Open Access Article
Andreas A.
Danopoulos
*ab,
Pierre
Braunstein
*b,
Kirill Yu.
Monakhov
*c,
Jan
van Leusen
c,
Paul
Kögerler
*cd,
Martin
Clémancey
e,
Jean-Marc
Latour
e,
Anass
Benayad
f,
Moniek
Tromp
g,
Elixabete
Rezabal
h and
Gilles
Frison
h
aInstitute for Advanced Study (USIAS), Université de Strasbourg, 67081 Strasbourg Cedex, France. E-mail: danopoulos@unistra.fr
bUniversité de Strasbourg, CNRS, CHIMIE UMR 7177, Laboratoire de Chimie de Coordination, Institut de Chimie, 4 rue Blaise Pascal, 67081 Strasbourg Cedex, France. E-mail: braunstein@unistra.fr
cInstitut für Anorganische Chemie, RWTH Aachen University, 52074 Aachen, Germany. E-mail: kirill.monakhov@ac.rwth-aachen.de; paul.koegerler@ac.rwth-aachen.de
dJülich-Aachen Research Alliance (JARA-FIT) and Peter Grünberg Institute 6, Forschungszentrum Jülich, 52425 Jülich, Germany
eLaboratoire de Chimie et Biologie des Métaux, Equipe de Physicochimie des Métaux en Biologie, UMR 5249 CNRS/CEA-DRF-BIG/Université Grenoble-Alpes, 17 rue des Martyrs, Grenoble 38054, France
fCEA/DRT/LITEN/DTNM/SEN/L2N, 38054 Grenoble Cedex 9, France
gVan't Hoff Institute for Molecular Sciences, Sustainable Materials Characterisation, University of Amsterdam, Amsterdam, The Netherlands
hLCM, CNRS, Ecole Polytechnique, Université Paris-Saclay, 91128 Palaiseau, France
First published on 21st December 2016
The linear, two-coordinate and isostructural heteroleptic [M(IPr){N(SiMe3)2}] (IPr = 1,3-bis(diisopropylphenyl)-imidazol-2-ylidene), formally MI complexes (M = Co, 3; Fe, 4) were obtained by the reduction of [M(IPr)Cl{N(SiMe3)2}] with KC8, or [Co(IPr){N(SiMe3)2}2] with mes*PH2, mes* = 2,4,6-tBu3C6H2. The magnetism of 3 and 4 implies CoII and FeII centres coupled to one ligand-delocalized electron, in line with XPS and XANES data; the ac susceptibility of 4 detected a pronounced frequency dependence due to slow magnetization relaxation. Reduction of [Fe(IPr)Cl{N(SiMe3)2}] with excess KC8 in toluene gave the heteronuclear ‘inverse-sandwich’ Fe–K complex 7, featuring η6-toluene sandwiched between one Fe0 and one K+ centre.
With the aim to rationally access linear two-coordinate complexes with particularly interesting magnetic properties,5 bulky ligands have been used. In general, departure from the exactly linear geometry due to intra- or inter-molecular interactions or other ligand effects has a deleterious effect on the desirably high unquenched orbital magnetism. Thus, alkyl (i.e. C(SiMe3)3e.g. in [M{C(SiMe3)3}2]−, M = Fe,6,7 M = Mn8) or amido ligands (i.e. –N(SiMemPh(3−m))2, m = 1, 2; N(SiMe3)(DiPP), DiPP = 2,6-diisopropylphenyl), have been employed successfully to support linear homoleptic 2-coordinate complexes of Fe,5a–c,3c,9 Co5d,9 and Ni5d,4a and to a lesser extent Cr,9d,10 Mn9b,10b and V;11 two-coordinate linear or quasi-linear homoleptic terphenyls, alkoxides and thiolates have also been described.1,12 Neutral homoleptic complexes with the ubiquitous –N(SiMe3)2 form dimers in the solid state or in solution through amido bridging, due to the insufficient steric bulk of the ligand but the anion [Fe{N(SiMe3)2}2]− (ref. 13) is mononuclear and linear. Lately, mononuclear, 2-coordinate homoleptic complexes of Fe,4g,14 Co,14b,15 Mn16 and Cr17 with the cAAC (cAAC = cyclic AlkylAmino Carbene), and cationic complexes with imidazol-2-ylidene ligands18 have also been reported.
Notwithstanding these synthetic successes, there is only a limited number of linear heteroleptic 2-coordinate complexes A,19B,4a,c,e,f,20C
21 and D
22 with 3dn (n < 10) metals (Scheme 1). Heteroleptic 3-coordinate NHC amido species have recently been appearing more often.23
Herein, we describe the stable, formally 12 and 11 valence electron, virtually linear, 2-coordinate, charge-neutral heteroleptic complexes [M(IPr){N(SiMe3)2}] (IPr = N,N′-bis(di-isopropylphenyl)imidazol-2-ylidene, M = Co (3), Fe (4)), respectively, and their remarkable magnetic behaviour; we propose an original, functional model for the interpretation of the latter, supported by probing the oxidation state of the metals with XPS and XANES techniques. We also include some preliminary reactivity studies of 3 and 4.
Interestingly, reduction of [Co(IMes){N(SiMe3)2}Cl] (IMes = 1,3-bis(mesityl)-imidazol-2-ylidene) with KC8 and of [Co(IMes){N(SiMe3)2}2] with mes*PH2 led to 3-coordinate [Co(IMes)2Cl]25 and intractable mixtures, respectively. Reaction of Co–NHC bis(trimethylsilyl)amide complexes with mesPH2 (mes = 2,4,6-Me3C6H2) yielded Co–NHC phosphinidene species.26
Reduction of [Fe(IPr){N(SiMe3)2}Cl] (2Cl) with excess KC8 in pentane gave 4 in moderate yields (Scheme 2), while reaction of [Fe(IPr){N(SiMe3)2}2] (2N) with mes*PH2 yielded only [Fe(a-IPr){N(SiMe3)2}2] (a-IPr = abnormally-bound IPr)27 and no 5. Both 3 and 4 are paramagnetic with shifted but relatively sharp lines observable in the 1H-NMR spectra, which could be assigned by integration.
Determination of the structure of 3 and 4 by X-ray diffraction (Fig. 1)‡ showed that the complexes are isostructural and virtually linear at metal (178.83(7)° and 178.2(2)°, respectively); the M–N(SiMe3)2 and the M–CNHC distances fall into the generally established ranges for these bond types.13,21,23d Attempts were undertaken to use metrical data in order to support metal oxidation state assignments in 3 and 4, which can formally be described as comprising either MI centres coordinated to one anionic amido and one neutral IPr ligand, or MII centres coordinated to one anionic amido and one radical anionic IPr. However, detailed comparison of the M–N(SiMe3)2 and M–CNHC bond distances in 3 and 4 with those in known two-coordinate homoleptic complexes, as well as other two-coordinate analogues (Scheme 1) was inconclusive. Correlation of the metrical data with a metal oxidation state was hampered by the scarcity of relevant identically substituted, two-coordinate complexes in different oxidation states, and the disparate trans-influence of the NHC and amido donor types. In the known homoleptic [Fe{N(SiMe3)2}2]−, [Fe(cyIDep)2]+ and [Fe(cAAC)2]+, cyIDep = 1,3-bis(2′,6′-diethylphenyl)-4,5-(CH2)4-imidazol-2-ylidene (all assigned as FeI complexes), the distances FeI–N = 1.9213(6) Å and FeI–CNHC = 1.971(5)–1.996(7) and 1.997(3) Å, respectively, are shorter than the corresponding in 4.4g,13,18b In addition, the M–N and M–CNHC distances in 3 and 4 are significantly shorter than those in the three-coordinate [MII(IPr)(N(SiMe3)2)2] (M = Fe, Co).23a,d,e However, the Fe–CNHC distance in 4 is virtually identical to that in the heteroleptic C (Scheme 1), in which FeI was implicated. In both 3 and 4 the heterocyclic rings are virtually planar (max. displacement from the mean 5-membered ring plane 0.002 Å); planar is also the environment of the amido N-atoms. In view of the lack of structural or computational data on plausible imidazole-2-ylidene radical anions28 and the previous discussion, it is futile to argue for metrical (ligand) oxidation states29 in 3 and 4. Close in energy, open-shell electronic structures may be attainable under specific conditions (temperature, solvent etc.) by fine ligand tuning, which may also potentially lead to ligand(s) noninnocence. Lastly, in the crystals of 3 and 4 there are no close contacts of the molecules with nearby atoms, implying sufficient electronic stabilization of the two-coordinate structures and the minimal role of sterics and/or dispersion forces serving this purpose. In the crystalline solids the metal centres are at ca. 9.485 and 9.471 Å apart, respectively (self-dilute).
![]() | ||
| Fig. 1 The structures of 3 (top) and 4 (bottom); H-atoms are omitted. Important bond lengths (Å) and angles (°): for 3, C1–Co1: 1.9423(18), C1–N2: 1.368(2), C1–N1: 1.368(2), C2–C3: 1.343(3), C2–N1: 1.381(2), N3–Co1–C1: 178.83(7). For 4, C1–N2: 1.366(7), C1–N1: 1.368(7), C1–Fe1: 2.015(6), C2–C3: 1.336(10), C2–N1: 1.388(8), N3–Fe1: 1.882(5), N3–Fe1: 1.882(5), N3–Fe1–C1: 178.2(2); additional details are given in the ESI.‡ | ||
To account for the observed values, we reasoned that 3 may be characterized by either a 3d74s1 configuration in a quintet state (S = 2), or by a configuration resulting from the transfer of one electron from the CoI to the ligand periphery, i.e. localized at one of the ligands or delocalized over the π system of the complex. Analogous considerations for 4 may lead to a 3d64s1 configuration in a sextet state (S = 5/2), or one electron transferred from FeI to one of the ligands, more likely the NHC. Consequently, three different electronic configuration scenarios have been investigated: (i) a single MI centre (3dN, M = CoI: N = 8; M = FeI: N = 7); (ii) a single MII centre (3dN−1) representing the upper limit for a high spin 3d6 electron configuration, with free ion values: L = 2, S = 2, J = 4, gJ = 3/2, μeff ≈ 6.71μB, to explore the unlikely possibility of decomposition; (iii) a single MII (3dN−1) centre interacting with one electron (S = 1/2) via Heisenberg–Dirac–van Vleck exchange coupling that is (de)localized over the ligand/complex (referred hereafter to as the ‘MII + e−’ scenario). In the latter case, maximum μeff ≈ 6.86μB, and μeff ≈ 6.93μB corresponding to “3d7 + e−”and “3d6 + e−”, respectively, could be envisaged.
Considering the C∞v ligand field symmetry to be in line with the linear heteroleptic nature of 3 and 4, we employed the ‘full’ model Hamiltonian implemented in CONDON 2.0.30 The corresponding Hamiltonian, neglecting contributions resulting in constant shifts of the total energy, is defined as:
in Ĥlf denote the relevant spherical tensors for a given ligand field symmetry and are directly related to the spherical harmonics Y0k. B0k are the (real) ligand field parameters in the Wybourne notation. The sum index i runs over all NM valence electrons of the corresponding metal center.
The least-squares fits for the scenarios (i) and (ii) did not yield even remotely acceptable solutions (including considerations of free ions, and physically unlikely parameters, see Fig. S8‡ for a selection of resulting fits). In contrast, the “MII + e−” scenario indeed reproduced the temperature-dependent susceptibility data (SQ = 1.0%; Fig. 2 and Table S6‡). We note that for 3, inclusion of the field-dependent magnetization data at 2.0 K (Fig. 2a, inset) reduces the overall fit quality (SQ = 7.7%). Fitting solely the Mmvs. B data would point (ii) as the preferred scenario, but then the χmT vs. T curve would not be reproduced at all (Fig. S8c‡). By employing the “MII + e−” scenario for 4, the least-squares fit yields a reasonable SQ = 2.8% (1.8% when excluding Mmvs. B data). The single ion contribution of the MII centre is highlighted in Fig. 2 for both complexes.
The scenario “MII + e−” yields small, antiferromagnetic exchange energies J = −0.1 cm−1 (3) and −0.5 cm−1 (4). These results indicate that the additional electron should be assigned to the ligands. Such parameters imply formal triplet (3) or quartet (4) ground states with respect to the whole molecule. We note that DFT calculations of 3 and 4 also support triplet and quartet electronic ground states, respectively; however, at the DFT theory level, only minimal spin delocalization on the ligand is predicted (see the ESI‡ for details).
Since the Heisenberg–Dirac–Van Vleck exchange formalism refers to localized electrons, an unpaired electron localized on a ligand atom is expected to induce a strong exchange interaction. In contrast, within the limitations of the model, the small magnitudes of J hint at (at least partially) delocalized electrons for which various effects, e.g. electron transfer, might compensate each other, yielding a small effective net value. Thus, the fits also reflect the inherent limitations of ligand field theory, where electrons localized at the metal interact with an electrostatic ligand field potential, neglecting further dynamic aspects generated by e.g. the conjugation of the π system.
The in-phase χ′m and out-of-phase χ′′m” components of the magnetic ac susceptibility data as a function of temperature (Fig. S9 (3) and S10‡ (4)) show out-of-phase signals for 4, up to ca. 15 K (<1000 Hz) but not for 3. The χ′′mvs. χ′m data (Fig. 2c) were analysed in terms of the generalized Debye expression31 (solid lines). The dependence of the magnetic relaxation time τ on T−1 is shown in Fig. 2d as the Arrhenius plot. The distribution of relaxation times α (0.04–0.17, mean value 0.12) suggests the existence of multiple relaxation pathways. Notably, the semi-logarithmic Arrhenius plot exhibits two quasi-linear segments between 3.0–5.5 K and 8.5–14 K. Fitting these to the Arrhenius expression τ = τ0·exp[Ueff/(kBT)] (attempt time τ0, effective energy barrier Ueff, Boltzmann constant kB) yields τ0 = (3.23 ± 0.03) × 10−4 s, Ueff = (0.89 ± 0.03) cm−1 for 3.0–5.5 K, and τ0 = (1.64 ± 0.80) × 10−6 s, Ueff = (31.0 ± 3.1) cm−1 for 8.5–14 K. Whereas the latter parameters are at the upper limit for typical Orbach relaxation in single ion magnets (SIM), the first is not, thus potentially describing a different process. Note that the effective energy barrier is of the same order as the exchange coupling parameter J. The first process might be thus linked to the potential exchange interaction of the Fe centre and the delocalized electron. We therefore considered two different models for fitting the entire temperature range (2.0–14 K): (a) quantum tunnelling, Orbach and Raman relaxation processes (τ−1 = B + τ0−1·exp[−Ueff/(kBT)] + C·Tn), and (b) quantum tunnelling, Orbach relaxation process and another Arrhenius-type relaxation process (τ−1 = B + τ0,1−1·exp[−Ueff,1/(kBT)] + τ0,2−1·exp[−Ueff,2/(kBT)]). Least-squares fits to (a) (Fig. 2 and S11,‡ blue dashed line) yield B = (1352 ± 152) s−1, τ0 = (4.56 ± 0.17) × 10−4 s, Ueff = (2.55 ± 0.52) cm−1, C = (2.88 ± 3.53) × 10−5 s−1 K−{n}, and n = 8.02 ± 0.53. Model (b) (Fig. 2 and S11,‡ red line) results in B = (1547 ± 87) s−1, τ0,1 = (1.01 ± 0.45) × 10−7 s, Ueff,1 = (56.6 ± 3.1) cm−1, τ0,2 = (4.03 ± 0.25) × 10−4 s, and Ueff,2 = (3.59 ± 0.50) cm−1. The fit parameters of (a) describe a system characterized by quantum tunnelling (B) and Raman (C) relaxation processes close to Kramers systems (n = 9), and a process which is not a typical SIM-type Orbach relaxation. Note that there are several other relaxation processes32 compared to the suggested exchange interaction, which are characterized by an Arrhenius law, e.g. the sum process. Model (b) essentially replaces Raman relaxation by Orbach relaxation, while the other fit parameters are almost unchanged. Due to a slightly better goodness-of-fit, the occurrence of an Orbach relaxation process typical of SIMs and fit parameters in the range for similar compounds,3a–c the SIM characteristics of 4 are better characterized by model (b) than (a), although the nature of the second Arrhenius-type process could not be fully resolved. We note that the magnetization relaxation dynamics of linear CoI compounds seem to be very sensitive to the ligand field, as was also observed by Meng et al.,18a which may also be related to the electron delocalization implied in the “MII + e−” scenario.
The structural analysis of 3 with Co K-edge EXAFS spectroscopy agreed with the crystal structure data (see ESI‡). Remarkably, and in agreement with XPS, the Co K-edge XANES revealed an overall charge of +2 on the Co, with the pre-edge features nicely supporting the linear structure. The Co K-edge XANES of 3, in comparison to some representative CoII and CoIII compounds, is given in Fig. 3 (the first derivatives of the XANES are given in Fig. S14‡). Although the XANES is known to be dependent on the metal oxidation state (charge) as well as the nature of the ligands and the coordination geometry, Fig. 3 suggests that 3 represents an overall Co2+ complex. The first pre-edge feature as seen for 3 can be assigned to pd hybridization, making the dipole forbidden s-to-d transition slightly visible, and the second pre-edge feature originates from hybridization of the Co-p and Co-d (with C-p mostly, and little mixing from N-p, as indicated by simulations (see Fig. S15 in ESI‡)). The high intensity of the second pre-edge is due to the empty character of the orbital, which is indicative of the linear Co. The simulations also suggest the channel of charge redistribution being the aromatic part (i.e. NHC) of the molecule.
The XPS spectrum of 4 is less well resolved but globally similar to that of 3. The 2p3/2 peak is unsymmetrical with a tail on the high-energy side and a broad feature, which is due to the presence of a satellite peak partly overlapping with the 2p3/2; the 2p1/2 peak behaves similarly. A deconvolution using a Gaussian/Lorentzian admixture allows distinguishing both components (Table S7‡). The energies of the 2p3/2 and 2p1/2 peaks and their satellites are consistent with a FeII centre.29,36 The small intensity of the satellite peaks does not support a high-spin Fe, however, due to the small energy difference between the 2p peaks and their satellites, the intensities of the latter are strongly dependent on the deconvolution mode and base line correction.
![]() | ||
| Fig. 4 The structure of 6; H-atoms are omitted. Important bond lengths (Å) and angles (°): C1–N1: 1.3921(17), C1–N2: 1.3965(18), C1–Co1: 1.8897(14), N3–Co1: 2.0387(12), N3–Co1′: 2.0421(12), Co1–Co1: 2.5765(4), N3–Co1–N3′: 101.70(4), C1–Co1–Co1′: 179.03(4), additional details are given in the ESI.‡ | ||
Reduction of 2Cl with excess KC8 in toluene afforded low yields of the yellow-green, paramagnetic 7 which could only be characterized crystallographically (Fig. 5).‡ It is a centrosymmetric, ‘inverse-sandwich’ tetranuclear heterometallic dimer, each monomer containing a η6-toluene sandwiched between one Fe0 (16e−) and one K+ centre; the two monomers are connected by two bridging –N(SiMe3)2 ligands ligated to the K cations.
![]() | ||
| Fig. 5 The structure of 7; H-atoms are omitted. Important bond lengths (Å) and angles (°): C1–N1: 1.368(4), C1–N2: 1.370(4), C1–Fe1: 1.988(3), C28–Fe1: 2.061(4), C29–Fe1: 2.070(4), Fe1–K1: 4.520, additional details are given in the ESI.‡ | ||
Complex 7 is a rare example of the Fe–NHC η6-arene complex;37,38 [Fe0(IPr)(diene)] species were also recently described.39 Interestingly, it could not be obtained by the reduction of 4 in toluene. It is also worth pointing out that mechanistically the formation of 7 is remarkable. The sole source of K is the reducing KC8 and of –N(SiMe3)2 the 2Cl featuring a direct Fe–N(SiMe3)2 bond. The formation of 7 may involve a consecutive insertion of the K(toluene) fragment into the Fe–N(SiMe3)2 of a transient [IPrFe0(N(SiMe3)2)]− with or without prior coordination to the metal; [IPrFe0(N(SiMe3)2)]− is a plausible initial product from the 2e− reduction of 2Cl. Fe–K assemblies have been recently studied with respect to Fe catalysed N2 activation.40
Although the experimental evidence for the nature of IPr in 3 and 4 is convincing, and DFT calculations support triplet and quartet electronic ground states, respectively, minimal spin delocalization on the ligand is predicted by this methodology (see the ESI‡ for details); this may be a consequence of the well-known limitations of this method with multi-reference structures. Therefore high-level ab initio calculations are desirable to provide a better insight into the electronic structures of 3 and 4.
CH, imid), −48.8 (4H, CH(CH3)2), −152.19 (12H, CH(CH3)2) ppm. The 1H-NMR spectrum in d8-THF remains unchanged. NMR spectroscopic analysis of the mother liquor after the second crop revealed the presence of 5 (31P δ −75.0 (d)) and a minor unidentified species, δ −60.0 (d of d, JP–H = 73.0 Hz).
CH, imid), −19.22 (4H, CH(CH3)2), −78.80 (12H, CH(CH3)2) ppm. The 1H-NMR spectrum in d8-THF remains unchanged.
Footnotes |
| † Dedicated to Professor Malcolm L. H. Green on the occasion of his 80th birthday. |
| ‡ Electronic supplementary information (ESI) available: General information on the synthetic methodology, studies of the reactivity, mechanistic considerations, crystal data and detailed list of metrical data and details of the DFT calculations, the magnetochemical measurements and the XPS, EXAFS and XANES studies. CCDC 1469263–1469265 and 1469268. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6dt03565e |
| This journal is © The Royal Society of Chemistry 2017 |