cyclooctatetraene,14 phosphines,15 or bulky Cp–-based16 ligands. Common to these strategies is the reliance on maximizing the coordination number, and, notably, low-coordinate TiII complexes continue to elude isolation. In particular, we found that no example of a four-coordinate, high-spin TiII d2 complex exists. Despite the aforementioned cases all being formally categorized as TiII, the strongly π-accepting nature of one or more of the ligands means that the metal does not typically behave as a diradical or powerful π base. Perhaps only in the case of trans-[TiCl2(TMEDA)2]4a,4b could one argue that the TiII ion represents a bona fide example of a high-spin d2 configuration, given the more innocent nature of the chloride and chelating diamine ligands. However, this complex is highly unstable in solution, undergoing dinuclearization or disproportionation reactions.17 More recently, Lin, Tilley, Ye, and co-workers as well as Deng and co-workers have used sterically bulky, cyclic (alkyl)(amino)carbene (cAAC) ligands to stabilize C2-symmetric Ti18 and Hf19 centers of the type [(cAAC)2MCl2], wherein a combination of π-accepting properties of the ligand and/or low symmetry electronically favors closed-shell ground states.
Given the propensity for mononuclear TiII to disproportionate, especially if bearing halide or pseudohalide ligands, we decided to reinvestigate its chemistry with the ubiquitous20 Tp– (scorpionate) ligand. Lee and co-workers established that two Tp– ligands can be substituted onto TiII, with each being κ3 (vide supra).6 However, the use of more sterically encumbering groups on the pyrazoles should allow one to isolate and stabilize tetrahedral geometries with divalent ions, which has been observed for most of the 3d metals. In fact, Petrov and co-workers recently reported the synthesis of a tetrahedral divalent VII complex, [(TptBu2)VCl] [TptBu2– = hydridotris(3,5-di(tert-butyl)pyrazol-1-yl)borate21], via reduction of the VIII precursor [(TptBu2)VCl2].22 Inspired by the work of Lee, Deng, Lin, Tilley, Ye, and Petrov, we have now isolated and fully characterized the only example of a high-spin, tetrahedral complex of TiII, by using a relatively weak-field ligand, in this case TptBu,Me– [TptBu,Me– = hydridotris(3-tert-butyl-5-methylpyrazol-1-yl)borate]. If one excludes the hypothetical compound [TpRScCl], this Ti complex would essentially complete the divalent tetrahedral transition-metal 3d series [TpRMCl] [TpR– = a sterically demanding κ3-tris(pyrazolyl)borate; M = Ti, V,22 Cr,23 Mn,24 Fe,25 Co,26 Ni,27 Cu,28 and Zn29]. Herein, we provide a reliable synthetic route to a mononuclear, high-spin TiII ion confined to a tetrahedral geometry and provide complete spectroscopic and magnetic data for this unique system, including high-frequency and -field electron paramagnetic resonance (HFEPR) and X-ray absorption spectroscopy (XAS). This panel of characterization is complemented by theoretical, electrochemical, and reactivity studies. We demonstrate how the TiII ion readily binds THF (reversibly) as well as N2. In fact, coordination even of a strong σ base/π acid, such as CNAd (Ad = 1-adamantyl) to the TiII ion does not perturb the high-spin nature of the metal ion, despite forming a rare example of a five-coordinate isocyanide adduct, [(TptBu,Me)TiCl(CNAd)]. Last, we showcase how the TiII ion can be oxidized to TiIV to form platforms of the type [(TptBu,Me)Ti=E(Cl)] (E2– = O, NSiMe3, N2CPh2, and NH).
Results and Discussion
Synthesis of Mononuclear TiIII Complexes with a Sterically Encumbering Trispyrazolylborate Ligand
Lee and co-workers reported how [TiCl2(TMEDA)2] could be cleanly transmetalated with 2 equiv of [KTp] to afford the octahedral TiII complex [TiTp2]. Intuitively, one would expect this reaction to proceed via the monomeric intermediate [TpTiCl]. Surprisingly, we found that no examples of mono-Tp-based TiII complexes exist, despite the VII complex [(TptBu2)VCl] having been recently reported by Petrov and co-workers.22 An initial investigation of salt metathesis between mer-[TiCl3(THF)3] and [KTptBu2] led to isolation of the TiIII complex [(κ1,κ1,η2-TptBu2)TiCl2], albeit in low (∼10%) yield and with the presence of impurities (Scheme 1). Single-crystal X-ray diffraction (XRD; Figure 1) studies revealed a highly distorted system where the TptBu2– ligand coordinates in an unusual κ1,κ1,η2 mode, which has been observed only in coordination polymers30 and for large metal ions that favor high coordination numbers, such as BaII,31 SmIII,32 YbII,33 and UIII.34 Here, the severe skewing of one pyrazolyl arm results in an almost side-on interaction between both of its N atoms and the metal ion. A moderate degree of elongation for one of the B–N bonds [B1–N2, 1.597(3) Å] compared to the other two [B1–N4 and B1–N6, 1.562(3) and 1.542(2) Å, respectively] can be observed. Consequently, the N–N distances in each ring reflect a subtle degree of activation of the pyrazolyl groups, with the side-on-bonded one possessing a longer distance [1.403(2) Å] compared to the other two [1.382(2) and 1.385(2) Å]. Distortion of one of the sterically encumbered pyrazolyl arms in [(κ1,κ1,η2-TptBu2)TiCl2] implies the possibility of decomposition reactions, involving B–N and N–N bond activation, especially in subsequent reduction chemistry, leading to the large and strongly reducing TiII ion. Not surprisingly, attempts to prepare pure samples of [(κ1,κ1,η2-TptBu2)TiCl2] invariably resulted in impure bulk material containing traces of other species, including free pyrazole. Given the larger ionic radius of TiII (100 pm) versus VII (93 pm), we decided to reduce some of the steric congestion in TptBu2– by replacing the tert-butyl group in the 5 position of each pyrazolyl arm with a methyl group. Theopold and co-workers have popularized this ligand scaffold with various 3d metals, including the early-transition-metal Cr in the formal oxidation states I–V.23,35
Scheme 1. Synthesis of [(κ1,κ1,η2-TptBu2)TiCl2].
Figure 1.

Two views of the solid-state structure of [(κ1,κ1,η2-TptBu2)TiCl2], showing distortion of one pyrazolyl arm of the TptBu2– ligand. Thermal ellipsoids are at 50% probability. Solvent and H atoms (except on B) are omitted.
Following a modified recipe to prepare [(κ1,κ1,η2-TptBu2)TiCl2], treatment of mer-[TiCl3(THF)3] with 1 equiv of [Tl(TptBu,Me)] in toluene over 13 h at 70 °C resulted in the smooth formation of [(TptBu,Me)TiCl2] (1), which was isolated in 72% yield as blue crystals (Scheme 2). The use of [Tl(TptBu,Me)] as opposed to the K+ salt improves the overall isolated yield because of the lower content of residual pyrazole that remains from synthesis of the alkali salt. X- and Q-band EPR spectroscopic studies of 1 in toluene glass afford typical S = 1/2 rhombic spectra, indicative of low symmetry around the TiIII ion (X band, g = [1.973, 1.923, 1.848], 12 K; Q-band, g = [1.973, 1.924, 1.848], 2 K; Figure 2). Likewise, solution Evans (μeff = 1.70 μB, 27 °C, C6D6), IR (νBH = 2569 cm–1), and UV–vis (d–d, 648 nm; ε = 35 M–1 cm–1) spectral studies are all in accordance with a monomeric d1 species. While the starting material, mer-[TiCl3(THF)3], shows a UV–vis absorption spectrum with two transitions due to a Jahn–Teller-distorted excited Eg state (668 and 763 nm, ε = 7 and 4 M–1 cm–1, respectively), complex 1 shows a single transition at 648 nm with a 5-fold increase in intensity in accord with the lower symmetry about the TiIII ion (Figure S51). XRD studies also confirmed a five-coordinate monomer more in line with a square-pyramidal geometry at the TiIII ion (τ5 = 0.27;36Figure 3). Moreover, no isomeric species resulting from borotropic shifts37 have been observed for 1, and significant twisting of the pyrazolyl arms is no longer as noticeable when judged by the B–N and N–N metrical parameters. The view down the Ti–B–H axis (Figure 3, right) clearly shows how twisting of the pyrazole arms in 1 is minimized compared to [(κ1,κ1,η2-TptBu2)TiCl2].
Scheme 2. Synthesis of Precursor 2via Reduction of 1, Reactivity Studies Involving THF, N2, CNAd, and Various Oxidants (N2CPh2, N3SiMe3, and N2O), as Well as the Bicyclic Amine Hdbabh to Form the Parent Imide 7.
Figure 2.
X-band (derivative, left) and Q-band (absorption, right) EPR spectra of 1 (black curves, experimental; red curves, simulated). Key experimental (black text) and S = 1/2 powder pattern simulation (red text) parameters are given (W = Gaussian line widths, hwhm). The Q-band spectrum exhibits an absorption, rather than first derivative, line shape due to rapid passage effects.38
Figure 3.

Solid-state structure of 1 from two different views. Thermal ellipsoids are at 50% probability. H atoms (except on B) are omitted.
Synthesis of a High-Spin, Tetrahedral TiII Complex and Exploration of Its Electronic Structure
A cyclic voltammogram of 1 revealed an irreversible one-electron anodic process with an anodic peak at 0.44 V (referenced to FeCp20/+ as 0.0 V), while a reversible one-electron cathodic process could be observed at −1.85 V (Figure 4). In Lee’s [Tp2Ti]PF6 complex, reversible one-electron oxidation and reduction processes occur at 1.20 and −1.28 V versus SCE, i.e., at 0.65 and −1.83 V versus FeCp20/+.39 These features indicate that reduction potentials are largely conserved between five-coordinate 1 and pseudooctahedral [Tp2Ti]PF6, and, importantly, both TiIII complexes provide access to the corresponding TiII ions in a reversible manner. In the case of 1, the nonreversible electrochemical features are likely due to significant structural distortion upon reduction, whereas the irreversible anodic feature might be due to the instability of a hypothetical four-coordinate [1]+ species, especially in the presence of electrolyte. Encouraged by the facile electrochemical reduction chemistry, we conducted a chemical reduction of 1 under Ar with KC8 in THF over 10 min, which resulted in an immediate color change from blue to intense green. Subsequent workup of the reaction mixture resulted in the formation of [(TptBu,Me)TiCl] (2) as light-green crystals in a respectable yield (78%; νBH = 2554 cm–1; Scheme 2). XRD studies revealed a slightly distorted tetrahedral TiII ion in each of the two crystallographically independent molecules (Table 1 and Figure 5, top). The chlorido ligand and Ti–B vector in the conformers of 2 range from being aligned in a nearly linear fashion (2-lin; τ4 = 0.75 and τδ = 0.7340) to being significantly bent (2-ben; τ4 = 0.69 and τδ = 0.57) with Cl–Ti–B = 174.6(1)° and 162.9(1)°, respectively. This tilt of the chlorido ligand has notable consequences on the solid-state electronic structure of 2 (Supporting Information, section 13). Figure 5 (bottom) shows the 3-fold axis along the Cl–Ti–B vector. As expected for the larger TiII ion, the average Ti–Cl distances in 2 exceed those in 1 (2.386 vs 2.307 Å), whereas the Ti–N distances remain practically unchanged (2.169 vs 2.178 Å). In fact, the Ti–Cl bonds in 2 fall within the longest 7% of crystallographically characterized Ti–Cl distances in the CSD (version 2020.1), reflecting the large size of the TiII ion in 2. Solution (Evans, μeff = 2.72 μB, 27 °C, C6D6) and direct-current (dc) SQUID (μeff = 2.61 μB, 27 °C; Table 2) magnetization measurements of 2 are consistent with a high-spin TiII d2 ion. Specifically, the SQUID magnetization measurements of 2 in the temperature range 2–300 K are consistent with an S = 1 species (for two independently isolated samples). Below 10 K, the magnetic moment drastically decreases to 2.24 μB (at 2 K), as shown in Figure 6, due to zero-field splitting (ZFS, vide infra). Variable-temperature and -field (VTVF) magnetization measurements between 1 and 5 T were also measured and simulated using a gavg value of 1.85 (Figure 6, inset, and Table 2). Being confined to a tetrahedral coordination geometry, complex 2 shows ligand-field transitions at lower energies than those in 1, extending into the near-IR (NIR) region of the UV–vis spectrum (λ = 391, 701, 857, 889, and 925 nm; ε = 1500, 180, 340, 320, and 230 M–1 cm–1, respectively; Figure S52).
Figure 4.
Scan-rate dependence for cyclic voltammograms of 1 under N2 in 0.1 M [nBu4N][PF6] in THF.
Table 1. Geometric Parameters for the Two Crystallographically Independent Molecules of Complex 2 (Pca21).
| geometry | τ4/τδ | B–Ti–Cl(deg) | Ti–Cl (Å) | Ti–Navg (Å) |
|---|---|---|---|---|
| 2-lin | 0.75/0.73 | 174.6(1) | 2.375(2) | 2.159 |
| 2-ben | 0.69/0.57 | 162.9(1) | 2.397(2) | 2.180 |
Figure 5.

Solid-state structures of the two crystallographically independent molecules in 2 (left, 2-lin; right, 2-ben). Thermal ellipsoids are at 50% probability. H atoms (except on B) are omitted.
Table 2. Electronic Structure Parameters for Complex 2 Extracted with SQUID Magnetometry and HFEPR Spectroscopy.
| SQUID | |||
|---|---|---|---|
| D (cm–1) | E/D | gavg | |
| dc, 1 T | –5.08 | 0 | 1.85 |
| VTVF | –4.86 | 0 | 1.85 |
| HFEPR | ||||||
|---|---|---|---|---|---|---|
| species | D (cm–1) | E (cm–1) | E/D | gx | gy | gz |
| 2-pos | +1.55(1) | +0.38(1) | 0.25 | 1.87(1) | 1.89(1) | 1.92(2) |
| 2-neg | –5.91(2) | –0.31(2) | 0.05 | 1.91(2) | 1.94(2) | 1.7(1) |
Figure 6.

Solid-state dc and VTVF SQUID (inset) magnetization studies of 2 at 1–5 T from 2 to 300 K.
Given the unique geometry of 2 and the even number of unpaired electrons resulting in an integer electronic spin, we carried out HFEPR spectroscopic measurements of powdered samples of 2 at low temperatures (Figure 7, top, 216 GHz; spectra at additional frequencies are shown in Figures S59–S62). The spectra are characteristic powder patterns for a triplet spin state (S = 1) and reveal the presence of two independent S = 1 species. The number and intensity of turning points in the spectra could be readily simulated at various frequencies using two sets of spin-Hamiltonian parameters (Table 2). These parameters were obtained from 2D maps of field versus frequency (Figure 7, bottom), with a tunable-frequency methodology applied.41 The sign of D was obtained from the simulation of single-frequency spectra such as those shown in Figure 7 and was found to be negative for one species, hence designated as 2-neg and positive for the other, hence 2-pos (E was assigned the same sign as that of D by convention). The spectroscopically determined D values compare well with the average value of the same parameter extracted from SQUID magnetometry (−5.08 cm–1; Table 2), and the presence of two crystallographically independent molecules in the asymmetric unit of the crystal structure of 2 readily explains the two S = 1 species observed by HFEPR. Apparently, the sign and magnitude of D is highly sensitive to the geometry of the TiII ion in 2. Ligand-field-theory (LFT) analysis using the angular overlap model provides an explanation for 2-neg and 2-pos having different magnitudes and signs of ZFS (Supporting Information, section 13). Conformer 2-lin, which more closely approximates trigonal (C3v) symmetry, is assigned to 2-pos, whereas the more distorted conformer 2-ben is assigned to 2-neg.
Figure 7.
Top: Solid-state HFEPR spectrum of 2 at 4.5 K and 216 GHz (black trace). The colored traces are simulations using spin-Hamiltonian parameters (Table 2) of 2-neg (bottom) and 2-pos (top); red traces are simulations using positive D; blue traces are negative D values. Bottom: 2D plot of turning points in the powder spectra of 2 as a function of the frequency, marked as squares: full squares, 2-neg; empty squares, 2-pos. The lines are simulated with red color representing the turning points with the external magnetic field, B, parallel to the x axis of the zfs tensor, blue: B || y, and black: B || z.
The dependence of D on molecular structure was further explored by quantum-chemistry computational means. The g and D tensors were calculated for 2 by employing hybrid density functional theory (DFT) calculations using the B3LYP hybrid density functional with the CP(PPP) basis set on the Ti atom and the ZORA-def2-TZVP(-f) basis set on all other atoms; the calculations were based on the atomic coordinates from the crystal structures. The two conformers of 2 (cf. Table 1) each gave a unique set of parameters, in accord with the appearance of two spin systems in the HFEPR data. Notably, as with classical LFT (Supporting Information, section 13), the calculations reproduce the oppositely signed D values, although the magnitudes of D and E/D poorly match the experimental data (Table 3), which was also the case using LFT. In light of this, complete-active-space self-consistent-field (CASSCF) calculations with the strongly contracted N-electron valence perturbation theory 2 (NEVPT242) dynamical correlation correction were undertaken. Multiconfigurational approaches, including CASSCF/NEVPT2, predict ZFS for transition-metal complexes more accurately than DFT.43 This is partially attributable to multiconfigurational methods giving a more accurate prediction of d–d excited-state energies, which are essential to accurately determining the spin–orbit coupling (SOC) contribution to the ZFS. The present calculations used the effective Hamiltonian theory and included both contributions from SOC and spin–spin coupling (SSC);44 the SSC contribution is not directly accounted for in the classical LFT model.
Table 3. Calculated Spin-Hamiltonian Parameters for Conformers of Complex 2.
| geometry | τ4 | method | g1 | g2 | g3 | D (cm–1) | E/D |
|---|---|---|---|---|---|---|---|
| 2-lin/2-pos | 0.75 | DFT | 1.966 | 1.971 | 1.973 | –5.58 | 0.029 |
| CASSCF/NEVPT2a | 1.792 | 1.829 | 1.916 | +2.87 | 0.19 | ||
| 2-ben/2-neg | 0.69 | DFT | 1.953 | 1.970 | 1.979 | +7.59 | 0.020 |
| CASSCF/NEVPT2a | 1.668 | 1.888 | 1.940 | –7.77 | 0.068 |
Calculated using CAS(2,5) with an effective Hamiltonian SOC contribution and including a SSC contribution. CASSCF calculations are state-averaged over all d–d states of the d2 configuration.
The CASSCF/NEVPT2 calculations were based on the quasi-restricted orbitals (QROs43d) generated by the B3LYP calculation. The active spaces comprised two electrons and five orbitals [CAS(5,2)] and were averaged over all singlet (15) and triplet (10) ligand-field states expected for a d2 ion in C3v symmetry. State energies and ground-state configurations are included in the Supporting Information. The spin-Hamiltonian parameters (g, D, and E/D) calculated using this approach conform better with experiment than values calculated using DFT (Table 3). Importantly, species 2-neg observed by HFEPR can be plausibly assigned to conformer 2-ben, while 2-pos can be assigned to conformer 2-lin, exactly as proposed from LFT.
Reactivity of Complex 2 toward N2, THF, and an Isocyanide
Transient TiII species are known to activate N2, but reports of isolable and well-defined TiII species that can activate this inert gas remain quite rare.8c,45 Potential synthetic applications of this activation chemistry include reductive splitting of N2.46 In order to prepare complex 2, manipulations must be carried out under a strictly pure Ar atmosphere, and solvents must be sparged thoroughly in order to avoid reactivity with residual N2. Indeed, we found that reduction of 1 with KC8 in THF under a N2 (as opposed to Ar) atmosphere, and with workup carried out essentially identically with that for 2, resulted in the clean formation of a new diamagnetic product. Moreover, exposure of 2 in a pentane solution to an atmosphere of N2 afforded the same species, quantitatively, over the course of minutes as a dark-green crystalline material. We should stress that, even in an Ar-filled glovebox, complex 2 gradually converts, in solution, to this new diamagnetic species, making it a potent scavenger of N2! By contrast, crystalline 2 shows no signs of conversion into this new species when stored under N2 over at least 24 h, reflecting the energetic penalties associated with structural changes in the solid state. The 1H and 13C NMR, in addition to IR (νBH = 2559 cm–1), spectral data of this new species feature a single pyrazolyl chemical environment, indicating that the new five-coordinate species undergoes rapid Berry pseudorotations in solution. XRD studies established formation of the bridged N2 species [(TptBu,Me)TiCl]2(η1,η1;μ2-N2) (3) having a linear Ti=N=N=Ti topology (Scheme 2 and Figure 8). The elongated N–N bond of 1.207(4) Å with respect to free N2 (1.0977 Å),47 the short Ti–N distances [1.805(3) and 1.812(3) Å], and the diamagnetic nature hints to this motif being somewhere in the N22– range. The two Cl ligands are oriented with dihedral angles of 85.43(7)–115.98(5)° (one Cl is disordered), consistent with complex 3 adopting both gauche and eclipsed conformations in the solid state. Dark-green 3 absorbs strongly across the entire UV–vis region with a maximum at 652 nm (ε = 580 M–1 cm–1; Figure S55), reflecting strong interaction between the bridging N2 ligand and Ti centers.
Figure 8.

Solid-state structure of 3. Thermal ellipsoids are at 50% probability. Solvent and H atoms (except on B) are omitted. Disorder in two tBu groups and Cl1′ is not shown.
Returning to the TiIII/TiII redox couple from the electrochemical studies of 1, it would be interesting to establish whether the identity of the reduced species might correspond to 2 or 3. During the synthesis of 2 and 3, we noted that the reduction of 1 with KC8 in THF, in both cases, afforded an intense-green solution, the color of which matches neither the light-green color of 2 nor the dark-green color of 3. In both syntheses, removal of the volatiles produces an intense-green film, and the addition of pentane induces crystallization of 2 (under Ar) and 3 (under N2) over the course of minutes. Surprisingly, an initial attempt at generating the 15N isotopologue of 3 by exposing 2 to 15N2 in THF over 1 h resulted in the isolation of only unconverted 2. Likewise, a complementary experiment involving the reduction of 1 with KC8 in THF under N2, followed by removal of the volatiles in vacuo and subsequent workup under Ar, afforded a mixture of 2 and 3 in a 70:30% proportion. To better understand the apparent inhibition of N2 binding to 2 in THF, we turned to UV–vis spectroscopic studies. Upon going from toluene to a THF solution, absorption from 2 at 391 nm (1500 M–1 cm–1) undergoes a slight redshift to 412 nm (380 M–1 cm–1). More importantly, the intense absorptions extending into the NIR region (λ = 701, 857, 889, and 925 nm and ε = 180, 340, 320, and 230 M–1 cm–1, respectively) disappear altogether and are replaced by a weaker and broad absorption at 606 nm (38 M–1 cm–1; Figure S53). Thus, titration of toluene solutions of 2 with THF (0–0.70 M) indicates that an equilibrium takes place between 2 and the Lewis base with an association constant in the range of 5–8 M–1 (Figure 9, left). In neat THF (12.3 M), this corresponds to 98–99% of 2 being present as the five-coordinate [(TptBu,Me)TiCl(THF)] (2-THF; Scheme 2 and Figure 9, right). Eventually, we isolated crystalline 2-THF (in the strict absence of N2) by prolonged storage of 2 in pentane containing low concentrations of THF (−35 °C). XRD studies verify that 2-THF is a rare example of a five-coordinate, mononuclear, and high-spin TiII ion (τ5 = 0.46). Moreover, the Ti–Cl [2.414(1) Å] and Ti–OTHF [2.251(2) Å] distances in 2-THF fall in the longest decile of crystallographically characterized Ti–Cl and Ti–OTHF distances in the CSD (version 2020.1), again reflecting the size of the TiII ion in 2-THF. Importantly, the reactivity and spectroscopic and structural data provide evidence that the reduction of 1 to ultimately produce 2 or 3 proceeds with 2-THF as a common intermediate (Scheme 2). It is apparent from our studies that THF blocks the binding site for N2 activation. This finding is quite influential because it suggests that reductions of metal ions, often conducted in polar solvents such as THF, might actually disfavor binding of N2 if the association constant of THF is large. In addition to this feature, the lability of the THF ligand is central for the TiII ion in subsequently attaining the tetrahedral geometry that is characteristic of complex 2.
Figure 9.

Left: UV–vis absorption spectra of 2 (1.6 mM) in toluene with increasing concentrations of THF. Right: Solid-state structure of 2-THF. Thermal ellipsoids are at 50% probability. Solvent and H atoms (except on B) are omitted.
Exposing complex 2 to isotopically enriched 15N2 in pentane afforded [(TptBu,Me)TiCl]2(η1,η1;μ2-15N2) (3-15N), and collecting 15N NMR spectroscopic data revealed a downfield chemical shift at 309 ppm [vs NH3(l) referenced at 0.0 ppm; top right in Figure 10]. The 15N NMR data indicate a reduced N2 ligand; known [{LnTi}2(η1,η1;μ2-15N2)] complexes display 15N resonances in the range of 322–558 ppm versus NH3(l).7h,7i,7k,7l A comparison of the IR spectra of 3-15N and 3 allows the identification of low-energy Ti–N stretching frequencies (14N/15N at 769/749 cm–1; Figure 10, top left), which gives a ratio (1.027) in good agreement with that calculated within the harmonic oscillator approximation for a 48Ti–14N/48Ti–15N isotopic system (calcd 1.0269). Resonance Raman spectroscopy obtained by excitation at 405 nm reveals a strongly red-shifted N–N vibration (14N/15N at 1401.1/1346.3 cm–1; Figure 10, bottom), indicating a highly reduced N2 unit consistent with a N22– ligand (informative N–N Raman stretching frequencies;48 N2H4, 1076 cm–1; trans-N2H2, 1529 cm–1). With these data, 3 can be formally construed as two TiIII ions bridged by a N22– ligand.
Figure 10.
Top left: IR spectral data for 3 (black) and 3-15N (red) along with a difference spectrum (blue). Top right: 15N NMR spectrum of 3-15N [with an external standard, AdC15N, referenced at 244 ppm vs NH3(l) referenced at 0.0 ppm49]. Bottom: Resonance Raman (405 nm excitation) data for 3 (black) and 3-15N (red) along with a difference spectrum (blue).
We also investigated the reactivity of 2 with a π acid, such as the isocyanide CNAd, to probe whether this reaction would yield a low-spin complex. Accordingly, treatment of 2 with CNAd resulted in clean formation of the adduct [(TptBu,Me)TiCl(CNAd)] (2-CNAd) in 66% yield as a dark-maroon crystalline material. The 1H NMR spectrum of 2-CNAd covers a ∼50 ppm chemical shift range (−8 to +40 ppm), similar to the range found for 2 (−2 to +46 ppm). Likewise, the magnetic moment extracted with Evans’ method (μeff = 2.72 μB, 28 °C, C6D6) attests to 2-CNAd still being high spin, consistent with the persistence of a TiII ion. Akin to 2-THF, an XRD study reaffirmed the connectivity in 2-CNAd, showing a rare example of a five-coordinate high-spin TiII complex (Figure 11). The linear Ti–C–N geometry [174.2(1)°] in 2-CNAd and the modest redshift of the C≡N stretching frequency versus free CNAd (2096 vs 2124 cm–1; Figure 11) reflect a moderate degree of π-back-bonding, in line with the high-spin nature of 2-CNAd. However, the isonitrile ligand is not entirely spectroscopically innocent because the UV–vis spectrum of 2-CNAd displays much more intense charge-transfer bands (λ < 450 nm and ε > 3000 M–1 cm–1; Figure S56) than complexes 2 and 2-THF.
Figure 11.

Left: Solid-state structure of 2-CNAd. Thermal ellipsoids are at 50% probability. Solvent and H atoms (except on B) are omitted. Right: IR spectral data of 2-CNAd (black) and free CNAd (gray).
XAS Spectra of Complexes 1–3 and 2-CNAd
Ti K-edge XAS spectra were obtained for compounds 1–3 and the titanium(II) isocyanide adduct, 2-CNAd (Scheme 2 and Figure 12). The rising-edge energies obtained from the first derivatives of the spectra are presented in Table 4. Compounds 1 and 3 exhibit inflection points at 4974.5 and 4973.9 eV, respectively, while 2 and 2-CNAd have inflection points at 4971.6 and 4971.1 eV, respectively. The ca. 3 eV gap between the pairs of rising-edge energies accords with the assignment of 2 and 2-CNAd as bona fide TiII species, whereas 1 and 3 are physically defined TiIII species. The XAS spectra also display prominent preedge absorption features, which were assigned with aid from time-dependent density functional theory (TDDFT) calculations (vide infra).
Figure 12.
Ti K-edge XAS data for complexes 1 (blue), 2 (black), 2-CNAd (gray), and 3 (red).
Table 4. Ti K-Edge XAS Preedge and Rising-Edge Energies for 1–3 and 2-CNAd.
| complex | preedge energy (eV) | rising-edge energy (eV) |
|---|---|---|
| 1 | 4968.0 | 4974.5 |
| 2 | 4966.8 | 4971.6 |
| 2-CNAd | 4966.6 | 4971.1 |
| 3 | 4967.3 | 4973.9 |
Electronic Structure Calculations
Hybrid DFT calculations using the B3LYP hybrid density functional with the CP(PPP) basis set on the Ti atom and the ZORA-def2-TZVP(-f) basis set on all other atoms were carried out using atomic coordinates from the crystal structures. The calculations were used as starting points for TDDFT calculations of the Ti K-edge XAS data. Calculated excitation energies correlated strongly and linearly with the experimental data (Figure S66; R2 = 0.97). The linear fit was used to shift the calculated energies such that the calculated and experimental spectra overlap (Figures 13 and 14). Spectra for 2 were calculated for both independent conformers in the crystal structure (2-lin and 2-ben; Table 1) but were found to be effectively superimposable.
Figure 13.

Overlap of the experimental and TDDFT-calculated [B3LYP and CP(PPP) on the Ti atom and ZORA-def2-TZVP(-f) on all other atoms] Ti K-edge XAS spectra of 2 (geometry for conformer 2-lin) with acceptor MOs for preedge transitions plotted at an isovalue of 0.03 au. Calculated data have been energy-shifted following the correlation given in Figure S66.
Figure 14.

Overlap of the experimental and TDDFT-calculated [B3LYP and CP(PPP) on the Ti atom and ZORA-def2-TZVP(-f) on all other atoms] Ti K-edge XAS spectra of 3 with acceptor MOs for preedge transitions plotted at an isovalue of 0.03 au. Calculated data have been energy-shifted following the correlation given in Figure S66.
Assignments of the preedge features in the Ti K-edge XAS data of 2 and 3 were facilitated using molecular orbital (MO) diagrams produced by the initial single-point DFT calculations. Complex 2 exhibits a 2 + 1 + 2 d–d splitting diagram, as would be expected for an effectively C3v system (Figure 15). The two d electrons reside within an effectively degenerate (e) level comprised of MOs of ca. 80% d character in accordance with the physical TiII d2 assignment suggested by the XAS data. Preedge excitations thus comprise several quadrupole-allowed Ti 1s → 3d excitations that gain intensity due to 4p mixing in the C3v environment.
Figure 15.

Truncated MO diagram depicting the ligand field for 2 (geometry for conformer 2-lin) using QROs generated following a spin-unrestricted (UKS) DFT calculation using the B3LYP hybrid density functional with the CP(PPP) basis set on the Ti atom and the ZORA-def2-TZVP(-f) basis set on all other atoms. Orbitals are plotted at an isovalue of 0.03 au.
The MO diagram of 3 (Figure 16) resulting from a spin-restricted hybrid DFT calculation depicts four electrons in two pseudodegenerate MOs of ∼50% Ti 3d and ∼30% N 2p parentages. An alternative electronic structure description for 3 involves the antiferromagnetic coupling of two TiIII ions through the N2 bridge. To explore these bonding scenarios, we carried out broken-symmetry (BS1,1) calculations on 3. However, unrestricted corresponding orbital analysis reveals that no pairs of α and β spin orbitals have low overlap (S < 0.75). Two pairs of spin orbitals have intermediate overlap (S ∼ 0.8); these are π-bonding combinations between two Ti 3d and two N 2p orbitals, which are effectively delocalized over the entire (Ti2N2)4+ fragment. Moreover, the calculated coupling constant50 is J ∼ −5000 cm–1, an excessively large magnitude to be considered magnetic exchange. Together, these values suggest that considering 3 as a pair of TiIII centers antiferromagnetically coupled through the bridging N2 fragment is inappropriate. Instead, the crystallographic, spectroscopic, and computational data for 3 are in line with a TiIII–N22––TiIII core having two covalent bonds between N2 and two Ti.
Figure 16.

Truncated MO diagram for 3 generated with a spin-restricted hybrid DFT calculation employing the B3LYP hybrid density functional, the CP(PPP) basis set on the Ti atom, and the ZORA-def2-TZVP(-f) basis set on all other atoms. Orbitals in black have large contributions from Ti 3d, while orbitals in gray have small (<10%) contributions from Ti 3d. Orbitals are plotted at an isovalue of 0.03 au.
Reactivity of Complexes 2 and 3 toward Small Molecules
Complex 2 is stable in the solid state at room temperature and can be stored over months without showing signs of decomposition. In a toluene or benzene solution, 2 is stable over several hours at room temperature, after which it degrades to a myriad of unrecognizable diamagnetic products (full degradation after 20 h). By contrast, 2 is stable for at least 24 h in a THF solution at room temperature (by forming 2-THF), reflecting the increase in stability achieved upon attainment of a five-coordinate geometry (vide supra). Complex 2 displays rich chemistry; for example, it rapidly reacts with N3SiMe3 to form the imide [(TptBu,Me)Ti≡NSiMe3(Cl)] (4) in 80% isolated yield along with N2 extrusion (Scheme 2). Complex 4 is diamagnetic and has been characterized by a combination of IR and 1H and 13C NMR spectra in addition to solid-state structural studies (Figure 17, top left). 29Si NMR spectroscopy reveals an upfield resonance (−13.5 ppm, vs SiMe4 referenced at 0.0 ppm), indicating that the “(TptBu,Me)Ti≡N–(Cl)” fragment in 4 is more electron-donating than a methyl group. The structure shows short Ti–N distances for the imide ligand [1.692(1)–1.700(2) Å] in accordance with metal–ligand multiple-bond character, whereas the Ti–N–Si angles [158.94(10)–160.24(9)°] indicate that the imide N is best described as an sp-hybridized motif. In contrast, when 2 is treated with N2CPh2, N2 is not expelled, and instead of the hypothetical carbene “[(TptBu,Me)Ti=CPh2(Cl)]”, the diamagnetic diazoalkane adduct [(TptBu,Me)Ti≡NNCPh2(Cl)] (5) is isolated in 91% yield. Neither thermolysis (100 °C, overnight) nor photolysis (full spectrum from a Xe lamp, 1 h) promotes N2 extrusion from 5. Thus, 5 behaves like Herberhold’s prototypical [Cp2Ti(N2CPh2)(PMe3)] complex51 but shows marked contrast to the more reactive [Cp2Ti(diazoalkane)] systems studied by Bergman, Andersen, and Chirik.52 While 4 shows Cs symmetry in solution (two sets of pyrazole resonances), 5 displays a symmetry similar to that of 3 in solution, based on multinuclear NMR spectral studies, reflecting the larger separation between the CPh2 and tBu groups in 5 compared to the SiMe3 and tBu groups in 4. A solid-state structure confirmed retention of the N2 unit in 5, as well as the presence of a short Ti–N multiple bond [1.722(1) Å; Figure 17, top right]. Bond distances within the diazoalkane ligand in 5 [C–N, 1.297(2) Å; N–N, 1.327(2) Å] correspond reasonably well with bond distances in the hydrazone, Ph2CNNH2 [C–N, 1.286(2) Å; N–N, 1.371(3) Å].53 This hints at a dianionic diazoalkane ligand; notably, free diaryldiazoalkanes display much shorter N–N bond distances (1.12–1.16 Å).54
Figure 17.

Top: Solid-state structures of 4 (two polymorphs with two and three crystallographically independent molecules, respectively) and 5. Bottom: Solid-state structures of 6 and 7 (two crystallographically independent molecules). Thermal ellipsoids are at 50% probability. Solvent and H atoms (except on B and N) are omitted.
We also explored the reactivity of the high-spin, tetrahedral TiII ion in 2 with other small molecules. N2O reacts immediately upon contact with a benzene solution of 2 to form the terminal oxo [(TptBu,Me)Ti≡O(Cl)] (6) along with 3 (Scheme 2), due to the concomitant release of N2 and trapping by 2 equiv of the azophile 2. The greater affinity of 2 for N2 than N2O most likely results in capture of the former. However, one could also argue for a bimolecular mechanism55 involving a N2O complex,56 “[(TptBu,Me)Ti(N2O)(Cl)]”, reacting with 2. Regardless, the reaction mixture progresses over 18 h to eventually form the oxo complex in quantitative spectroscopic yield, and workup results in colorless crystals of 6 in 75% isolated yield (Scheme 2). Independently, treating 3 with 1 atm of N2O cleanly produces 6, thus rendering the former species a masked form of TiII ion.8e,57 The Ti≡O bond in 6 is remarkably short [1.621(1) Å; Figure 17, bottom left], possibly in line with triple-bond character; this is also manifested in the 1H NMR spectrum of 6, where the Me and tBu groups display broad resonances because of restricted rotation (e.g., via a Berry pseudorotation) around the central Ti≡O unit. This feature might contribute to the more restricted rotation in 4 compared to 3 and 5, but it is noteworthy that the 1H NMR spectrum of 7 (a titanium imide akin to 4, vide infra) shows only a single pyrazole environment.
The bicyclic amine 2,3:5,6-dibenzo-7-azabicyclo[2.2.1]hepta-2,5-diene (Hdbabh)58 delivers a parent imide moiety to the TiII ion of 2 by forming [(TptBu,Me)Ti≡NH(Cl)] (7) in near-quantitative spectroscopic yield (Scheme 2). Separation of 7 from the side product anthracene can be achieved via filtration of the reaction mixture through activated charcoal, providing yellow crystals of 7 in 32% isolated yield. The 1H–15N HSQC NMR spectrum of 7 unambiguously identifies the parent imide group through a cross-peak between a 1H triplet (4.97 ppm) and a downfield 15N nucleus (458 ppm; Figure S36). The same conclusion can be derived from the IR spectrum of 7, which exhibits a strong NH stretching vibration (3296 cm–1). Whereas terminal titanium–oxo ligands are not uncommon, terminal parent imides are quite a rare motif for group 4 transition metals.59 A solid-state structural study of 7 (Figure 17, bottom right) confirms the terminal parent imide motif and a geometry similar to that of 4 and 6. The imide H for both independent molecules was located in the difference Fourier map, and the Ti–N bond distances [1.674(4)–1.680(4) Å] fall within the range for terminal titanium imide complexes [1.627(8)–1.747(2) Å].59a−59d The transfer of a parent imide group to low-valent Ti, thereby forming a mononuclear group 4 transition-metal imide, is unprecedented. In a few instances, this type of reaction has been reported for group 5 metals, via NH group transfer from 2-methylaziridine to VIII and TaIII ions.60 Moreover, NH transfer reactions from aziridines or Hdbabh are known to generate transient imide functionalities, which are implicated in the formation of amide and nitride complexes of Ti, W, and Fe.61 The few known examples of titanium complexes having the parent imide group have been reported by several routes: deprotonation of NH3,59a protonation of a nitride salt,59d or via a nitridyl radical that effects H-atom abstraction59b,59c Along these lines, an alternative, nitridyl-mediated, route to 7 consists of treatment of the TiIII precursor 1 with NaN3 in THF. However, this reaction suffers from the sacrificial use of the [TptBu,MeTi] fragment as a H-atom source as well as the partial substitution of chloride for azide, thus giving mixtures of the corresponding azide and chloride imide complexes [(TptBu,Me)Ti≡NH(X)] [X– = N3 or Cl (7)] as well as other poorly defined decomposition products. The azide imide complex was identified via the characteristic 1H NMR resonance from the NH group (triplet at 3.58 ppm; Figure S33).
With the availability of structural and vibrational data for the new complexes 1, 2, adducts of 2, and the derived products (3–7), some general conclusions about their geometries and electronic structure can be inferred (Table 5). With 1 as the exception (τ5 = 0.27), the values of τ5 fall closely around 0.5, such that complexes 2-THF, 2-CNAd, and 3–7 do not lend themselves to a classification as being either trigonal-bipyramidal or square-pyramidal. The divalent nature of 2-THF and 2-CNAd can be directly seen from the long Ti–E and Ti–Cl distances compared to complexes 3–7, both reflecting the single-bond character of the Ti–E bonds and the large size of the TiII ion. Complexes 2-THF, 2-CNAd, and 3–7 underline the strong tendency for complex 2 to undergo reactions that render the Ti ion five-coordinate. As the reverse to this, complexes 4 and 7 might, in principle, eliminate small molecules (Me3SiCl and HCl) to generate a neutral titanium nitride functionality, but this mode of reactivity is not operational for the (TptBu,Me)Ti≡E(Cl) scaffold. Finally, inspection of the molecular structures of all complexes reported herein shows that the TiII complexes have Ti–Cl and B–H vectors oriented in an antiparallel fashion, whereas the TiIV complexes have the same vectors aligned nearly perpendicular. Indeed, the average B–Ti–Cl angles correlate well with the average Ti–B separations, as shown in Figure 18, revealing that all complexes group into three distinct regions when judged by their formal oxidation states. The link between metric variation and the formal oxidation state likely reflects the increasing ionic radius encountered upon going from TiIV to TiII; thus, a reduced Ti center tends to slide out of the TptBu,Me– ligand, accounting for the increase in Ti–B separations and the straightening of B–Ti–Cl angles. Importantly, complex 3 falls between the TiII and TiIII regions in Figure 18, augmenting support for the assignment of 3 as consisting of two TiIII centers covalently bound through an N22– ligand.
Table 5. Geometric Parameters and Vibrational Spectroscopic Data for Five-Coordinate Complexes 1, 2-THF, 2-CNAd, and 3–7.
| complex | E | τ5 | Ti–Cl (Å) | Ti–E (Å) | νBH (cm–1) |
|---|---|---|---|---|---|
| 1 | 0.27 | 2.2816(6)–2.3310(6) | 2569 | ||
| 2-THF | O | 0.46 | 2.414(1) | 2.251(2) | |
| 2-CNAd | C | 0.54 | 2.4123(5) | 2.207(1) | 2557 |
| 3 | N | 0.46–0.50 | 2.317(2)–2.360(2) | 1.805(3)–1.812(3) | 2559 |
| 4 | N | 0.46–0.59 | 2.3440(6)–2.3567(6) | 1.692(1)–1.700(2) | 2542 |
| 5 | N | 0.48 | 2.3486(4) | 1.722(1) | 2545 |
| 6 | O | 0.54 | 2.3468(6) | 1.621(1) | 2554 |
| 7 | N | 0.46–0.50 | 2.340(1)–2.348(1) | 1.674(4)–1.680(4) | 2543 |
Figure 18.
Average Ti–B distances versus Cl–Ti–B angles for the Ti complexes reported herein, representing formal oxidation states II–IV.
Conclusions
In summary, we have demonstrated the use of a sterically encumbering tris(pyrazolyl)borate ligand, TptBu,Me–, to generate the TiIII precursor, 1. Tuning of the steric profile of the Tp ligand (replacement of the 5-tBu group with Me) proved essential for obtaining a TiIII precursor with sufficient purity and stability for subsequent chemistry. Reduction of 1 under Ar furnishes a new geometry with a unique electronic structure: a stable tetrahedral TiII ion, 2, having two unpaired electrons. The Ti K-edge XAS spectrum of 2 is shifted to lower energy relative to structurally similar complexes bearing formally TiIII centers, consistent with 2 having a more reduced metal center. Complex 2 exhibits a 3A2 electronic ground state because of its confinement to a tetrahedral coordination geometry, as demonstrated by the magnetic moments extracted from SQUID magnetometric studies; the spin-only d2 behavior of 2 persists to temperatures as low as 10 K, whereupon ZFS results in a drastically lowered magnetic moment. HFEPR spectroscopic studies corroborate that 2 is a bona fide example of an S = 1 TiII ion via the observation of two spin triplet systems (ZFS parameters: D = −5.91 and +1.55 cm–1, respectively). The origin of two S = 1 systems can be traced to the two crystallographically independent molecules (conformers) in 2, which display geometric distortions around the tetrahedral TiII ion, e.g., B–Ti–Cl angles spanning 162.9(1)–174.6(1)°. This assertion is supported by classical LFT and by CASSCF/NEVPT2 calculations, which closely reproduce these ZFS parameters and depict excited-state manifolds consistent with d2 ions in an effective C3v environment. Complex 2 is a potent scavenger of N2, thus producing the dinuclear diamagnetic complex 3, where the N2 unit coordinates in a linear fashion between two Ti centers. The use of 15N2 to generate isotopically labeled 3-15N2 confirms atmospheric N2 as the source of the bridging N2 ligand. Labeling studies further identify a N–N stretching frequency by Raman spectroscopy (14N/15N at 1401.1/1346.3 cm–1), a Ti–N stretching frequency by IR vibrational spectroscopy (14N/15N at 769/749 cm–1), and a 15N NMR resonance at 309 ppm. Ti K-edge XAS supported by TDDFT calculations accord with these complexes bearing physically defined TiIII d1 centers. Taken together, these results show that the bonding in 3 can be viewed as a covalent interaction between two Ti centers and N2, producing a TiIII–N22––TiIII core. A combination of electrochemical, reactivity, spectroscopic, and structural studies demonstrate that THF coordinates reversibly to 2 with a fairly high association constant on the order of 5–8 M–1; the so-formed 2-THF adduct is impervious to N2 binding, implying the intermediacy of a four-coordinate geometry in reactions, where 1 is reduced to generate 2 as well as 3. Even when using a strong-field ligand, such as CNAd, complex 2 forms a five-coordinate adduct, 2-CNAd, while preserving the high-spin nature of the TiII ion, owing to only a moderate degree of π-back-bonding, as seen from the linear Ti–C–N geometry [174.2(1)°] and slight redshift of ν(C≡N) (28 cm–1). The availability of an open coordination site as well as two energetically high-lying d electrons renders tetrahedral 2 a versatile precursor to a family of five-coordinate TiIV complexes of the type [(TptBu,Me)Ti≡E(Cl)], with E2– = O, NH, NSiMe3, and N2CPh2. These products result from O-atom transfer from N2O and NH group transfer from the bicyclic amine Hdbabh to generate respectively the corresponding terminal oxide and parent terminal imide complexes. While N3SiMe3 reacts with 2 under N2 extrusion, N2CPh2 retains an intact N=N=C spine up to 100 °C.
We are presently scrutinizing the unique electronic structure of additional tetrahedral TiII complexes, as well as examining their potential for generating unusual metal–ligand multiple bonds.
Acknowledgments
A.R. gratefully acknowledges The Carlsberg Foundation (Grant CF18-0613) and the Independent Research Fund Denmark (Grant 9036-00015B) for funding. We thank the U.S. National Science Foundation (NSF; Grants CHE-0848248 and CHE-1152123 to D.J.M. and Grant CHE-1454455 to K.M.L.) and the University of Pennsylvania for financial support of this research. Part of this work was performed at the National High Magnetic Field Laboratory, which is supported by NSF Cooperative Agreement DMR-1644779 and the State of Florida. EPR studies at Northwestern University were supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences, under Contract DE-SC0019342 (to Prof. Brian M. Hoffman). XAS data were obtained at Stanford Synchrotron Radiation Lightsource (SSRL), which is supported by the U.S. DOE, Office of Science, Office of Basic Energy Sciences, under Contract DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the U.S. DOE, Office of Biological and Environmental Research, and by National Institutes of Health (NIH)/National Institute of General Medical Sciences (including P41GM103393). The authors also acknowledge the NIH Supplemental Awards 3R01GM118510-03S1 and 3R01GM087605-06S1 and financial support of the Vagelos Institute for Energy Sciences and Technology for the purchase of NMR instrument NEO600. K.M. acknowledges generous funding from FAU Erlangen-Nürnberg.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c02586.
NMR, XRD, cyclic voltammetry, SQUID, X- and Q-band EPR, HFEPR, Ti K-edge XAS, and computational data (PDF)
Accession Codes
CCDC 2005728–2005738 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge viawww.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Author Contributions
The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
The authors declare no competing financial interest.
Supplementary Material
References
- a McMurry J. E. Carbonyl-coupling reactions using low-valent titanium. Chem. Rev. 1989, 89, 1513–1524. 10.1021/cr00097a007. [DOI] [Google Scholar]; b Kulinkovich O. G.; Sviridov S. V.; Vasilevski D. A. Titanium(IV) Isopropoxide-Catalyzed Formation of 1-Substituted Cyclopropanols in the Reaction of Ethylmagnesium Bromide with Methyl Alkanecarboxylates. Synthesis 1991, 1991, 234–234. 10.1055/s-1991-26431. [DOI] [Google Scholar]; c Aleandri L. E.; Bogdanovicć B.; Gaidies A.; Jones D. J.; Liao S.; Michalowicz A.; Rozière J.; Schott A. [Ti(MgCl)2· xTHF]q: a reagent for the McMurry reaction and a novel inorganic Grignard complex. J. Organomet. Chem. 1993, 459, 87–93. 10.1016/0022-328X(93)86059-Q. [DOI] [Google Scholar]; d Fürstner A. Chemistry of and with Highly Reactive Metals. Angew. Chem. Angew. Chem., Int. Ed. Engl. 1993, 32, 164–189. 10.1002/anie.199301641. [DOI] [Google Scholar]; e Kulinkovich O. G.; de Meijere A. 1,n-Dicarbanionic Titanium Intermediates from Monocarbanionic Organometallics and Their Application in Organic Synthesis. Chem. Rev. 2000, 100, 2789–2834. 10.1021/cr980046z. [DOI] [PubMed] [Google Scholar]
- a Gilbert Z. W.; Hue R. J.; Tonks I. A. Catalytic formal [2 + 2+1] synthesis of pyrroles from alkynes and diazenes via TiII/TiIV redox catalysis. Nat. Chem. 2016, 8, 63–68. 10.1038/nchem.2386. [DOI] [PubMed] [Google Scholar]; b Davis-Gilbert Z. W.; Yao L. J.; Tonks I. A. Ti-Catalyzed Multicomponent Oxidative Carboamination of Alkynes with Alkenes and Diazenes. J. Am. Chem. Soc. 2016, 138, 14570–14573. 10.1021/jacs.6b09939. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Chiu H.-C.; Tonks I. A. Trimethylsilyl-Protected Alkynes as Selective Cross-Coupling Partners in Titanium-Catalyzed [2 + 2+1] Pyrrole Synthesis. Angew. Chem., Int. Ed. 2018, 57, 6090–6094. 10.1002/anie.201800595. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Davis-Gilbert Z. W.; Wen X.; Goodpaster J. D.; Tonks I. A. Mechanism of Ti-Catalyzed Oxidative Nitrene Transfer in [2 + 2 + 1] Pyrrole Synthesis from Alkynes and Azobenzene. J. Am. Chem. Soc. 2018, 140, 7267–7281. 10.1021/jacs.8b03546. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Pearce A. J.; See X. Y.; Tonks I. A. Oxidative nitrene transfer from azides to alkynes via Ti(II)/Ti(IV) redox catalysis: formal [2 + 2+1] synthesis of pyrroles. Chem. Commun. 2018, 54, 6891–6894. 10.1039/C8CC02623H. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Beaumier E. P.; McGreal M. E.; Pancoast A. R.; Wilson R. H.; Moore J. T.; Graziano B. J.; Goodpaster J. D.; Tonks I. A. Carbodiimide Synthesis via Ti-Catalyzed Nitrene Transfer from Diazenes to Isocyanides. ACS Catal. 2019, 9, 11753–11762. 10.1021/acscatal.9b04107. [DOI] [PMC free article] [PubMed] [Google Scholar]; g Chiu H.-C.; See X. Y.; Tonks I. A. Dative Directing Group Effects in Ti-Catalyzed [2 + 2+1] Pyrrole Synthesis: Chemo- and Regioselective Alkyne Heterocoupling. ACS Catal. 2019, 9, 216–223. 10.1021/acscatal.8b04669. [DOI] [PMC free article] [PubMed] [Google Scholar]; h Reiner B. R.; Tonks I. A. Group 4 Diarylmetallocenes as Bespoke Aryne Precursors for Titanium-Catalyzed [2 + 2 + 2] Cycloaddition of Arynes and Alkynes. Inorg. Chem. 2019, 58, 10508–10515. 10.1021/acs.inorgchem.9b01082. [DOI] [PMC free article] [PubMed] [Google Scholar]; i Beaumier E. P.; Gordon C. P.; Harkins R. P.; McGreal M. E.; Wen X.; Copéret C.; Goodpaster J. D.; Tonks I. A. Cp2Ti(κ2-tBuNCNtBu): A Complex with an Unusual κ2 Coordination Mode of a Heterocumulene Featuring a Free Carbene. J. Am. Chem. Soc. 2020, 142, 8006–8018. 10.1021/jacs.0c02487. [DOI] [PMC free article] [PubMed] [Google Scholar]; j Pearce A. J.; Harkins R. P.; Reiner B. R.; Wotal A. C.; Dunscomb R. J.; Tonks I. A. Multicomponent Pyrazole Synthesis from Alkynes, Nitriles, and Titanium Imido Complexes via Oxidatively Induced N–N Bond Coupling. J. Am. Chem. Soc. 2020, 142, 4390–4399. 10.1021/jacs.9b13173. [DOI] [PMC free article] [PubMed] [Google Scholar]; k Beaumier E. P.; Ott A. A.; Wen X.; Davis-Gilbert Z. W.; Wheeler T. A.; Topczewski J. J.; Goodpaster J. D.; Tonks I. A. Ti-catalyzed ring-opening oxidative amination of methylenecyclopropanes with diazenes. Chem. Sci. 2020, 11, 7204. 10.1039/D0SC01998D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Girolami G. S.; Wilkinson G.; Galas A. M. R.; Thornton-Pett M.; Hursthouse M. B. Synthesis and Properties of the Divalent 1,2-Bis(dimethylphosphino)ethane (dmpe) Complexes MCl2(dmpe)2 and MMe2(dmpe)2 (M = Ti, V, Cr, Mn, or Fe). X-Ray Crystal Structures of MCl2(dmpe)2 (M = Ti, V, or Cr), MnBr2(dmpe)2, TiMe1.3Cl0.7(dmpe)2, and CrMe2(dmpe)2. J. Chem. Soc., Dalton Trans. 1985, 1339–1348. 10.1039/dt9850001339. [DOI] [Google Scholar]; b Jensen J. A.; Wilson S. R.; Schultz A. J.; Girolami G. S. Divalent Titanium Chemistry. Synthesis, Reactivity, and X-ray and Neutron Diffraction Studies of Ti(BH4)2(dmpe)2 and Ti(CH3)2(dmpe)2. J. Am. Chem. Soc. 1987, 109, 8094–8096. 10.1021/ja00260a029. [DOI] [Google Scholar]; c Jensen J. A.; Girolami G. S. Synthesis, Characterization, and X-ray Crystal Structures of the Divalent Titanium Complex Ti(η2-BH4)2(dmpe)2 and the unidentate tetrahydroborate complex V(η1-BH4)2(dmpe)2. Inorg. Chem. 1989, 28, 2107–2113. 10.1021/ic00310a019. [DOI] [Google Scholar]; d Morris R. J.; Girolami G. S. On the π-Donor Ability of Early Transition Metals: Evidence That Trialkylphosphines Can Engage in π-Back-Bonding and X-ray Structure of the Titanium(II) Phenoxide Ti(OPh)2(dmpe)2. Inorg. Chem. 1990, 29, 4167–4169. 10.1021/ic00346a001. [DOI] [Google Scholar]
- a Edema J. J. H.; Duchateau R.; Gambarotta S.; Hynes R.; Gabe E. Novel Titanium(II) Amine Complexes L4TiCl2 [L = 1/2N,N,N′,N′-tetramethylethylenediamine (TMEDA), 1/2N,N,N′-trimethylethylenediamine, pyridine, 1/2 2,2′-bipyridine]: Synthesis and Crystal Structure of Monomeric trans-(TMEDA)2TiCl2. Inorg. Chem. 1991, 30, 154–156. 10.1021/ic00002a002. [DOI] [Google Scholar]; b Zolnhofer E. M.; Wijeratne G. B.; Jackson T. A.; Fortier S.; Heinemann F. W.; Meyer K.; Krzystek J.; Ozarowski A.; Mindiola D. J.; Telser J. Electronic Structure and Magnetic Properties of a Titanium(II) Coordination Complex. Inorg. Chem. 2020, 59, 6187–6201. 10.1021/acs.inorgchem.0c00311. [DOI] [PubMed] [Google Scholar]; c Araya M. A.; Cotton F. A.; Matonic J. H.; Murillo C. A. An Efficient Reduction Process Leading to Titanium(II) and Niobium(II): Preparation and Structural Characterization of trans-MCl2(py)4 Compounds, M = Ti, Nb, and Mn. Inorg. Chem. 1995, 34, 5424–5428. 10.1021/ic00126a009. [DOI] [Google Scholar]; d Wijeratne G. B.; Zolnhofer E. M.; Fortier S.; Grant L. N.; Carroll P. J.; Chen C.-H.; Meyer K.; Krzystek J.; Ozarowski A.; Jackson T. A.; Mindiola D. J.; Telser J. Electronic Structure and Reactivity of a Well-Defined Mononuclear Complex of Ti(II). Inorg. Chem. 2015, 54, 10380–10397. 10.1021/acs.inorgchem.5b01796. [DOI] [PubMed] [Google Scholar]
- a Woo L. K.; Hays J. A.; Young V. G.; Day C. L.; Caron C.; D’Souza F.; Kadish K. M. Synthesis, Characterization, Substitution, and Atom-Transfer Reactions of (η2-Alkyne)(tetratolylporphyrinato)titanium(II). X-ray Structure of trans-Bis(4-picoline)(tetratolylporphyrinato)titanium(II). Inorg. Chem. 1993, 32, 4186–4192. 10.1021/ic00072a007. [DOI] [Google Scholar]; b Thorman J. L.; Young V. G.; Boyd P. D. W.; Guzei I. A.; Woo L. K. Atom Transfer Reactions of (TTP)Ti(η2-3-hexyne): Synthesis and Molecular Structure of trans-(TTP)Ti[OP(Oct)3]2. Inorg. Chem. 2001, 40, 499–506. 10.1021/ic0003426. [DOI] [PubMed] [Google Scholar]; c Wang X.; Gray S. D.; Chen J.; Woo L. K. Facile Syntheses of Titanium(II), Tin(II), and Vanadium(II) Porphyrin Complexes through Homogeneous Reduction. Reactivity of trans-(TTP)TiL2 (L = THF, t-BuNC). Inorg. Chem. 1998, 37, 5–9. 10.1021/ic970961n. [DOI] [PubMed] [Google Scholar]
- Kayal A.; Kuncheria J.; Lee S. C. Bis[hydrotris(pyrazol-1-yl)borato]titanium(II): a stable Tp2M complex of singular reactivity. Chem. Commun. 2001, 2482–2483. 10.1039/b108115b. [DOI] [PubMed] [Google Scholar]
- a Hagadorn J. R.; Arnold J. Low-Valent Chemistry of Titanium Benzamidinates Leading to New Ti μ-N2, μ-O, Alkyl Derivatives, and the Cyclometalation of TMEDA. J. Am. Chem. Soc. 1996, 118, 893–894. 10.1021/ja953449e. [DOI] [Google Scholar]; b Hagadorn J. R.; Arnold J. Titanium(II), -(III), and -(IV) Complexes Supported by Benzamidinate Ligands. Organometallics 1998, 17, 1355–1368. 10.1021/om970933c. [DOI] [Google Scholar]; c Baumann R.; Stumpf R.; Davis W. M.; Liang L.-C.; Schrock R. R. Titanium and Zirconium Complexes That Contain the Tridentate Diamido Ligands [(i-PrN-o-C6H4)2O]2- ([i-PrNON]2-) and [(C6H11N-o-C6H4)2O]2- ([CyNON]2-). J. Am. Chem. Soc. 1999, 121, 7822–7836. 10.1021/ja983549v. [DOI] [Google Scholar]; d Bai G.; Wei P.; Stephan D. W. Reductions of β-Diketiminato–Titanium(III) Complexes. Organometallics 2006, 25, 2649–2655. 10.1021/om060076p. [DOI] [Google Scholar]; e Chomitz W. A.; Arnold J. Transition metal dinitrogen complexes supported by a versatile monoanionic [N2P2] ligand. Chem. Commun. 2007, 4797–4799. 10.1039/b709763h. [DOI] [PubMed] [Google Scholar]; f Hanna T. E.; Bernskoetter W. H.; Bouwkamp M. W.; Lobkovsky E.; Chirik P. J. Bis(cyclopentadienyl) Titanium Dinitrogen Chemistry: Synthesis and Characterization of a Side-on Bound Haptomer. Organometallics 2007, 26, 2431–2438. 10.1021/om0611913. [DOI] [Google Scholar]; g Semproni S. P.; Milsmann C.; Chirik P. J. Side-on Dinitrogen Complexes of Titanocenes with Disubstituted Cyclopentadienyl Ligands: Synthesis, Structure, and Spectroscopic Characterization. Organometallics 2012, 31, 3672–3682. 10.1021/om300156z. [DOI] [Google Scholar]; h Kurogi T.; Ishida Y.; Kawaguchi H. Synthesis of titanium and zirconium complexes supported by a p-terphenoxide ligand and their reactions with N2, CO2 and CS2. Chem. Commun. 2013, 49, 11755–11757. 10.1039/c3cc47284a. [DOI] [PubMed] [Google Scholar]; i Doyle L. R.; Wooles A. J.; Jenkins L. C.; Tuna F.; McInnes E. J. L.; Liddle S. T. Catalytic Dinitrogen Reduction to Ammonia at a Triamidoamine–Titanium Complex. Angew. Chem., Int. Ed. 2018, 57, 6314–6318. 10.1002/anie.201802576. [DOI] [PMC free article] [PubMed] [Google Scholar]; j Sekiguchi Y.; Meng F.; Tanaka H.; Eizawa A.; Arashiba K.; Nakajima K.; Yoshizawa K.; Nishibayashi Y. Synthesis and reactivity of titanium- and zirconium-dinitrogen complexes bearing anionic pyrrole-based PNP-type pincer ligands. Dalton Trans. 2018, 47, 11322–11326. 10.1039/C8DT02739K. [DOI] [PubMed] [Google Scholar]; k Ghana P.; van Krüchten F. D.; Spaniol T. P.; van Leusen J.; Kögerler P.; Okuda J. Conversion of dinitrogen to tris(trimethylsilyl)amine catalyzed by titanium triamido-amine complexes. Chem. Commun. 2019, 55, 3231–3234. 10.1039/C8CC09742A. [DOI] [PubMed] [Google Scholar]; l Nakanishi Y.; Ishida Y.; Kawaguchi H. Nitrogen–Carbon Bond Formation by Reactions of a Titanium–Potassium Dinitrogen Complex with Carbon Dioxide, tert-Butyl Isocyanate, and Phenylallene. Angew. Chem., Int. Ed. 2017, 56, 9193–9197. 10.1002/anie.201704286. [DOI] [PubMed] [Google Scholar]
- a Atwood J. L.; Stone K. E.; Alt H. G.; Hrncir D. C.; Rausch M. D. The crystal structure of dicarbonyldicyclopentadienyltitanium(II), (η5-C5H5)2Ti(CO)2. J. Organomet. Chem. 1977, 132, 367–375. 10.1016/S0022-328X(00)93722-7. [DOI] [Google Scholar]; b Hollis T. K.; Ahn Y. J.; Tham F. S. Low-Valent Titanium Bis(phospholyl) Chemistry: A Configurationally Stable Chiral Phosphatitanocene. Organometallics 2003, 22, 1432–1436. 10.1021/om020961h. [DOI] [Google Scholar]; c Hanna T. E.; Lobkovsky E.; Chirik P. J. Mono(dinitrogen) and Carbon Monoxide Adducts of Bis(cyclopentadienyl) Titanium Sandwiches. J. Am. Chem. Soc. 2006, 128, 6018–6019. 10.1021/ja061213c. [DOI] [PubMed] [Google Scholar]; d Kilpatrick A. F. R.; Cloke F. G. N. Reductive deoxygenation of CO2 by a bimetallic titanium bis(pentalene) complex. Chem. Commun. 2014, 50, 2769–2771. 10.1039/C3CC48379G. [DOI] [PubMed] [Google Scholar]; e Aguilar-Calderón J. R.; Metta-Magaña A. J.; Noll B.; Fortier S. C(sp3)–H Oxidative Addition and Transfer Hydrogenation Chemistry of a Titanium(II) Synthon: Mimicry of Late-Metal Type Reactivity. Angew. Chem., Int. Ed. 2016, 55, 14101–14105. 10.1002/anie.201607441. [DOI] [PubMed] [Google Scholar]
- a Kool L. B.; Rausch M. D.; Herberhold M.; Alt H. G.; Thewalt U.; Honold B. Diamagnetic isocyanide complexes of titanium, zirconium, and hafnium. Organometallics 1986, 5, 2465–2468. 10.1021/om00143a010. [DOI] [Google Scholar]; b Cuenca T.; Gómez R.; Goḿez-Sal P.; Royo P. Synthesis and characterization of ansa-dimethylsilylbiscyclopentadienyl titanium(II) complexes. Crystal structure of [Ti{Me2Si(C5H4)2}{CN(2,6-Me2C6H3)}2]. J. Organomet. Chem. 1993, 454, 105–111. 10.1016/0022-328X(93)83230-S. [DOI] [Google Scholar]; c Haehnel M.; Ruhmann M.; Theilmann O.; Roy S.; Beweries T.; Arndt P.; Spannenberg A.; Villinger A.; Jemmis E. D.; Schulz A.; Rosenthal U. Reactions of Titanocene Bis(trimethylsilyl)acetylene Complexes with Carbodiimides: An Experimental and Theoretical Study of Complexation versus C–N Bond Activation. J. Am. Chem. Soc. 2012, 134, 15979–15991. 10.1021/ja3070649. [DOI] [PubMed] [Google Scholar]; d Altenburger K.; Arndt P.; Becker L.; Reiß F.; Burlakov V. V.; Spannenberg A.; Baumann W.; Rosenthal U. Multiple and Highly Selective Alkyne–Isonitrile C–C and C–N Couplings at Group 4 Metallocenes. Chem. - Eur. J. 2016, 22, 9169–9180. 10.1002/chem.201601465. [DOI] [PubMed] [Google Scholar]
- a McPherson A. M.; Fieselmann B. F.; Lichtenberger D. L.; McPherson G. L.; Stucky G. D. Electronic Properties of Bis(η5-cyclopentadienyl)titanium 2,2′-Bipyridyl. A Singlet Molecule with a Low-Lying Triplet Excited State. J. Am. Chem. Soc. 1979, 101, 3425–3430. 10.1021/ja00507a001. [DOI] [Google Scholar]; b Durfee L. D.; Fanwick P. E.; Rothwell I. P.; Folting K.; Huffman J. C. Reductive Elimination Pathways to Low Valent Titanium Aryl Oxide Complexes. J. Am. Chem. Soc. 1987, 109, 4720–4722. 10.1021/ja00249a045. [DOI] [PubMed] [Google Scholar]; c Piglosiewicz I. M.; Beckhaus R.; Saak W.; Haase D. Dehydroaromatization of Quinoxalines: One-Step Syntheses of Trinuclear 1,6,7,12,13,18-Hexaazatrinaphthylene Titanium Complexes. J. Am. Chem. Soc. 2005, 127, 14190–14191. 10.1021/ja054740p. [DOI] [PubMed] [Google Scholar]
- a Corbin D. R.; Willis W. S.; Duesler E. N.; Stucky G. D. Intramolecular Electron-Transfer Induced Carbon-Hydrogen Bond Dissociation in Methyl-Substituted 1,10-Phenanthroline Complexes of Bis(η5-cyclopentadienyl)titanium. J. Am. Chem. Soc. 1980, 102, 5969–5971. 10.1021/ja00538a074. [DOI] [Google Scholar]; b Wolff C.; Gottschlich A.; England J.; Wieghardt K.; Saak W.; Haase D.; Beckhaus R. , Molecular and Electronic Structures of Mononuclear and Dinuclear Titanium Complexes Containing π-Radical Anions of 2,2′-Bipyridine and 1,10-Phenanthroline: An Experimental and DFT Computational Study. Inorg. Chem. 2015, 54, 4811–4820. 10.1021/acs.inorgchem.5b00285. [DOI] [PubMed] [Google Scholar]
- a Schmid G.; Thewalt U.; Polášek M.; Mach K.; Sedmera P. η5-Pentabenzylcyclopentadienyl derivatives of titanium (IV), (III), and (II). The crystal structures of (η5-C5H5)(η5-C5Bz5)TiCl2 (Bz = benzyl), (η5-C5H5)(η5- C5Bz5)TiCl, and (η5-C5H5)(η5-C5Bz5)Ti[η2-(CSiMe3)2]. J. Organomet. Chem. 1994, 482, 231–241. 10.1016/0022-328X(94)88206-1. [DOI] [Google Scholar]; b Noor A.; Kempe R. Acetylenetitanium Complex Stabilized by Aminopyridinato Ligands. Eur. J. Inorg. Chem. 2008, 2008, 2377–2381. 10.1002/ejic.200800132. [DOI] [Google Scholar]; c Lamač M.; Spannenberg A.; Baumann W.; Jiao H.; Fischer C.; Hansen S.; Arndt P.; Rosenthal U. Si–H Bond Activation of Alkynylsilanes by Group 4 Metallocene Complexes. J. Am. Chem. Soc. 2010, 132, 4369–4380. 10.1021/ja910527w. [DOI] [PubMed] [Google Scholar]; d Haehnel M.; Hansen S.; Schubert K.; Arndt P.; Spannenberg A.; Jiao H.; Rosenthal U. Synthesis, Characterization and Reactivity of Group 4 Metallocene Bis(diphenylphosphino)acetylene Complexes—A Reactivity and Bonding Study. J. Am. Chem. Soc. 2013, 135, 17556–17565. 10.1021/ja409320k. [DOI] [PubMed] [Google Scholar]; e Pinkas J.; Gyepes R.; Císařová I.; Kubišta J.; Horáček M.; Mach K. Displacement of ethene from the decamethyltitanocene-ethene complex with internal alkynes, substituent-dependent alkyne-to-allene rearrangement, and the electronic transition relevant to the back-bonding interaction. Dalton Trans. 2015, 44, 7276–7291. 10.1039/C5DT00351B. [DOI] [PubMed] [Google Scholar]
- a Cohen S. A.; Auburn P. R.; Bercaw J. E. Structure and reactivity of bis(pentamethylcyclopentadienyl)(ethylene)titanium(II), a simple olefin adduct of titanium. J. Am. Chem. Soc. 1983, 105, 1136–1143. 10.1021/ja00343a012. [DOI] [Google Scholar]; b Horáček M.; Kupfer V.; Thewalt U.; Štěpnička P.; Polášek M.; Mach K. Bis[η5-tetramethyl(trimethylsilyl)cyclopentadienyl]titanium(II) and Its π-Complexes with Bis(trimethylsilyl)acetylene and Ethylene. Organometallics 1999, 18, 3572–3578. 10.1021/om990286k. [DOI] [Google Scholar]; c Pinkas J.; Císařová I.; Gyepes R.; Kubišta J.; Horáček M.; Mach K. Ethene Complexes of Bulky Titanocenes, Their Thermolysis, and Their Reactivity toward 2-Butyne. Organometallics 2012, 31, 5478–5493. 10.1021/om300461k. [DOI] [Google Scholar]
- a Breil H.; Wilke G. Di(cyclooctatetraene)titanium and Tri(cyclooctatetraene)dititanium. Angew. Chem., Int. Ed. Engl. 1966, 5, 898–899. 10.1002/anie.196608982. [DOI] [Google Scholar]; b Dietrich H.; Soltwisch M. Crystal Structure of Bis(cyclooctatetraene)titanium. Angew. Chem., Int. Ed. Engl. 1969, 8, 765–765. 10.1002/anie.196907652. [DOI] [Google Scholar]; c Cloke F. G. N.; Green J. C.; Hitchcock P. B.; Joseph S. C. P.; Mountford P.; Kaltsoyannis N.; McCamley A. Molecular and electronic structures of bis[1,4-bis(trimethylsilyl)cyclooctatetraene] sandwich complexes of titanium and zirconium. J. Chem. Soc., Dalton Trans. 1994, 2867–2874. 10.1039/dt9940002867. [DOI] [Google Scholar]; d Horáček M.; Hiller J.; Thewalt U.; Štěpnička P.; Mach K. Bis[(η8-cyclooctatetraene)titanium] complex with perpendicularly bridging bis(trimethylsilyl)acetylene. J. Organomet. Chem. 1998, 571, 77–82. 10.1016/S0022-328X(98)00870-5. [DOI] [Google Scholar]
- a Sikora D. J.; Rausch M. D.; Rogers R. D.; Atwood J. L. Formation and Molecular Structure of Bis(η5-cyclopentadienyl)bis(trifluorophosphine)titanium. J. Am. Chem. Soc. 1981, 103, 982–984. 10.1021/ja00394a065. [DOI] [Google Scholar]; b Arp H.; Baumgartner J.; Marschner C.; Zark P.; Müller T. Coordination Chemistry of Cyclic Disilylated Stannylenes and Plumbylenes to Group 4 Metallocenes. J. Am. Chem. Soc. 2012, 134, 10864–10875. 10.1021/ja301547x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Hitchcock P. B.; Kerton F. M.; Lawless G. A. The Elusive Titanocene. J. Am. Chem. Soc. 1998, 120, 10264–10265. 10.1021/ja981934e. [DOI] [Google Scholar]; b Kilpatrick A. F. R.; Green J. C.; Cloke F. G. N.; Tsoureas N. Bis(pentalene)di-titanium: a bent double-sandwich complex with a very short Ti–Ti bond. Chem. Commun. 2013, 49, 9434–9436. 10.1039/c3cc45187a. [DOI] [PubMed] [Google Scholar]
- Edema J. J. H.; Gambarotta S.; Duchateau R.; Bensimon C. Labile triangulo-trititanium(II) and trivanadium(II) clusters. Inorg. Chem. 1991, 30, 3585–3587. 10.1021/ic00019a001. [DOI] [Google Scholar]
- Ma W.; Zhang J.-X.; Lin Z.; Tilley T. D.; Ye Q. Synthesis, structure and DFT calculations of mononuclear cyclic (alkyl)(amino) carbene supported titanium(II) complexes. Dalton Trans. 2019, 48, 14962–14965. 10.1039/C9DT03342D. [DOI] [PubMed] [Google Scholar]
- Liu Q.; Chen Q.; Leng X.; Deng Q.-H.; Deng L. Hafnium(II) Complexes with Cyclic (Alkyl)(amino)carbene Ligation. Organometallics 2018, 37, 4186–4188. 10.1021/acs.organomet.8b00688. [DOI] [Google Scholar]
- a Trofimenko S. Coordination chemistry of pyrazole-derived ligands. Chem. Rev. 1972, 72, 497–509. 10.1021/cr60279a003. [DOI] [Google Scholar]; b Trofimenko S. Recent Advances in Poly(pyrazolyl)borate (Scorpionate) Chemistry. Chem. Rev. 1993, 93, 943–980. 10.1021/cr00019a006. [DOI] [Google Scholar]
- Dowling C. M.; Leslie D.; Chisholm M. H.; Parkin G. The Synthesis and Structural Characterization of the Sterically Demanding Tris(3,5-di-t-butylpyrazolyl)hydroborato Ligand, [TpBut2]: A Highly Twisted, Propeller-Like, Ligand System. Main Group Chem. 1995, 1, 29–52. 10.1080/13583149512331338245. [DOI] [Google Scholar]
- Petrov P. A.; Smolentsev A. I.; Bogomyakov A. S.; Konchenko S. N. Novel vanadium complexes supported by a bulky tris(pyrazolyl)borate ligand. Polyhedron 2017, 129, 60–64. 10.1016/j.poly.2017.03.033. [DOI] [Google Scholar]
- Kersten J. L.; Kucharczyk R. R.; Yap G. P. A.; Rheingold A. L.; Theopold K. H. [(TptBu,Me)CrR]: A New Class of Mononuclear, Coordinatively Unsaturated Chromium(II) Alkyls with cis-Divacant Octahedral Structure. Chem. - Eur. J. 1997, 3, 1668–1674. 10.1002/chem.19970031016. [DOI] [Google Scholar]
- a Brunker T. J.; Hascall T.; Cowley A. R.; Rees L. H.; O’Hare D. Variable Coordination Modes of Hydrotris(3-isopropyl-4-bromopyrazolyl)borate (Tp′) in Fe(II), Mn(II), Cr(II), and Cr(III) Complexes: Formation of MTp′Cl (M = Fe and Mn), Structural Isomerism in CrTp′2, and the Observation of Tp′ – as an Uncoordinated Anion. Inorg. Chem. 2001, 40, 3170–3176. 10.1021/ic010125k. [DOI] [PubMed] [Google Scholar]; b Nabika M.; Seki Y.; Miyatake T.; Ishikawa Y.; Okamoto K.-i.; Fujisawa K. Manganese Catalysts with Scorpionate Ligands for Olefin Polymerization. Organometallics 2004, 23, 4335–4337. 10.1021/om049756n. [DOI] [Google Scholar]; c Tietz T.; Limberg C.; Stößer R.; Ziemer B. Four-Coordinate Trispyrazolylboratomanganese and -iron Complexes with a Pyrazolato Co-ligand: Syntheses and Properties as Oxidation Catalysts. Chem. - Eur. J. 2011, 17, 10010–10020. 10.1002/chem.201100343. [DOI] [PubMed] [Google Scholar]; d Gorrell I. B.; Parkin G. (Tris-(3-tert-butylpyrazolyl)hydroborato)manganese(II), -iron(II), -cobalt(II), and -nickel(II) Halide Derivatives: Facile Abstraction of Fluoride from [BF4]−. Inorg. Chem. 1990, 29, 2452–2456. 10.1021/ic00338a013. [DOI] [Google Scholar]
- a Jové F. A.; Pariya C.; Scoblete M.; Yap G. P. A.; Theopold K. H. A Family of Four-Coordinate Iron(II) Complexes Bearing the Sterically Hindered Tris(pyrazolyl)borato Ligand TptBu,Me. Chem. - Eur. J. 2011, 17, 1310–1318. 10.1002/chem.201001792. [DOI] [PubMed] [Google Scholar]; b Fujisawa K.; Soma S.; Kurihara H.; Ohta A.; Dong H. T.; Minakawa Y.; Zhao J.; Alp E. E.; Hu M. Y.; Lehnert N. Stable Ferrous Mononitroxyl {FeNO}8 Complex with a Hindered Hydrotris(pyrazolyl)borate Coligand: Structure, Spectroscopic Characterization, and Reactivity Toward NO and O2. Inorg. Chem. 2019, 58, 4059–4062. 10.1021/acs.inorgchem.9b00107. [DOI] [PubMed] [Google Scholar]
- a Uehara K.; Hikichi S.; Akita M. Highly labile cationic tris-acetonitrile complexes, [TpRM(NCMe)3]OTf (M = Ni, Co; TpR: hydrotrispyrazolylborato, R = Ph, Me and iPr2): versatile precursors for TpR-containing nickel and cobalt complexes. J. Chem. Soc., Dalton Trans. 2002, 3529–3538. 10.1039/b203377c. [DOI] [Google Scholar]; b Rheingold A. L.; Liable-Sands L. M.; Golan J. A.; Trofimenko S. Remote Rotamer Control: The Effect of a 4-tert-Butyl Group on the Coordination Chemistry of TpR Ligands. Eur. J. Inorg. Chem. 2003, 2003, 2767–2773. 10.1002/ejic.200300005. [DOI] [Google Scholar]; c Rheingold A. L.; Liable-Sands L. M.; Golen J. A.; Yap G. P. A.; Trofimenko S. The coordination chemistry of the hydrotris(3-diphenylmethylpyrazol-1-yl)borate (TpCHPh2) ligand. Dalton Trans. 2004, 598–604. 10.1039/b313785f. [DOI] [PubMed] [Google Scholar]; d Trofimenko S.; Rheingold A. L.; Liable Sands L. M. Coordination Chemistry of Novel Scorpionate Ligands Based on 3-Cyclohexylpyrazole and 3-Cyclohexyl-4-bromopyrazole. Inorg. Chem. 2002, 41, 1889–1896. 10.1021/ic011186l. [DOI] [PubMed] [Google Scholar]; e Krzystek J.; Swenson D. C.; Zvyagin S. A.; Smirnov D.; Ozarowski A.; Telser J. Cobalt(II) “Scorpionate” Complexes as Models for Cobalt-Substituted Zinc Enzymes: Electronic Structure Investigation by High-Frequency and -Field Electron Paramagnetic Resonance Spectroscopy. J. Am. Chem. Soc. 2010, 132, 5241–5253. 10.1021/ja910766w. [DOI] [PubMed] [Google Scholar]
- a Guo S.; Ding E.; Yin Y.; Yu K. Synthesis and structures of tris(pyrazolyl)hydroborato metal complexes as structural model compounds of carbonicanhydrase. Polyhedron 1998, 17, 3841–3849. 10.1016/S0277-5387(98)00139-9. [DOI] [Google Scholar]; b Belderraín T. R.; Paneque M.; Carmona E.; Gutiérrez-Puebla E.; Monge M. A.; Ruiz-Valero C. Three-Center, Two-Electron M···H–B Bonds in Complexes of Ni, Co, and Fe and the Dihydrobis(3-tert-butylpyrazolyl)borate Ligand. Inorg. Chem. 2002, 41, 425–428. 10.1021/ic010598r. [DOI] [PubMed] [Google Scholar]; c Kunrath F. A.; de Souza R. F.; Casagrande O. L.; Brooks N. R.; Young V. G. Highly Selective Nickel Ethylene Oligomerization Catalysts Based on Sterically Hindered Tris(pyrazolyl)borate Ligands. Organometallics 2003, 22, 4739–4743. 10.1021/om034035u. [DOI] [Google Scholar]; d Frampton A. K.; Gartland K.; Piro N. A.; Kassel W. S.; Dougherty W. G. Structural characterization and electrochemical properties of nickel (II) complexes bearing sterically bulky hydrotris(3-phenyl)- and hydrotris(3-tert-butylpyrazol-1-yl)borato ligands. Polyhedron 2016, 114, 172–178. 10.1016/j.poly.2015.11.032. [DOI] [Google Scholar]
- a Kitajima N.; Fujisawa K.; Morooka Y. Tetrahedral copper(II) complexes supported by a hindered pyrazolylborate. Formation of the thiolato complex, which closely mimics the spectroscopic characteristics of blue copper proteins. J. Am. Chem. Soc. 1990, 112, 3210–3212. 10.1021/ja00164a052. [DOI] [Google Scholar]; b Yoon K.; Parkin G. Tris(3-t-butyl-5-methylpyrazolyl)hydroborato derivatives of copper and thallium: The structural influence of a 5-methyl substituent. Polyhedron 1995, 14, 811–821. 10.1016/0277-5387(94)00300-4. [DOI] [Google Scholar]; c Higashimura H.; Kubota M.; Shiga A.; Fujisawa K.; Moro-oka Y.; Uyama H.; Kobayashi S. Radical-Controlled” Oxidative Polymerization of 4-Phenoxyphenol by a Tyrosinase Model Complex Catalyst to Poly(1,4-phenylene oxide). Macromolecules 2000, 33, 1986–1995. 10.1021/ma991635p. [DOI] [Google Scholar]; d Fujisawa K.; Tada N.; Ishikawa Y.; Higashimura H.; Miyashita Y.; Okamoto K.-i. The most hindered hydrotris(pyrazolyl)borate ligand, X-ray structure of chlorocopper(II) complex: [Cu(Cl){HB(3-Ad-5-Pripz)3}] as compared with [Cu(Cl){HB(3-But-5-Pripz)3}]. Inorg. Chem. Commun. 2004, 7, 209–212. 10.1016/j.inoche.2003.10.030. [DOI] [Google Scholar]
- a Yoon K.; Parkin G. Artificial Manipulation of Apparent Bond Lengths as Determined by Single-Crystal X-ray Diffraction. J. Am. Chem. Soc. 1991, 113, 8414–8418. 10.1021/ja00022a032. [DOI] [Google Scholar]; b Yoon K.; Parkin G. Resolved and Unresolved Crystallographic Disorder Between {η3-HB(3-Butpz)3}ZnCN and {η3-HB(3-Butpz)3}ZnX (X = Cl, Br, I). Inorg. Chem. 1992, 31, 1656–1662. 10.1021/ic00035a026. [DOI] [Google Scholar]
- a Adams H.; Batten S. R.; Davies G. M.; Duriska M. B.; Jeffery J. C.; Jensen P.; Lu J.; Motson G. R.; Coles S. J.; Hursthouse M. B.; Ward M. D. New bis-, tris- and tetrakis(pyrazolyl)borate ligands with 3-pyridyl and 4-pyridyl substituents: synthesis and coordination chemistry. Dalton Trans. 2005, 1910–1923. 10.1039/b502892b. [DOI] [PubMed] [Google Scholar]; b Pérez Olmo C.; Böhmerle K.; Steinfeld G.; Vahrenkamp H. New Polar Pyrazolylborate Ligands and Their Basic Zinc Complex Chemistry. Eur. J. Inorg. Chem. 2006, 2006, 3869–3877. 10.1002/ejic.200600382. [DOI] [Google Scholar]; c Vitze H.; Bolte M.; Lerner H.-W.; Wagner M. Third-Generation Scorpionates [RBpz3]− – How Influential Is the Nondonor Substituent R?. Eur. J. Inorg. Chem. 2016, 2016, 2443–2454. 10.1002/ejic.201500801. [DOI] [Google Scholar]
- a Shi X.; Liu Z.; Cheng J. Barium tetraalkylaluminate complexes supported by the super-bulky hydrotris(pyrazolyl)borate ligand. Dalton Trans. 2019, 48, 17919–17924. 10.1039/C9DT04182F. [DOI] [PubMed] [Google Scholar]; b Shi X.; Cheng J. Reversible addition and hydrogenation of 1,1-diphenylethylene with a barium complex. Dalton Trans. 2019, 48, 8565–8568. 10.1039/C9DT01874C. [DOI] [PubMed] [Google Scholar]
- Hillier A. C.; Liu S.-Y.; Sella A.; Elsegood M. R. J. Lanthanide Chalcogenolate Complexes: Synthesis and Crystal Structures of the Isoleptic Series [Sm(TpMe,Me)2ER] (E = O, S, Se, Te; TpMe,Me = tris-3,5-Dimethylpyrazolylborate). Inorg. Chem. 2000, 39, 2635–2644. 10.1021/ic9914793. [DOI] [PubMed] [Google Scholar]
- Ferrence G. M.; McDonald R.; Morissette M.; Takats J. Mixed pyrazolylborate/cyclopentadienyl derivatives of divalent lanthanides: synthesis and structure of (TptBu,Me)Yb(C5H4R) (R = H, SiMe3). J. Organomet. Chem. 2000, 596, 95–101. 10.1016/S0022-328X(99)00559-8. [DOI] [Google Scholar]
- Sun Y.; McDonald R.; Takats J.; Day V. W.; Eberspacher T. A. Synthesis and Structure of Bis[hydrotris(3,5-dimethylpyrazolyl)borato]iodouranium(III), U[HB(3,5-Me2pz)3]2I: Unprecedented Side-On Interaction Involving a Hydrotris(pyrazolyl)borate Ligand. Inorg. Chem. 1994, 33, 4433–4434. 10.1021/ic00098a005. [DOI] [Google Scholar]
- a Qin K.; Incarvito C. D.; Rheingold A. L.; Theopold K. H. Hydrogen Atom Abstraction by a Chromium(IV) Oxo Complex Derived from O2. J. Am. Chem. Soc. 2002, 124, 14008–14009. 10.1021/ja028382r. [DOI] [PubMed] [Google Scholar]; b Qin K.; Incarvito C. D.; Rheingold A. L.; Theopold K. H. A Structurally Characterized Chromium(III) Superoxide Complex Features “Side-on” Bonding. Angew. Chem., Int. Ed. 2002, 41, 2333–2335. . [DOI] [PubMed] [Google Scholar]; c Akturk E. S.; Yap G. P. A.; Theopold K. H. Mechanism-based design of labile precursors for chromium(I) chemistry. Chem. Commun. 2015, 51, 15402–15405. 10.1039/C5CC05993C. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Akturk E. S.; Yap G. P. A.; Theopold K. H. Dioxygen Activation by Non-Adiabatic Oxidative Addition to a Single Metal Center. Angew. Chem., Int. Ed. 2015, 54, 14974–14977. 10.1002/anie.201508777. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Hess A.; Hörz M. R.; Liable-Sands L. M.; Lindner D. C.; Rheingold A. L.; Theopold K. H. Insertion of O2 into a Chromium–Phenyl Bond: Mechanism of Formation of the Paramagnetic d2 Oxo Complex [TptBu,MeCrIV(O)OPh]. Angew. Chem., Int. Ed. 1999, 38, 166–168. . [DOI] [Google Scholar]
- Addison A. W.; Rao T. N.; Reedijk J.; van Rijn J.; Verschoor G. C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349–1356. 10.1039/DT9840001349. [DOI] [Google Scholar]
- Biagini P.; Calderazzo F.; Marchetti F.; Romano A. M.; Spera S. Synthesis and structural characterization of sterically crowded hydridotris(pyrazolyl)borato complexes: Unusual double 1,2-borotropic shift at a titanium centre. J. Organomet. Chem. 2006, 691, 4172–4180. 10.1016/j.jorganchem.2006.06.026. [DOI] [Google Scholar]
- a Mailer C.; Taylor C. P. S. Rapid adiabatic passage EPR of ferricytochrome c: Signal enhancement and determination of the spin-lattice relaxation time. Biochim. Biophys. Acta, Protein Struct. 1973, 322, 195–203. 10.1016/0005-2795(73)90293-6. [DOI] [PubMed] [Google Scholar]; b Mailer C.; Hoffman B. M. Tumbling of an adsorbed nitroxide using rapid adiabatic passage. J. Phys. Chem. 1976, 80, 842–846. 10.1021/j100549a015. [DOI] [Google Scholar]
- Aranzaes J. R.; Daniel M.-C.; Astruc D. Metallocenes as references for the determination of redox potentials by cyclic voltammetry - Permethylated iron and cobalt sandwich complexes, inhibition by polyamine dendrimers, and the role of hydroxy-containing ferrocenes. Can. J. Chem. 2006, 84, 288–299. 10.1139/v05-262. [DOI] [Google Scholar]
- a Yang L.; Powell D. R.; Houser R. P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 955–964. 10.1039/B617136B. [DOI] [PubMed] [Google Scholar]; b Reineke M. H.; Sampson M. D.; Rheingold A. L.; Kubiak C. P. Synthesis and Structural Studies of Nickel(0) Tetracarbene Complexes with the Introduction of a New Four-Coordinate Geometric Index, τδ. Inorg. Chem. 2015, 54, 3211–3217. 10.1021/ic502792q. [DOI] [PubMed] [Google Scholar]
- Krzystek J.; Zvyagin S. A.; Ozarowski A.; Trofimenko S.; Telser J. Tunable-frequency high-field electron paramagnetic resonance. J. Magn. Reson. 2006, 178, 174–183. 10.1016/j.jmr.2005.09.007. [DOI] [PubMed] [Google Scholar]
- a Angeli C.; Cimiraglia R.; Evangelisti S.; Leininger T.; Malrieu J.-P. Introduction of n-electron valence states for multireference perturbation theory. J. Chem. Phys. 2001, 114, 10252–10264. 10.1063/1.1361246. [DOI] [Google Scholar]; b Angeli C.; Cimiraglia R.; Malrieu J.-P. N-electron valence state perturbation theory: a fast implementation of the strongly contracted variant. Chem. Phys. Lett. 2001, 350, 297–305. 10.1016/S0009-2614(01)01303-3. [DOI] [Google Scholar]; c Angeli C.; Cimiraglia R.; Malrieu J.-P. n-electron valence state perturbation theory: A spinless formulation and an efficient implementation of the strongly contracted and of the partially contracted variants. J. Chem. Phys. 2002, 117, 9138–9153. 10.1063/1.1515317. [DOI] [Google Scholar]
- a Liakos D. G.; Ganyushin D.; Neese F. A Multiconfigurational ab Initio Study of the Zero-Field Splitting in the Di- and Trivalent Hexaquo–Chromium Complexes. Inorg. Chem. 2009, 48, 10572–10580. 10.1021/ic901063y. [DOI] [PubMed] [Google Scholar]; b Sinnecker S.; Neese F. Spin–Spin Contributions to the Zero-Field Splitting Tensor in Organic Triplets, Carbenes and Biradicals – A Density Functional and Ab Initio Study. J. Phys. Chem. A 2006, 110, 12267–12275. 10.1021/jp0643303. [DOI] [PubMed] [Google Scholar]; c Duboc C.; Ganyushin D.; Sivalingam K.; Collomb M.-N.; Neese F. Systematic Theoretical Study of the Zero-Field Splitting in Coordination Complexes of Mn(III). Density Functional Theory versus Multireference Wave Function Approaches. J. Phys. Chem. A 2010, 114, 10750–10758. 10.1021/jp107823s. [DOI] [PubMed] [Google Scholar]; d Neese F. Importance of Direct Spin–Spin Coupling and Spin-Flip Excitations for the Zero-Field Splittings of Transition Metal Complexes: A Case Study. J. Am. Chem. Soc. 2006, 128, 10213–10222. 10.1021/ja061798a. [DOI] [PubMed] [Google Scholar]
- a Lang L.; Neese F. Spin-dependent properties in the framework of the dynamic correlation dressed complete active space method. J. Chem. Phys. 2019, 150, 104104. 10.1063/1.5085203. [DOI] [PubMed] [Google Scholar]; b Maurice R.; Bastardis R.; Graaf C. d.; Suaud N.; Mallah T.; Guihéry N. Universal Theoretical Approach to Extract Anisotropic Spin Hamiltonians. J. Chem. Theory Comput. 2009, 5, 2977–2984. 10.1021/ct900326e. [DOI] [PubMed] [Google Scholar]
- a Sanner R. D.; Duggan D. M.; McKenzie T. C.; Marsh R. E.; Bercaw J. E. Structure and magnetism of μ-Dinitrogen-bis(bis(pentamethylcyclopentadienyl)titanium(II)), {(η5-C5(CH3)5)2Ti}2N2. J. Am. Chem. Soc. 1976, 98, 8358–8365. 10.1021/ja00442a008. [DOI] [Google Scholar]; b Berry D. H.; Procopio L. J.; Carroll P. J. Molecular Structure of {Cp2Ti(PMe3)}2(μ-N2), a Titanocene Dinitrogen Complex. Organometallics 1988, 7, 570–572. 10.1021/om00092a059. [DOI] [Google Scholar]; c Duchateau R.; Gambarotta S.; Beydoun N.; Bensimon C. , Side-on versus end-on coordination of dinitrogen to titanium(II) and mixed-valence titanium(I)/titanium(II) amido complexes. J. Am. Chem. Soc. 1991, 113, 8986–8988. 10.1021/ja00023a080. [DOI] [Google Scholar]; d Beydoun N.; Duchateau R.; Gambarotta S. , Synthesis and characterization of a thermally robust titanium dinitrogen complex. The crystal structure of [{[(Me3Si)2N]TiCl(pyridine)2}2(μ-η:η′-N2)]. J. Chem. Soc., Chem. Commun. 1992, 244–246. 10.1039/C39920000244. [DOI] [Google Scholar]; e Hanna T. E.; Lobkovsky E.; Chirik P. J. , Dinitrogen Activation by Titanium Sandwich Complexes. J. Am. Chem. Soc. 2004, 126, 14688–14689. 10.1021/ja045884r. [DOI] [PubMed] [Google Scholar]; f Hanna T. E.; Lobkovsky E.; Chirik P. J. , Dinitrogen Complexes of Bis(cyclopentadienyl) Titanium Derivatives: Structural Diversity Arising from Substituent Manipulation. Organometallics 2009, 28, 4079–4088. 10.1021/om900282u. [DOI] [Google Scholar]
- a Morello L.; Yu P.; Carmichael C. D.; Patrick B. O.; Fryzuk M. D. Formation of Phosphorus–Nitrogen Bonds by Reduction of a Titanium Phosphine Complex under Molecular Nitrogen. J. Am. Chem. Soc. 2005, 127, 12796–12797. 10.1021/ja054467r. [DOI] [PubMed] [Google Scholar]; b Shima T.; Hu S.; Luo G.; Kang X.; Luo Y.; Hou Z. Dinitrogen Cleavage and Hydrogenation by a Trinuclear Titanium Polyhydride Complex. Science 2013, 340, 1549–1552. 10.1126/science.1238663. [DOI] [PubMed] [Google Scholar]; c Mo Z.; Shima T.; Hou Z. Synthesis and Diverse Transformations of a Dinitrogen Dititanium Hydride Complex Bearing Rigid Acridane-Based PNP-Pincer Ligands. Angew. Chem., Int. Ed. 2020, 59, 8635–8644. 10.1002/anie.201916171. [DOI] [PubMed] [Google Scholar]
- Huber K. P.; Herzberg G.. Molecular Spectra and Molecular Structure—IV. Constants of Diatomic Molecules; Van Nostrand Reinhold: New York, 1979. [Google Scholar]
- a Durig J. R.; Griffin M. G.; Macnamee R. W. Raman spectra of gases. XV: Hydrazine and hydrazine-d4. J. Raman Spectrosc. 1975, 3, 133–141. 10.1002/jrs.1250030204. [DOI] [Google Scholar]; b Craig N. C.; Levin I. W. Vibrational assignment and potential function for trans-diazene (diimide): Predictions for cis-diazene. J. Chem. Phys. 1979, 71, 400–407. 10.1063/1.438084. [DOI] [Google Scholar]
- a Figueroa J. S.; Piro N. A.; Clough C. R.; Cummins C. C. A Nitridoniobium(V) Reagent That Effects Acid Chloride to Organic Nitrile Conversion: Synthesis via Heterodinuclear (Nb/Mo) Dinitrogen Cleavage, Mechanistic Insights, and Recycling. J. Am. Chem. Soc. 2006, 128, 940–950. 10.1021/ja056408j. [DOI] [PubMed] [Google Scholar]; b Gdula R. L.; Johnson M. J. A. Highly Active Molybdenum–Alkylidyne Catalysts for Alkyne Metathesis: Synthesis from the Nitrides by Metathesis with Alkynes. J. Am. Chem. Soc. 2006, 128, 9614–9615. 10.1021/ja058036k. [DOI] [PubMed] [Google Scholar]; c Bailey B. C.; Fout A. R.; Fan H.; Tomaszewski J.; Huffman J. C.; Gary J. B.; Johnson M. J. A.; Mindiola D. J. Snapshots of an Alkylidyne for Nitride Triple-Bond Metathesis. J. Am. Chem. Soc. 2007, 129, 2234–2235. 10.1021/ja0689684. [DOI] [PubMed] [Google Scholar]
- Using the following convention for defining the coupling constant: J = (HSE – BSE)/(HS⟨S2⟩ – BS⟨S2⟩).; a Yamaguchi K.; Takahara Y.; Fueno T.. Ab-Initio Molecular Orbital Studies of Structure and Reactivity of Transition Metal-OXO Compounds; Springer: Dordrecht, The Netherlands, 1986; pp 155–184. [Google Scholar]; b Soda T.; Kitagawa Y.; Onishi T.; Takano Y.; Shigeta Y.; Nagao H.; Yoshioka Y.; Yamaguchi K. Ab initio computations of effective exchange integrals for H–H, H–He–H and Mn2O2 complex: comparison of broken-symmetry approaches. Chem. Phys. Lett. 2000, 319, 223–230. 10.1016/S0009-2614(00)00166-4. [DOI] [Google Scholar]
- Kool L. B.; Rausch M. D.; Alt H. G.; Herberhold M.; Hill A. F.; Thewalt U.; Wolf B. A diazoalkane complex of titanium. J. Chem. Soc., Chem. Commun. 1986, 408–409. 10.1039/c39860000408. [DOI] [Google Scholar]
- a Polse J. L.; Andersen R. A.; Bergman R. G. Synthesis, Structure, and Reactivity Studies of an η2-N2-Titanium Diazoalkane Complex. Generation and Trapping of a Carbene Complex Intermediate. J. Am. Chem. Soc. 1996, 118, 8737–8738. 10.1021/ja9614981. [DOI] [Google Scholar]; b Polse J. L.; Kaplan A. W.; Andersen R. A.; Bergman R. G. Synthesis of an η2-N2-Titanium Diazoalkane Complex with Both Imido- and Metal Carbene-Like Reactivity Patterns. J. Am. Chem. Soc. 1998, 120, 6316–6328. 10.1021/ja974303d. [DOI] [Google Scholar]; c Hanna T. E.; Keresztes I.; Lobkovsky E.; Bernskoetter W. H.; Chirik P. J. Synthesis of a Base-Free Titanium Imido and a Transient Alkylidene from a Titanocene Dinitrogen Complex. Studies on Ti=NR Hydrogenation, Nitrene Group Transfer, and Comparison of 1,2-Addition Rates. Organometallics 2004, 23, 3448–3458. 10.1021/om049817h. [DOI] [Google Scholar]; d Kaplan A. W.; Polse J. L.; Ball G. E.; Andersen R. A.; Bergman R. G. Synthesis, Structure, and Reactivity of η2-N2-Aryldiazoalkane Titanium Complexes: Cleavage of the N–N Bond. J. Am. Chem. Soc. 1998, 120, 11649–11662. 10.1021/ja981340b. [DOI] [Google Scholar]
- Anandha Babu G.; Perumal Ramasamy R.; Ramasamy P.; Natarajan S. Studies on the crystal growth, crystal structure, optical and thermal properties of an organic crystal: Benzophenone hydrazone. J. Cryst. Growth 2009, 311, 3461–3465. 10.1016/j.jcrysgro.2009.04.007. [DOI] [Google Scholar]
- a Davis P. J.; Harris L.; Karim A.; Thompson A. L.; Gilpin M.; Moloney M. G.; Pound M. J.; Thompson C. Substituted diaryldiazomethanes and diazofluorenes: structure, reactivity and stability. Tetrahedron Lett. 2011, 52, 1553–1556. 10.1016/j.tetlet.2011.01.116. [DOI] [Google Scholar]; b Nazran A. S.; Lee F. L.; Gabe E. J.; Lepage Y.; Northcott D. J.; Park J. M.; Griller D. Structures of dimesitylcarbene and related compounds. J. Phys. Chem. 1984, 88, 5251–5254. 10.1021/j150666a028. [DOI] [Google Scholar]
- a Armor J. N.; Taube H. Reduction of nitrous oxide in the presence of pentaammineaquoruthenium(II). J. Am. Chem. Soc. 1971, 93, 6476–6480. 10.1021/ja00753a023. [DOI] [Google Scholar]; b Laplaza C. E.; Odom A. L.; Davis W. M.; Cummins C. C.; Protasiewicz J. D. Cleavage of the Nitrous Oxide NN Bond by a Tris(amido)molybdenum(III) Complex. J. Am. Chem. Soc. 1995, 117, 4999–5000. 10.1021/ja00122a033. [DOI] [Google Scholar]; c Cherry J.-P. F.; Johnson A. R.; Baraldo L. M.; Tsai Y.-C.; Cummins C. C.; Kryatov S. V.; Rybak-Akimova E. V.; Capps K. B.; Hoff C. D.; Haar C. M.; Nolan S. P. On the Origin of Selective Nitrous Oxide N–N Bond Cleavage by Three-Coordinate Molybdenum(III) Complexes. J. Am. Chem. Soc. 2001, 123, 7271–7286. 10.1021/ja0031063. [DOI] [PubMed] [Google Scholar]; d Palluccio T. D.; Rybak-Akimova E. V.; Majumdar S.; Cai X.; Chui M.; Temprado M.; Silvia J. S.; Cozzolino A. F.; Tofan D.; Velian A.; Cummins C. C.; Captain B.; Hoff C. D. Thermodynamic and Kinetic Study of Cleavage of the N–O Bond of N-Oxides by a Vanadium(III) Complex: Enhanced Oxygen Atom Transfer Reaction Rates for Adducts of Nitrous Oxide and Mesityl Nitrile Oxide. J. Am. Chem. Soc. 2013, 135, 11357–11372. 10.1021/ja405395z. [DOI] [PubMed] [Google Scholar]
- a Pamplin C. B.; Ma E. S. F.; Safari N.; Rettig S. J.; James B. R. The Nitrous Oxide Complex, RuCl2(η1-N2O)(P–N)(PPh3) (P–N = [o-(N,N-Dimethylamino)phenyl]diphenylphosphine); Low Temperature Conversion of N2O to N2 and O2. J. Am. Chem. Soc. 2001, 123, 8596–8597. 10.1021/ja0106319. [DOI] [PubMed] [Google Scholar]; b Piro N. A.; Lichterman M. F.; Harman W. H.; Chang C. J. A Structurally Characterized Nitrous Oxide Complex of Vanadium. J. Am. Chem. Soc. 2011, 133, 2108–2111. 10.1021/ja110798w. [DOI] [PubMed] [Google Scholar]; c Zhuravlev V.; Malinowski P. J. A Stable Crystalline Copper(I)–N2O Complex Stabilized as the Salt of a Weakly Coordinating Anion. Angew. Chem., Int. Ed. 2018, 57, 11697–11700. 10.1002/anie.201806836. [DOI] [PubMed] [Google Scholar]; d Gyton M. R.; Leforestier B.; Chaplin A. B. Rhodium(I) Pincer Complexes of Nitrous Oxide. Angew. Chem., Int. Ed. 2019, 58, 15295–15298. 10.1002/anie.201908333. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Mokhtarzadeh C. C.; Chan C.; Moore C. E.; Rheingold A. L.; Figueroa J. S. Side-On Coordination of Nitrous Oxide to a Mononuclear Cobalt Center. J. Am. Chem. Soc. 2019, 141, 15003–15007. 10.1021/jacs.9b08241. [DOI] [PubMed] [Google Scholar]
- a Graham T. W.; Kickham J.; Courtenay S.; Wei P.; Stephan D. W. Reduction of Titanium(IV)-Phosphinimide Complexes: Routes to Ti(III) Dimers, Ti(IV)-Metallacycles, and Ti(II) Species. Organometallics 2004, 23, 3309–3318. 10.1021/om049826q. [DOI] [Google Scholar]; b Boynton J. N.; Guo J.-D.; Grandjean F.; Fettinger J. C.; Nagase S.; Long G. J.; Power P. P. Synthesis and Characterization of the Titanium Bisamide Ti{N(H)AriPr6}2 (AriPr6 = C6H3-2,6-(C6H2-2,4,6-iPr3)2 and Its TiCl{N(H)AriPr6}2 Precursor: Ti(II) → Ti(IV) Cyclization. Inorg. Chem. 2013, 52, 14216–14223. 10.1021/ic4021355. [DOI] [PubMed] [Google Scholar]; c Aguilar-Calderón J. R.; Murillo J.; Gomez-Torres A.; Saucedo C.; Jordan A.; Metta-Magaña A. J.; Pink M.; Fortier S. Redox Character and Small Molecule Reactivity of a Masked Titanium(II) Synthon. Organometallics 2020, 39, 295–311. 10.1021/acs.organomet.9b00637. [DOI] [Google Scholar]; d Gómez-Torres A.; Aguilar-Calderón J. R.; Encerrado-Manriquez A. M.; Pink M.; Metta-Magaña A. J.; Lee W.-Y.; Fortier S. Titanium-Mediated Catalytic Hydrogenation of Monocyclic and Polycyclic Arenes. Chem. - Eur. J. 2020, 26, 2803–2807. 10.1002/chem.201905466. [DOI] [PubMed] [Google Scholar]; e Gómez-Torres A.; Aguilar-Calderón J. R.; Saucedo C.; Jordan A.; Metta-Magaña A.; Pinter B.; Fortier S. Reversible oxidative-addition and reductive-elimination of thiophene from a titanium complex and its thermally-induced hydrodesulphurization chemistry. Chem. Commun. 2020, 56, 1545–1548. 10.1039/C9CC09267F. [DOI] [PubMed] [Google Scholar]
- a Carpino L. A.; Padykula R. E.; Barr D. E.; Hall F. H.; Krause J. G.; Dufresne R. F.; Thoman C. J. Synthesis, Characterization, and Thermolysis of 7-Amino-7-azabenzonorbornadienes. J. Org. Chem. 1988, 53, 2565–2572. 10.1021/jo00246a031. [DOI] [Google Scholar]; b Mindiola D. J.; Cummins C. C. Deprotonated 2,3:5,6-Dibenzo-7-aza bicyclo[2.2.1]hepta-2,5-diene as a Nitrido Nitrogen Source by Anthracene Elimination: Synthesis of an Iodide(nitride)chromium(VI) Complex. Angew. Chem., Int. Ed. 1998, 37, 945–947. . [DOI] [PubMed] [Google Scholar]
- a McKarns P. J.; Yap G. P. A.; Rheingold A. L.; Winter C. H. Synthesis, Structure, and Characterization of the Hydrogen-Substituted Imido Complex TiCl2(NH)(OPPh3)2. Inorg. Chem. 1996, 35, 5968–5969. 10.1021/ic960595s. [DOI] [Google Scholar]; b Tran B. L.; Washington M. P.; Henckel D. A.; Gao X.; Park H.; Pink M.; Mindiola D. J. A four coordinate parent imide via a titanium nitridyl. Chem. Commun. 2012, 48, 1529–1531. 10.1039/C1CC14574F. [DOI] [PubMed] [Google Scholar]; c Thompson R.; Chen C.-H.; Pink M.; Wu G.; Mindiola D. J. A Nitrido Salt Reagent of Titanium. J. Am. Chem. Soc. 2014, 136, 8197–8200. 10.1021/ja504020t. [DOI] [PubMed] [Google Scholar]; d Grant L. N.; Pinter B.; Kurogi T.; Carroll M. E.; Wu G.; Manor B. C.; Carroll P. J.; Mindiola D. J. Molecular titanium nitrides: nucleophiles unleashed. Chem. Sci. 2017, 8, 1209–1224. 10.1039/C6SC03422E. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Grant L. N.; Pinter B.; Gu J.; Mindiola D. J. Molecular Zirconium Nitride Super Base from a Mononuclear Parent Imide. J. Am. Chem. Soc. 2018, 140, 17399–17403. 10.1021/jacs.8b11198. [DOI] [PubMed] [Google Scholar]
- a Cummins C. C.; Schrock R. R.; Davis W. M. Synthesis of Terminal Vanadium(V) Imido, Oxo, Sulfido, Selenido, and Tellurido Complexes by Imido Group or Chalcogenide Atom Transfer to Trigonal Monopyramidal V[N3N] (N3N = [(Me3SiNCH2CH2)3N]3-). Inorg. Chem. 1994, 33, 1448–1457. 10.1021/ic00085a038. [DOI] [Google Scholar]; b Smythe N. C.; Schrock R. R.; Müller P.; Weare W. W. Synthesis of [(HIPTNCH2CH2)3N]V Compounds (HIPT = 3,5-(2,4,6-i-Pr3C6H2)2C6H3) and an Evaluation of Vanadium for the Reduction of Dinitrogen to Ammonia. Inorg. Chem. 2006, 45, 9197–9205. 10.1021/ic061554r. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Hulley E. B.; Bonanno J. B.; Wolczanski P. T.; Cundari T. R.; Lobkovsky E. B. Pnictogen-Hydride Activation by (silox)3Ta (silox = tBu3SiO); Attempts to Circumvent the Constraints of Orbital Symmetry in N2 Activation. Inorg. Chem. 2010, 49, 8524–8544. 10.1021/ic101147x. [DOI] [PubMed] [Google Scholar]
- a Hanna T. E.; Lobkovsky E.; Chirik P. J. N–H Group Transfer and Oxidative Addition Chemistry Promoted by Isolable Bis(cyclopentadienyl)titanium Sandwich Complexes. Eur. J. Inorg. Chem. 2007, 2007, 2677–2685. 10.1002/ejic.200601134. [DOI] [Google Scholar]; b Veige A. S.; Slaughter L. M.; Lobkovsky E. B.; Wolczanski P. T.; Matsunaga N.; Decker S. A.; Cundari T. R. Symmetry and Geometry Considerations of Atom Transfer: Deoxygenation of (silox)3WNO and R3PO (R = Me, Ph, tBu) by (silox)3M (M = V, NbL (L = PMe3, 4-Picoline), Ta; silox = tBu3SiO). Inorg. Chem. 2003, 42, 6204–6224. 10.1021/ic0300114. [DOI] [PubMed] [Google Scholar]; c Bowman A. C.; Bart S. C.; Heinemann F. W.; Meyer K.; Chirik P. J. Synthesis of Bis(imino)pyridine Iron Amide and Ammonia Compounds from an N–H Transfer Agent. Inorg. Chem. 2009, 48, 5587–5589. 10.1021/ic9003017. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








