Abstract
The β-diketiminate supporting group, [ArNCRCHCRNAr]−, stabilizes low coordination number complexes. Four such complexes, where R = tert-butyl, Ar = 2,6-diisopropylphenyl, are studied: (nacnactBu)ML, where M = FeII, CoII and L = Cl, CH3. These are denoted FeCl, FeCH3, CoCl, and CoCH3 and have been previously reported and structurally characterized. The two FeII complexes (S = 2) have also been previously characterized by Mössbauer spectroscopy, but only indirect assessment of the ligand-field splitting and zero-field splitting (zfs) parameters was available. Here, EPR spectroscopy is used, both conventional field-domain for the CoII complexes (with S = 3/2) and frequency-domain, far-infrared magnetic resonance spectroscopy (FIRMS) for all four complexes. The CoII complexes were also studied by magnetometry. These studies allow accurate determination of the zfs parameters. The two FeII complexes are similar with nearly axial zfs and large magnitude zfs given by D = −37 ± 1 cm−1 for both. The two CoII complexes likewise exhibit large and nearly axial zfs, but surprisingly CoCl has positive D = +55 cm−1 while CoCH3 has negative D = −49 cm−1. Theoretical methods were used to probe the electronic structures of the four complexes, which explain the experimental spectra and the zfs parameters.
Graphical Abstract

Introduction
First row transition metal (3d) complexes with low-coordination numbers, defined here as two- or three-coordinate, have generated considerable interest in the inorganic chemistry community.1–13 In particular, the area of single molecule magnets (SMMs)14–17 has been greatly advanced by investigations of low-coordinate 3d ion complexes.18–39 The properties of SMMs depend crucially on the details of the ligand-field splitting of the orbitals and the zero-field splitting (zfs) of the magnetic sublevels,40–42 adding strong motivation to quantitatively determine these parameters.
One N-donor ligand that has been particularly effective at stabilizing three-coordinate complexes is the β-diketiminate ion,43–46 often denoted nacnac, due to its formal derivation from β-diketonate, or acac.47 An advantage of nacnac over acac is the ability to incorporate steric bulk on the imine/iminate substituents,45 to prevent formation of four- and six-coordinate transition metal complexes that are common with β-diketonates,48 e.g., (acac)2M, M = Cu, Pd; (acac)3M, M = Ti, V, Cr, Mn, Fe, Co, etc.; the CSD (version 5.43, update 4) yields 355 structures of 4-coordinate bis(acac) complexes (of which 245 are M = Cu) and 378 structures of 6-coordinate tris(acac) complexes. In contrast, there are only two structurally characterized complexes of (nacnacR,R′)3M; with M = CrIII (CSD: IJEVUP, IJEWAW)49 and with YIII (CSD: XINMAJ).50 In the former, the N-substituents are benzyl (nacnacBn,Me) and in the latter phenyl (nacnacPh,Me). Four-coordinate bis(nacnac) complexes are relatively plentiful, with 108 structures, although this includes tetraazamacrocylic complexes such as these of NiII.51 There are 48 structures of (nacacR,Me)2M transition metal complexes with two bidentate ligands. With bulky N-substituents, most commonly 2,6-diisopropylphenyl (DiPP) groups, the three-coordinate complexes of general formula (nacnacR,R′)MX are often isolable, with the CSD yielding 401 structures wherein M = Fe, Co, Ni, Cu, Zn (primarily), and X = C, N, P, O, S, halide, etc. Specifically with the DiPP substituent, there are 255 structures of (nacnacDiPP,R′)MX. Thus, this is a useful scaffold for constructing systematic series of three-coordinate complexes.
Our focus here is on two 3d ions of great importance in inorganic chemistry in general, and in the area of SMMs in particular, namely FeII and CoII. These ions feature rich d-orbital manifolds when they have high-spin d6 and d7 electronic configurations, respectively. Cobalt(II) complexes are among the most extensively studied in connection with SMM behavior, as has been very recently reviewed,52 but SMM examples of FeII, particularly in low-coordination number, are also plentiful such as in these recent examples.31, 33, 34, 36 For the detailed study here, we utilize two ligands in the third coordination position, X = Cl and CH3. These represent respectively a moderate σ- and π-donor and a strong σ-donor. Both have cylindrical symmetry (i.e., when z is chosen as the M-X σ-bond vector, the x and y directions are equivalent). These four complexes, which have been previously reported by some of us,4, 53, 54 are testbeds for investigating the electronic structure of 3-coordinate complexes of high-spin 3d ions. In these complexes, an extremely bulky β-diketiminate ligand is used, 2,2,6,6-tetramethyl-3,5-bis(2,6-diisopropylphenylimido)heptane anion, nacnacDiPP,tBu, which we will abbreviate as nacnactBu. For further simplicity, the four (nacnacDiPP,tBu)MX (M= Fe, Co; X = Cl, CH3) complexes will be referred to respectively as FeCl, FeCH3, CoCl, and CoCH3 and collectively as MX and pairwise by metal as FeX or CoX and by ligand as MCl or MCH3.
Using applied-field Mössbauer effect spectroscopy, certain aspects of the electronic structure of the FeX complexes have been investigated previously;4 however, the computational tools available at that time were insufficient for a deeper study by quantum chemical theory. Mössbauer gives an indirect assessment of the ligand-field transitions of the iron complexes, and is of course inapplicable to the cobalt complexes. Here, we use a more direct measurement of ligand-field parameters, far-infrared magnetic resonance spectroscopy (FIRMS)55, 56. FIRMS directly yields the zfs of high-spin systems, which is key towards understanding their electronic structure. The results obtained herein give insights that may be applicable to other high-spin ions and to other complexes of the versatile and popular β-diketiminate ligand platform.
Experimental Section
Synthesis.
The complexes FeCl, FeCH3, CoCl, and CoCH3 were prepared as previously reported.4, 54, 57 All sample handling was done under an inert atmosphere.
Electronic absorption spectroscopy.
Electronic absorption spectra of all complexes were recorded in toluene solution on a Cary 60 spectrophotometer.
Caution! Care should be taken in the presence of high magnetic fields and in the use of cryogenic fluids.
X- and Q-Band EPR/ENDOR spectroscopy.
X-band (~9.5 GHz) spectra of CoCl and CoCH3 in toluene frozen solution and as pure powders were recorded on a modified Bruker E109 spectrometer equipped with an Oxford cryostat. Q-band (35 GHz) EPR and ENDOR spectra only of frozen solution samples were recorded at 2 K on CW58 and pulsed59 spectrometers previously described, the latter using the Davies pulse sequence60 for 14N ENDOR. CW EPR spectra under these conditions are in rapid passage and thus exhibit an absorption lineshape.61, 62 EPR simulations used the program QPOW63, 64 and ENDOR simulations used the locally written program DDPOWHE.
Magnetometry.
Magnetic measurements for CoX were performed using a Quantum Design MPMS 3 magnetometer. All samples were prepared under a N2 atmosphere in polyethylene capsules and were solid powders restrained with eicosane in a gelatin capsule. Ferromagnetic impurities were ruled out by inspection of the 100 K magnetization data that showed no curvature in the field range of 0 – 7 T. DC magnetic susceptibility measurements were collected in the temperature range of 2 – 300 K under an applied magnetic field of 0.1 T. Variable-temperature, variable-field (VTVH) magnetization measurements were collected in the temperature range of 2–10 K under applied magnetic fields of 1 – 7 T, in 1 T increments. DC magnetic susceptibility measurements were corrected for diamagnetism, estimated using Pascals constants.65 Magnetic susceptibility and VTVH magnetization data were simulated using the program MagProp in DAVE 2.0.66
FIRMS.
FIRMS experiments were performed at NHMFL using a Bruker Vertex 80v FT-IR spectrometer coupled with a 17 T vertical-bore superconducting magnet in a Voigt configuration (light propagation perpendicular to the external magnetic field). The experimental setup employs broadband terahertz radiation emitted by an Hg arc lamp. The radiation transmitted through the sample is detected by a composite silicon bolometer (Infrared Laboratories) mounted at the end of the quasi-optical transmission line. Both the sample and bolometer are cooled by low-pressure helium gas to a temperature of 5.5 K. To obtain air-free measurements, the samples were loaded in the sample holder in an argon-filled glovebox. The microcrystalline powder (~3 – 5 mg) was bonded by n-eicosane and sandwiched between two n-eicosane layers for protection from oxygen and moisture. Sample loading in the FIRMS spectrometer was performed under a flow of N2. After collection, the samples were exposed to ambient conditions for two days, after which the measurements were recollected to ensure the originally observed absorption peaks were not attributable to sample degradation from contact with oxygen or moisture (Figures S8 and S9). The intensity spectra of each sample were measured in the spectral region between 14 and 730 cm−1 (0.42 − 22 THz) with a resolution of 0.3 cm−1 (9 GHz). To discern the magnetic absorptions, the spectra were normalized by dividing with the reference spectrum, which is the average spectrum for all magnetic fields. Such normalized transmittance spectra are sensitive only to intensity changes induced by the magnetic field and therefore are not obscured by nonmagnetic vibrational absorption features. The data analysis was implemented using an in-house written MATLAB code and the EPR simulation software package EasySpin,67, 68 which uses a standard spin Hamiltonian for S = 2 (FeX) and S = 3/2 (CoX).69
Ligand Field Theory (LFT) Calculations.
Calculations employed the locally written programs DDN and DDNFIT and the Ligfield software by Bendix (Copenhagen U., Denmark).70 These programs employ all 210 microstates for d6 (FeCl, FeCH3), and all 120 microstates for d7 (CoCl, CoCH3) with the angular overlap model (AOM)71–73 to describe σ- and π-bonding using respectively the parameters εσ and επ (which can be anisotropic: επ−s and επ−c).
Quantum Chemical Theory (QCT) Calculations.
All calculations were performed using the Orca 4.2 program package.74 Density Functional Theory (DFT) was used to calculate the 57Fe quadrupole splitting (ΔEQ) and isomer shift (δ) for FeX, and the 57Fe and 59Co A-tensors and for CoX the 14N A- and P-tensors. Calculations were performed using the atomic coordinates from the reported X-ray structures using the B3LYP/ CP(PPP) (Fe, Co), def2-TZVP (N, Cl, coordinated C), def2-SVP (C/H) functional/basis set combination.75–78 The calculated electron density at the Fe nuclei were converted into 57Fe Mössbauer isomer shift values using the calibration reported by Römelt et. al. 79 For these calculations the spin-orbit coupling operator was computed using the mean-field approximation (SOMF). Time Dependent DFT (TD-DFT) calculations were performed with the same functional/basis set described above except for Fe/Co which were changed to def2-tzvp. The TD-DFT calculations included 250 roots and the computed UV-visible absorption spectrum is reported in the SI.
The state averaged-CASSCF (SA-CASSCF) calculations used the minimum active space of six (FeX) or seven (CoX) electrons in the five 3d orbitals and, for FeX, included all five quintet and all 45 triplet states while the CoX compounds considered all 10 quartet and all 40 doublet states.
The resolution of the identity approximation and auxiliary basis sets generated using the ‘autoaux’ command were used in all CASSCF calculations.80 Scalar relativistic effects were accounted for by the second-order Douglas-Kroll-Hess (DKH) procedure and appropriate basis sets dkh-def2-TZVP (Fe, N, Cl, coordinated C) and dkh- def2-SVP (C/H).81 The converged wavefunctions were then subjected to N-electron valence perturbation theory to second order (NEVPT2) to account for dynamic correlation.82 Example ORCA input files are shown in the Supporting Information.
Results and Discussion
Structures.
The crystal structures of all four complexes have been reported previously: FeCl,53 (CSD: REWZUO), FeCH3,4 (CSD: XOXHUN), and CoCl (CSD: XUNTAB) and CoCH3 (CSD: XUNTEF).54 All four complexes have a crystallographic two-fold symmetry axis coincident with the X-M vector so that the two ∠N-M-X (X = Cl, C) are equal and the molecules have roughly C2v point group symmetry. This C2 axis would normally be defined as the z axis, but given that these complexes are derived from trigonal planar geometry, we define the z axis as the pseudo three-fold axis (i.e., normal to the molecular plane) and the actual C2 axis is defined as x. This assignment, which was used also by Andres et al. for the FeII complexes,4 leads to a slight redefinition of the d orbital representations in C2v symmetry as described elsewhere and is shown in Table S7.83 Andres et al. also presented a d orbital energy scheme for both the “parent” trigonal planar (i.e., a hypothetical MX3 complex with D3h (or C3h) symmetry84) and the actual structure. As such a diagram is extremely useful for both the FeII and CoII complexes under study here, we reproduce it with slight modifications. Note that if there were only σ-bonding, then the degenerate orbitals would be lowest in energy; however, π-donation from the N donors (and from Cl in MCl) raises above ; the σ-antibonding degenerate orbitals are always highest in energy. Quantitative diagrams are given in Figures 10 and 11 for FeX and CoX, respectively. The hypothetical trigonal CoII complex is Jahn-Teller effect (JTE) active (4E″ ground state), so it would distort as in actual trigonal complexes, such as MoIII.85 However, there is not threefold symmetry in the present FeII and CoII complexes, because the bidentate β-diketiminate ligand constrains ∠N-M-N to roughly 95°. As a result, none of the (nacnac)MX complexes have an orbitally degenerate JTE active ground state; the ground state for FeX (X = Cl, CH3) is 5A1 and for CoX is 4A2.
Figure 10.

The left-hand side shows a qualitative orbital energy level diagram for FeX with the dominate, non-Aufbau, ground state configuration shown. The inset shows the axis definition which has been chosen to be consistent with previous studies of these compounds (see also Figure 1). The transitions and associated excited states are colored to indicate the orbital angular momentum operator responsible for their interaction with the ground state (red, ; blue, ; green, ). The diagram on the right shows the relative energies of the low-lying excited states, color coded as on the left.
Figure 11.

The left-hand side shows a qualitative orbital energy level diagram for CoX with the dominate ground state configuration shown. The inset defines the choice of axis which has been chosen to be consistent with previous studies of the Fe analogues of these compounds (see also Figure 1). The transitions and associated excited states are colored to indicate the orbital angular momentum operator responsible for their interaction with the ground state (red, ; blue, ; green, ). The diagram on the right shows the energies of the low-lying excited states for CoCl and CoCH3, color coded as on the left, and the reorganization of these states that is responsible for the change in the sign of D between CoCl and CoCH3.
Conventional (field-domain) EPR spectroscopy.
X-band EPR spectra using parallel mode detection86–90 were reported by Andres et al. for powder FeCl and FeCH3.4 The two FeII complexes each exhibited an X-band (9.27 GHz) signal at low field (maxima at 35.6 mT and 59.4 mT for FeCl and FeCH3, respectively; see Figure 8 in Andres et al.4). This high g′ value (~18.6 and ~11.2, respectively) signal arises from a transition within the mS = ±2 doublet,86, 88, 91 which is the spin ground state (i.e., D < 0) as indicated by its temperature dependence. The energy levels for both S = 3/2 and S = 2 systems are shown in Figure 2.
Figure 2.

Energy levels for spin sublevel states in quartet (left) and quintet (right) systems with D < 0. For illustrative purposes, the rhombicity is small (|E/D| = 0.1 for both). An EPR transition within the small splitting (~0.03|D|) between the mS = ±2 levels is observed at X-band for FeCl and FeCH3.4 The transitions indicated by arrows can be observed by FIRMS (both transitions to the mS = ±1 levels for S = 2 are readily observable while that to the mS = 0 level is less likely and is thus shown as a dotted line).
The X-band spectrum of CoCl in toluene solution is shown in Figure 3; the Supporting Information shows the X-band spectrum of powder CoCl (Figure S3) as well as the corresponding Q-band spectra (CW and pulsed; Figure S4). The Q-band toluene frozen solution spectrum is essentially the same as at X-band except for the loss of hyperfine resolution due to g-strain,92 but with slightly better determination of g values. The X-band spectra for the solid and frozen solution are essentially the same, except for the inevitable loss of resolution in the magnetically non-dilute powder, which indicates that the solid state (XRD) structure is maintained in toluene solution. Qualitatively, the conventional EPR spectra of CoCl are characteristic of an S = 3/2 system with zfs (2D*, D* = (D2 + 3E2)1/2) that is large in energy relative to the microwave quantum (~1.17 cm−1 at Q-band) and with D > 0, i.e., mS = ±1/2 ground state.91, 93, 94 The simulations thus employ an effective spin, S′ = 1/2, with effective g′ values, as opposed to the real S = 3/2. Use of perturbation theory formulas91, 95 allows an estimate as to the real g values along with the zfs rhombicity, |E/D|. These formulas give |E/D| = 0.065 and gx = 2.580, gy = 2.585, gz = 2.000 (giso = 2.39), which parameter set affords g′ = [4.644, 5.655, 1.975], coinciding with the experimental g′ = [4.62 – 4.70, 5.64 – 5.68, 1.96 – 1.97] (here ordered as g′x, g′y, g′z, rather than as g′max, g′mid, g′min, as in the simulation) with the range due to the use of two frequencies and different sample preparations. The real g tensor is thus essentially axial with g⊥ > 2, g∥ ≈ 2, and 2 < giso < 2.5, which is typical for d7 systems.
Figure 3.

Experimental X-band EPR spectra of CoCl as a toluene solution (black trace) recorded at 10 K (9.364 GHz). Simulation of the solution spectrum (red trace) uses: S′ = 1/2, g′ = [5.64, 4.60, 1.97] (defined simply by g′max, g′mid, g′min), A′(59Co) = [700, 400, 210] MHz (A′ collinear with g′), W (half-width at half-maximum (hwhm), Gaussian) = 320, 280, 60 MHz. The inset shows an expansion of the g′∥ region for the solution with the same simulation parameters.
The 59Co hyperfine coupling is very well resolved at g′∥ with an average splitting of 8.4 mT. The 59Co hyperfine coupling tensor determined by simulation is also an effective one, A′(59Co), i.e., defined in terms of coupling to S′ = 1/2, rather than to S = 3/2.96 It can be converted to a real (i.e., intrinsic) A(59Co) by multiplying each component by gi /g′i,97 so that A(59Co) ≈ [220, 320, 210] MHz (here ordered as Ax, Ay, Az, as with g), which gives Aiso ≈ 250 MHz. EPR spectra of high-spin CoII typically exhibit broad linewidths so that 59Co hyperfine coupling is unresolved.98 This is presumably a function of g- and A-strain (i.e., a distribution in these parameters due to structural heterogeneity) as well as superimposed ligand hyperfine coupling (typically from 14N, but also 31P98). One example where A(59Co) was well resolved is a five-coordinate complex with only O-donors (I = 0 ligands), pentakis(2-picoline N-oxide)cobalt(II) perchlorate, prepared as a doped powder (< 0.1 mol%) in the isomorphous ZnII host.98 This system exhibits g′ = [5.96, 3.56, 1.91] – thus similar values to CoCl, and hyperfine structure was resolved in both the gmax and gmin (gz, g∥) regions corresponding to an average A(59Co) = 243 MHz, essentially the same Aiso value as seen here. An EPR study by Tierney and co-workers of dihydrido[diphenyl]bis([3,4,5-methyl]-1-pyrazolyl)borate ([Ph2]BpnMe) complexes of CoII presented species with both mS = ±1/2 and mS = ±3/2 signals with resolved A(59Co) on certain features,99 despite the 14N-donor ligands. For example, (Ph2Bp)2Co had g′ = [5.50, 4.59, 2.00] – very close to CoCl, while Bp2Co had g′ = [4.73, 4.67, 2.03] – far more axially symmetric than CoCl, but with A(59Co)z = 298 MHz,99 still similar to that for CoCl.100 The hyperfine coupling for CoCl is also near those measured for low-spin CoII macrocyclic complexes with N4 coordination.101 Quantitative analysis of the spin Hamiltonian parameters for CoCl is given below in the computational section.
The situation with CoCH3 is quite different. In this case, conventional EPR spectra were very difficult to obtain due to the extreme air sensitivity of the complex, which required the use of sealed tubes, precluding Q-band measurements. Nevertheless, it was clear that CoCH3 exhibited spectra characteristic of an S = 3/2 system with D < 0 (i.e., mS = ±3/2 ground state). Such X-band spectra are distinctive in that they exhibit a very large g′∥ (i.e., at very low field, usually with resolved 59Co hyperfine coupling) and a very small g′⊥ (i.e., at very high field – by X-band standards). An example of such an EPR spectrum was reported for the three-coordinate CoII NHC complex [Co(CH2SiMe3)2(IPr)] (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene), which gave g′ = [8.85, 1.89, 1.10] with well-resolved 59Co hyperfine coupling at low field (splitting of 17.7 mT; the spectrum was not simulated).102 Figure 4 presents the low field region of the X-band EPR spectrum of CoCH3. In contrast to the NHC complex,102 no features attributable to g′⊥ (g′mid or g′min) were definitively observed (see Figure S5). These g′ value turning points may lie beyond the maximum field of the X-band spectrometer (~600 mT, g′ ≈ 1.1). Nevertheless, both g′z and A′z (A′max) are reasonably well determined and give values close to those reported for the NHC complex (17.7 mT102 versus an average splitting of 16.5 mT in CoCH3). If we assume gz = 2.85, based on the magnetometry (see below), then the real hyperfine coupling, Az ≈ (1850 MHz)(2.85/8.5) = 620 MHz, as opposed to 435 MHz using gz = 2.0. The real value for the NHC complex is likely similar. For further comparison, four-coordinate, homoleptic CoII complexes, [Co(OAsMePh2)4](ClO4)2 98 and [Co(NH2CSNH2)4](NO3)2,103 studied as powders doped into their corresponding ZnII hosts exhibited X-band spectra similar to that seen for CoCH3, with neither clearly providing g′mid or g′min values. The arsine oxide complex yielded g′z ≈ 8.1 and A′z = 1586 MHz;98 the thiourea complex was not analyzed quantitatively, but the resolved splitting appears to be ~20 mT,103 and thus consistent with the other cases. The key qualitative finding from conventional EPR of CoX is that the sign of D is opposite between CoCl (D > 0) and CoCH3 (D < 0), which is quantitatively analyzed in the computational section below.
Figure 4.

Experimental X-band EPR spectrum of CoCH3 in frozen toluene solution (black trace) recorded at 10 K (9.3616 GHz). Simulation (red trace) uses: S′ = 1/2, g′ = [8.50, 1.2, 1.0] (defined simply by g′max, g′mid, g′min; the last two g′ values are essentially arbitrary), A′(59Co) = [1950, 400, 400] MHz (A′ collinear with g′), W (half-width at half-maximum, Gaussian) = 400, 500, 500 MHz. It is not possible to match exactly the experimental lineshape, but the hyperfine splitting pattern is reproduced by the simulation. As with g′, the last two A′ components and linewidths are arbitrary as there is no reliable experimental data for their determination (see also Figure S5).
ENDOR Spectroscopy.
As described above, due to their previous investigation by conventional EPR,4 no such experiments were undertaken here on FeX. As also indicated above, conventional X-band EPR spectroscopy in frozen solution was fruitful for CoX, although Q-band EPR was not feasible for CoCH3. CoCl, in contrast, could be studied by Q-band EPR (Figure S4) and thus by ENDOR spectroscopy at this frequency as well. Signals due to 1H in CoCl are seen using CW 35 GHz ENDOR as shown in Figure S6 and discussed in Supporting Information. More important are the 14N signals from the nacnac ligand. Pulsed (Davies) ENDOR spectra for CoCl recorded across its EPR envelope are shown in Figure 5. These spectra could be analyzed quantitatively by ENDOR simulation using S′ = 1/2 and g’ as above, now also with A′(14N) and P(14N) (the nqc is a purely nuclear interaction, involving no electronic spin terms, so that it is determined regardless of whether S or S′ is used). The simulations assume that the two nacnac 14N are equivalent, which is true in the solid-state crystal structure. That the X-band EPR spectra in a powder and in solution are the same (see Figure S3) suggests that the two 14N are equivalent in solution as well. Such an assumption may be an oversimplification given that 14N ENDOR can, in ideal cases (e.g., single crystal studies of hemes / porphyrins that included 15N-enrichment104, 105) reveal slight magnetic differences among structurally equivalent ligands. The situation with CoCl is far from this ideal not only in lacking 15N-enrichment and single-crystals, but also having severe disadvantages with respect to the desirable, yet common, situation of S = 1/2 with hfc small relative to EPR linewidth: the high-spin state of CoCl so that the g/g′ factor operates but is not accounted for, and the 59Co hfc, which complicates the orientation selection ability of ENDOR106, 107 by making a given g′ (i.e., the field at which ENDOR is recorded) shifted/split by the 59Co hfc making the EPR linewidth used for simulation less meaningful. Nevertheless, a reasonable reproduction of the 14N ENDOR pattern is achieved as shown in Figure 5. The fit parameters are: A′(14N) = [11.8, 21.4, 7.3] MHz, P(14N) = [+1.26, −0.80, −0.46] MHz. The A′(14N) can be converted to A(14N) ≈ [11.8(2.580/5.68), 21.4(2.585/4.70), 7.3] MHz = [5.4, 11.8, 7.1] MHz, so A(14N)iso ≈ 8.2 MHz. Despite the widespread use of the nacac ligand platform, this represents, to our knowledge, the first determination of 14N hfc for the β-ketiminate donors. For comparison, such data for more “ENDOR-friendly” S = 1/2 systems such as (nacnac)CuCl108 and (nacnac)NiL, L = CO, thf,109 would be useful; especially the NiI complexes that are absent any non-1H hfc except from 14N. The closest comparison that can be made to the present A(14N)iso value is the result of Tierney and co-workers110 who found for Tp2Co that A(pyrazolyl-2-14N)iso = 11.8 MHz. The electronic structure of 6-coordinate Co(II) is much more complicated than that for CoX due to unquenched orbital angular momentum,110–113 but this A(14N)iso value is in the range of that observed here. Also relevant is the work of Walsby et al. on a CoII-substituted ZnII protein, Finger 3 of Transcription Factor IIIA, 97 which has a Cys2His2 coordination site. They found A(14N) = 7.2 MHz, close to what is seen here, and they point out that this value is close to that seen for histidine imidazole N coordinated to CuII, when the appropriate scaling factor of 2S = 3 is used:97 AS=1/2 = AS=3/2(3) = 24 MHz in our case.
Figure 5.

Pulsed 35 GHz 14N ENDOR spectra of CoCl in toluene frozen solution (black traces) with simulations (red traces). Experimental parameters: temperature, 2 K; microwave frequency, 34.846 GHz; Davies sequence with tπ = 80 ns, τ = 600 ns, trf = 15 μs with random hopping of rf, repetition rate, 20 ms; typically 10 scans. Simulation parameters: S′ = 1/2, A′(14N) = [12.2, 20.6, 7.1] MHz, P(14N) = [+1.26, −0.80, −0.46] MHz, WNMR = 0.5 MHz, WEPR = 800 MHz (both Gaussian, hwhm). The broad, isotropic EPR linewidth is an attempt to model the 59Co hfc. The A′(14N) and P(14N) tensors are each rotated by Euler angle α = 130° with respect to the g′ tensor so that the (A′,P)z (out-of-plane) direction remains along g′z, but (A′,P)x,y (in-plane) is along the N-Co bond (see Figure 1).
To contextualize the observed 14N nqc, we turn to the metalloporphyrin literature, namely single-crystal ENDOR studies by Brown and Hoffman on CuII(TPP) (TPP = 1,5,10,15-tetraphenylporphyrin)114 and by Scholes et al. on aquometmyoglobin (FeIII(PPIX), S = 5/2).104 For Cu(TPP) (doped into a Zn(TPP)(H2O) host), P(14N) = [−0.619, +0.926, (−0.307)] MHz114, 115 For myoglobin, the average of the four heme nitrogen donors gave P(14N) = [−0.77, +1.04, −0.27] MHz. These values for porphyrin pyrrole N donors are not only close to each other, despite the coordinated metal ions’ size, charge, and spin state, but in the same range as that observed here for β-diketiminate N donors. Note that, unlike in these single-crystal studies, we have less certainty as to the relative orientations of the g, A, or P tensors with respect to each other or to the molecular frame of reference, although we make assumptions based on the coordinate system in Figure 1.
Figure 1.

Qualitative d orbital energy diagram for (nacnac)MX complexes. The left diagram is for an idealized trigonal planar geometry (i.e., θ = 120°) with labels for D3h symmetry (thus ignoring the difference between N and X ligands). In the case of only σ-bonding, the orbitals would be lowest in energy; out-of-plane π-bonding (donation) by the N and X ligands raises them above in energy. The right diagram is for the idealized real geometry: planar but no longer trigonal (i.e., θ ≈ 130°). The labels are for C2v symmetry, but with the z axis of the Cartesian coordinate frame out of plane and the x axis along the C2 axis to correspond to the D3h definition. This leads to having a2 representation and having b2 representation, the reverse of the standard C2v definition. Orbital occupancy is shown with black arrows for FeII (d6, S = 2) and the magenta arrow additionally for CoII (d7, S = 3/2). Adapted with permission from Figure 10 in Andres et al., J. Am. Chem. Soc. 2002, 124, 3012 – 3025. Copyright 2002 American Chemical Society.
Magnetometry.
In contrast to conventional EPR spectroscopy, magnetometry (here DC susceptibility and VTVH-magnetization) can in principle directly provide the magnitude of D. We collected the dc magnetic susceptibility data for both CoCl and CoCH3, which are presented in the insets of Figure 6. For CoCl, the 300 K χMT value of 3.0 cm3K/mol supports a giso value of 2.52 and is consistent with the anisotropic g-values extracted from EPR spectroscopy (EPR analysis supports a giso value of 2.39). Similarly, CoCH3 displays a 300 χMT value of 2.62 cm3K/mol, consistent with giso = 2.36. For both CoCl and CoCH3, their χMT values begin to decrease below 150 K, ultimately reaching values of 1.97 cm3K/mol (CoCl) and 2.06 cm3K/mol (CoCH3) at 2 K. An initial estimation of their axial zero-field splitting parameters (D) was determined by fitting their dc susceptibility data (Table 1). For CoCl, this fitting yielded D = +59(3) cm−1 and g⊥ and g∥ values of 2.65 and 2.16, respectively. For CoCH3, this fitting yielded D = −91(5) cm−1 and g⊥ and g∥ values of 2.02 and 2.88, respectively. To gain a better estimate of the spin Hamiltonian parameters, we collected and fit the variable temperature, variable field (VTVH) magnetization data for CoCl and CoCH3 (Figure 6). Our best simulations afforded D = +55(2) cm−1, and g⊥ and g∥ values of 2.62 and 2.08, respectively, for CoCl, and D = −91(5) cm−1, and g⊥ and g∥ values of 2.17 and 2.85, respectively, for CoCH3. These bulk magnetization data qualitatively agree with the analysis of the EPR spectra for CoCl and CoCH3, whose spectra were consistent with CoCl possessing a large and positive D value, while CoCH3 possesses a large and negative value of D. More quantitatively, use of these g values from magnetization in the perturbation theory equations91, 95 gives for CoCH3 a viable range from g′x = g′y = 0, g′z = 8.55 for E/D = 0 to g′x = 0.50, g′y = 0.54, g′z = 8.50 for E/D = 0.08. This result suggests the futility of observing g′⊥ for CoCH3 by conventional EPR.
Figure 6.

Variable temperature, variable field (VTVH) magnetization data of CoCl (left) and CoCH3 (right) each collected in the temperature range of 2 – 10 K and field range 1 – 7 T. The 0.1 T dc susceptibility for both compounds are presented in the insets. The black traces are best fits to each of the VTVH magnetization and dc susceptibility data (see Table 1 for fit parameters).
Table 1.
Spin Hamiltonian parameters for MX complexes with CASSCF/NEVPT2 + SOC calculated parameters.
| Complex, technique | D (cm−1), E (cm−1), |E/D| | [gx, gy, gz], giso | g′x, g′y, g′z e |
|---|---|---|---|
| FeCl | |||
| FIRMS | −38.05 ± 0.1, −2.05 ± 0.1, 0.054 | [2.2, 2.2, 2.5], 2.3 | --- |
| Conventional EPR a | --- | --- | ---, ---, 10.9 |
| Calculated | –51.5, −2.1, 0.04 | [1.90, 2.01, 2.90], 2.27 | |
| FeCH3 | |||
| FIRMS | −36.92 ± 0.5, −1.22 ± 0.5, 0.033 | [2.2, 2.2, 2.5], 2.3 | --- |
| Conventional EPR a | --- | --- | ---, ---, 11.4 |
| Calculated | –50.1, −2.0, 0.04 | [1.91, 2.03, 2.89], 2.28 | |
| CoCl | |||
| FIRMS b | 55.2 ± 0.2, 0, 0 | 2.5 | --- |
| Magnetometry c | +55 ± 2, 0, 0 | [2.62, 2.62, 2.08], 2.44 | 5.24, 5.24, 2.08 |
| Conventional EPR d | > 0, ---, 0.065 | --- | 4.66(4), 5.66(2), 1.965(5) |
| Calculated e | +65.5, 4.6, 0.07 | [2.79, 2.85, 1.95], 2.53 | 4.98, 6.27, 1.92 |
| CoCH3 | |||
| FIRMS b | −49.4 ± 0.2, 0 | 2.5 | --- |
| Magnetometry c | −91 ± 5, 0, 0 | [2.17, 2.17, 2.85], 2.40 | 0, 0, 8.55 |
| Conventional EPR d | < 0, ---, --- | --- | ---, ---, 8.50 |
| Calculated e | –122.4, −4.9, 0.04 | [1.05, 2.09, 3.59], 2.24 | 0.13, 0.24, 10.75 |
Taken from Andres et al.4 using X-band EPR with parallel mode detection.
Only the D* parameter and g′∥ can be evaluated from our experimental FIRMS data. For CoCl, the most favorable assignment for D is given, but the range 55 cm−1 ≤ |D| ≤ 59 cm−1 covers both possible assignments. In contrast, FIRMS for CoCH3 is especially complicated by spin-phonon interactions. The most favorable assignment for D is given, but the range ~50 cm−1 ≤ |D| ≤ ~85 cm−1 encompasses all possible assignments. The positive sign of D for CoCl and negative sign for CoCH3 is inferred from their X-band EPR spectra, and is corroborated by magnetometry and calculations.
The values given are those from fits of VTVH magnetization data. Fits of DC susceptibility measurements gave for CoCl, D = +59(3) cm−1, g = 2.65, g∥ = 2.16 (giso = 2.49); for CoCH3, D = −91(5) cm−1, g = 2.02, g∥ = 2.88 (giso = 2.31). Perturbation theory equations91, 95 give g′x,y,z values (S′ = 1/2) derived from gx,y,z values obtained from fits of magnetometry using S = 3/2. Using the parameters from DC susceptibility the results are g′ = [5.30, 5.30, 2.16] for CoCl and g′z = 8.64 for CoCH3.
X-band EPR provides |E/D| = 0.065 for CoCl from the splitting of the g′ feature (see text) using the perturbation theory equations. The range of g′ values comprises both X- and Q-band EPR measurements as well as different sample preparations. This determination is impossible for CoCH3 as only the g′∥ feature is observed.
The g′ value ordering is taken here to match experiment using the conventional assignments in EPR and magnetometry where g∥ ≡ gz and g ≡ gx,y.
Far-infrared magnetic spectroscopy (FIRMS).
FIRMS allows direct evaluation of the zfs in an S = 2 system such as found for FeCl and FeCH3 which includes determination of the zfs rhombicity.94 Measurement of FeX powder samples with no applied field gives spectra with two absorption peaks, which are observed at 108.3 and 120.6 cm−1 for FeCl and at 107.2 and 114.5 cm−1 for FeCH3. Note that other signals are observed for which the frequency is independent of the magnetic field, and thus these are attributed to vibrational bands (phonons). The spectra are presented in Figure 7 (top), with additional spectra in Supporting Information (Figures S7 – S10). In an S = 2 system with D < 0 and E ≠ 0 (assumed E < 0, to correspond to the sign of D), these transition energies correspond to ~3|D − E| and ~3|D + E|, respectively (see Figure 2). The spin Hamiltonian parameters for the four studied complexes are given in Table 1.
Figure 7.

FIRMS color maps for FeCl (top, left), FeCH3 (top, right), CoCl (bottom, left), and CoCH3 (bottom, right) each collected at 5.5 K. The magnitude of the field-induced variation in the transmission spectrum is depicted in a color scale (see inset in left panels), which is same within each FeX and CoX pair of compounds. The part of the spectrum with large experimental error (>3%) is indicated in white. For FeX, the dashed lines indicate spectral positions of the magnetic resonance with the external magnetic field aligned along the zfs (D) tensor x (red lines), y (blue), and z (black) principal axes; for CoX, only axial fits were used so magenta lines indicate the field aligned along the x,y (perpendicular, ⊥) direction, with black lines again for the field aligned with z (parallel, ∥). These lines were generated visually and not by any automated fitting routine. Due to spin-phonon coupling, the non-magnetic transitions (phonons) show up as vertical lines and are often quite intense (dark blue). More traditional in appearance single-beam transmission far-IR spectra are presented in Figures S7 and S8, which respectively show the effects of applied field and of air exposure. Figure S9 presents color maps for each complex on a wider energy range (80 – 220 cm−1) and also shows the effects of air exposure. Figure S10 shows an attempt to identify the |S, mS〉 = |2, ±2〉 → |0〉 transition (ΔmS = 2; see Figure 2) in FeCH3.
In the case of an S = 3/2 spin system, such as CoCl and CoCH3, information on the rhombicity of the zfs (D) tensor cannot be obtained, only the zero-field energy gap between the mS = ±1/2 and ±3/2 Kramers doublets (see Figure 3), here denoted 2D* (alternatively as Δ). These spectra are presented in Figure 8 (bottom), with additional spectra in Figures S7 – S9. Although large field-induced changes in the transmission are observed for both these compounds (see Figure S7), 2D* values are challenging to extract because all the spectra are affected by strong spin-phonon coupling effects.116 As a result, the ground state in these complexes is vibronic (crystal field plus phonon (vibrational mode)), therefore resulting in a hybridization of the crystal field levels leading to a complex FIRMS pattern compared to what would be expected using a simple S = 3/2 spin Hamiltonian model, namely a single absorption at 2D* (Figure 2), which would split in applied field due to the Zeeman effect. Nevertheless, the complex FIRMS pattern in Figure 7 can be interpreted by inspection of absorption peaks in the FIR transmission spectra (e.g., Figures S7 and S8) as well as the fields/frequencies where the more pronounced phonon peaks (i.e., dark vertical lines) exhibit crossing behavior, namely a drop in intensity due to a spin-phonon interaction at that point. An example is seen at ~106 cm−1 and ~4 T for CoCl. From this analysis we suggest the inter-doublet energy gap (see Figure 2) to be 2D* = 110.4 cm−1 for CoCl, which agrees well with the value from magnetometry (2D* = 114(4) cm−1; see Table 1). CoCl exhibits an additional, nearby feature at 117.2 cm−1 that might also be due to magnetic resonance absorption, which also agrees with magnetometry and calculations (see below). The situation for CoCH3 is more ambiguous. There are zero-field absorptions at 98.8 cm−1, 112 cm−1, 137 cm−1, and 169.3 cm−1 (see Figures S8 and S9 for the higher energy region). Among these, we favor assignment of the band at 98.8 cm−1 to 2D*, as given in Table 1, but none of the others can be totally ruled out. Thus, FIRMS suggests that for CoCH3, −85 cm−1 ≤ D ≤ −50 cm−1, with the negative sign based on magnetometry and conventional EPR.
Figure 8.

Electronic absorption spectra (main figure, Vis-NIR region; inset UV-Vis region) on a wavenumber (energy) scale of FeX (left panel) and CoX (right panel) recorded at room temperature in toluene solution. Complete UV data are unavailable for FeCl.
Electronic absorption spectra.
The electronic absorption spectra of FeX and CoX recorded at room temperature in toluene solution are shown in Figure 8 on an energy (wavenumbers, cm−1) scale. These spectra are shown on a wavelength scale in Figures S1 and S2 respectively for FeX and CoX. A simple ligand field theory (LFT) discussion of these spectra is given below followed by a definitive explanation using quantum chemical theory (QCT), specifically time-dependent density functional theory (TD-DFT).
Ligand field theory (LFT): optical spectra and zfs.
A quantitative analysis of the electronic structure of the MX series is provided using QCT in the following section, but we first discuss classical LFT because it is still instructive. We begin with an idealized trigonal planar geometry as shown in Figure 1 (left panel), Even in this relatively high D3h symmetry the number of states is very large as shown in Table S1 (Supporting Information). To explain qualitatively the electronic absorption spectra of the complexes we employ only spin-allowed d-d transitions as their possible origin. Considering first the FeX complexes (Figures 8 (left) and S1), there are visible absorption bands at 559 nm (17 890 cm−1; ε = 1700 mol−1 L cm−1) for FeCl and at 517 nm (19 340 cm−1; ε = 590 mol−1 L cm−1) for FeCH3. In D3h symmetry, with only σ-bonding and ignoring the JTE, the ground state is 5E″, with the β electron of d6 in ) with the first excited state being 5A1′ (β electron in ) and then 5E′ (β electron in ) followed by the numerous triplet and singlet excited states. The transition 5E″ → 5E′ is dipole allowed with z polarization (5E″ → 5A1′ is forbidden; 5A1′ → 5E′ is dipole allowed with polarization) so this could be the origin of the visible band. However, this assignment would require unreasonably large bonding parameters (εσ = 15 902 and 19 191 cm−1 for FeCl and FeCH3, respectively) so that a simple σ-only bonding model disfavors a d-d assignment for the visible band in FeX. Another option is to include π-bonding, which could be either donating or accepting. This is a serious complication in the real FeX complexes as the methyl ligand would have no π-bonding, the chlorido ligand would be cylindrical (επ−s = επ−c), and π-bonding involving the nacnac with its sp2 nitrogen ligands would likely be only out-of-plane (επ−s = 0, επ−c ≠ 0). To maintain D3h symmetry in the present model, we use the same π-bonding for all three ligands. In plane π-bonding could be included to lower εσ (e.g., επ−s = 0.2εσ gives εσ = 13 572 cm−1 for FeCH3), but its value is still too large. Out-of-plane π-bonding could instead be included, which has the advantage that with sufficient π-donation, the ground state becomes 5A1′, as in Figure 1 (left). The problem is that this effect brings the 5E′ excited state lower in energy – further from the observed band energy making fitting even less viable. Given the total failure of the trigonally symmetric D3h model, it is not worthwhile to modify it to closer to the real, C2v symmetry. We conclude that the visible band in FeX is likely a charge transfer (CT) band (whether metal-to-ligand (MLCT) or ligand-to-metal (LMCT) is uncertain), which is supported by its relatively high molar absorption coefficient.
In the CoX complexes (see Figures 8 (right) and S2) there are visible bands at 514 and 540 nm (19 455 and 18 520 cm−1; both ε ≈ 150 mol−1 L cm−1) and at 635 nm (15 750 cm−1; sh) and 687 nm (14 555 cm−1; ε = 130 mol−1 L cm−1) for CoCl and at 565 nm (17 700 cm−1; ε = 130 mol−1 L cm−1) and at 640 nm (15 625 cm−1; shoulder (sh)) and 725 nm (13 790 cm−1; ε = 160 mol−1 L cm−1) for CoCH3. These lower molar absorption coefficients support the assignment of these bands as being d-d transitions. In contrast to FeX, the quartet electronic states of CoX are more complicated. The free-ion 4F splits into 4A2′, 4E″, 4A1″ and 4A2″ (degenerate in D3h), and 4E′, with the states derived from 4P, 4A2′ and 4E″, higher in energy. Table S2 (Supporting Information) illustrates the states for idealized CoX considering only σ-bonding (εσ = 7000 cm−1) and with Racah parameters at 70% of their free-ion values.117 Using this very simplified, idealized model, one can more quantitatively rationalize the observed electronic transitions for both CoCl and CoCH3 as being viable as d-d transitions, in contrast to the situation for FeX. This is done using three possible sets of assignments: a) the lower energy visible band corresponds to 4A2′ → 4A1″,4A2″ and the higher energy band to 4A2′ → 4E′ (both allowed); b) the lower to 4A2′ → 4E′ and the higher to 4A2′ → 4E″(P) (forbidden in D3h, but allowed in the lower real symmetry); and c) the lower to 4A2′ → 4A1″,4A2″ and the higher to 4A2′ → 4E″(P). The transitions 4A2′ → 4E″(F) and 4A2′ → 4A2′(P) would be respectively too low and too high in energy to be observed (see Table S2) as well as being dipole forbidden in D3h. Fits were made using these three models with variable εσ and with the Racah B parameter either fixed at 70% or variable (Racah C in this model is large since doublet states were ignored). As expected, fits with fixed B were not very successful, yet model (b) agreed reasonably well with experimental data. Allowing B to vary led to perfect fits for both models (a) and (b); however, the fit values for model (a) were less realistic in that values for B were low and for εσ very high. In contrast, model (b) gave perfect fits with reasonable values for both parameters. The results are given in Table S3 (Supporting Information). LFT thus provides an idea as to the origin of the electronic transitions observed for the MX and in particular CoX complexes.
The next step is to use the idealized C2v geometry with the above bonding parameters as a guideline and explore the ability of LFT to model the zfs. Use of the actual ∠X-Co-N = 130.6 ± 0.3o and εσ(X) ≠ εσ(N) (επ(X,N) = 0) successfully fits the absorption bands (Table S4) with the B value and bonding parameters still reasonable. For both CoCl and CoCH3, εσ(X) > εσ(N). This is expected for the methyl anion, and in the case of chloride, this parameter also includes π-donation that is not specifically accounted for so as to avoid overparameterization. We acknowledge that π-bonding involving the chloride is key in that it removes the degeneracy of the and orbitals, as discussed in the QCT section, but we cannot quantify it here based on the available data. The results of these fits can then be used with inclusion of SOC to attempt to reproduce the spin Hamiltonian parameters. In this case, the Racah C parameter and the SOC constant ζ are both chosen to have the same reduction from their free-ion values as the fits obtained for B; though this is an oversimplification, it is useful for illustration. For CoCl, this model gives g′ = [5.73, 2.84, 1.76] and 2D* = 41 cm−1, ignoring the small rhombicity gives D ≈ 20 cm−1 – lower than the FIRMS value (Table 1), but this could be increased by an larger ζ value.118 The sign is positive based on the spin magnitudes (lowest doublet is 〈Sz2〉 = ±0.44; the higher doublet has 〈Sz2〉 = ±1.43). For CoCH3, use of ζ = 425 cm−1 gives g′ = [7.02, 1.76, 1.27] and 2D* = 74 cm−1. Surprisingly, although the g′ tensor is not unreasonable, giving one large and two small components, the spin magnitudes (lowest doublet 〈Sz2〉 = ±0.32; higher doublet 〈Sz2〉 = ±1.31) do not support a negative D value.
Quantum chemical theory (QCT): Time-dependent DFT (TD-DFT).
The above LFT section discusses the electronic absorption spectra of the MX series. For the FeX complexes, it was proposed that the observed bands were due to transitions that involved the ligands, rather than d-d transitions that could be modelled using LFT. As shown in Figures S11 and S12, this is indeed the case. Figure S11 presents the calculated spectra for the entire MX series so that it is readily apparent that the CoX complexes exhibit d-d transitions in the visible-NIR region, while the FeX complexes lack these. In particular, the energy of the visible band at 559 nm (17 900 cm−1) in FeCl (Figure 8) likely corresponds to the lowest energy LMCT transition calculated by TD-DFT (Figure S11) at 20 418 cm−1 (Figure S12). For FeCH3, no such distinct band is calculated, but the LMCT shoulders extend into the region where the band at 517 nm (19 340 cm−1; Figure 8) is observed. For the CoX complexes, the visible bands were assigned to d-d transitions and analyzed approximately using LFT, which description is confirmed by TD-DFT. The longer wavelength, more intense, visible bands at 687 nm (14 555 cm−1) and 725 nm (13 790 cm−1) for CoCl and CoCH3, respectively, are d-d in character, with the electron density changes (see Figure S12) entirely on the CoII center. Their energies are matched by calculated bands (Figure S11) at 15 707 cm−1 and 13 426 cm−1 respectively for CoCl and CoCH3. Additionally, the shorter wavelength, less intense, pair of visible bands for CoCl at 514 nm (19 455 cm−1) and 540 nm (18 520 cm−1) are matched by a calculated band at 19 421 cm−1 (see Figure S11, which gives further discussion on this point).
Quantum chemical theory (QCT): ab initio calculated g-values and zfs.
Complete active space self-consistent field (CASSCF) calculations were performed on the series of complexes using the atomic coordinates derived from previously reported x-ray structures. The results of these ab initio calculations were mapped onto a ligand field Hamiltonian using the ab initio ligand field theory (AILFT) procedure.119 The AILFT orbital level diagram is presented in Figure 9 and the relevant parameters (3d single electron orbital energies, Racah and ζ values) are listed in Table 2. The Racah parameters determined by AILFT are greater than their free-ion values (by ~15 – 20% in B and 2 – 6% in C),117 so they should not be used in absolute sense, but are useful in comparison among the MX series in terms of showing that each pair of complexes with the same metal ion has essentially the same parameters despite the difference between X = Cl and CH3. The SOC constants so determined are ~93% of the free-ion values across all four complexes.
Figure 9.

AILFT d-orbital energy level diagram derived from CASSCF calculations for FeCl, FeCH3, CoCl, and CoCH3. This quantitative ordering is the same as that shown qualitatively in Figure 1, except therein the splitting between and orbitals is increased for illustrative purposes with lowest as in Andres et al. 4 This apparent discrepancy is the consequence of non-Aufbau occupation as described in the text.
Table 2.
CASSCF + SOC derived AILFT parameters (in cm−1) for MX series.
The AILFT analysis suggests that in both FeCl and FeCH3 the orbital is lowest in energy. However, the lowest state is shown to have a dominant configuration (~90%) where the orbital is doubly occupied (Figures 9 and 10 (left)). This non-Aufbau ground state indicates that there is a strong competition between electron repulsion and the ligand field, where the energy of the ground state is minimized when the lowest one-electron AILFT orbital remains singly occupied and the second lowest orbital is doubly occupied such that the lowest energy configuration is . This is the same orbital occupancy proposed previously by Andres et al.,4 but their analysis did not reveal the non-Aufbau occupation pattern so that their Aufbau occupancy, as shown in Figure 1, is .
After the inclusion of spin orbit coupling, the calculations for both FeCl and FeCH3 predict negative axial values of the zero-field parameter, D, and with |E/D| = 0.04 for both compounds. This is consistent with their X-band EPR spectra, namely that these showed low field (high g′) transitions.4 For FeCl the ground state quasi-doublet is calculated to be separated by ~0.27 cm−1 (i.e., in the X-band EPR energy range) with the first excited states at 138 and 151 cm−1. The calculations on FeCH3 predict a nearly identical separation of the ground quasi-doublet (0.26 cm−1) and first excited quasi-doublet at 135 and 148 cm−1. These values are all reasonably close to the experimental values and reproduce the slightly larger energy transitions in the FIRMS results of FeCl compared to the FeCH3. The origin of this large zfs is understood by examining the low-lying excited states and their interaction with the ground state via spin-orbit coupling (Figure 10). By far the largest contribution to zfs is the coupling between the nearly degenerate orbital pair ( and ). This is consistent with the explanation previously given by Andres et al.4
As seen in Figure 9, the AILFT orbital level diagram predicts the same qualitative ordering of the d-orbitals in CoCl and CoCH3 as their Fe analogues. Interestingly, in both CoX complexes the CASSCF-AILFT derived ground state contains two major configurations (55% in CoCl, 45% CoCH3) and (42% in both CoCl and CoCH3). The experimentally determined and CASSCF/NEVPT2 calculated (in parenthesis) g′-values of CoCl are gx′ = 5.64 (6.12), gy′ = 4.62 (4.94), and gz′ = 1.97 (1.92). These values are consistent with a positive value of D, where the calculated value of D = +65.5 cm−1 (|E/D| = 0.07 – in good agreement with the experimental value from X-band EPR of CoCl) corresponding to a zero-field gap of 2D* = 132 cm−1, a slight overestimation of the experimental value. As discussed above, the X-band EPR spectrum of CoCH3 shows only a single observable g′-value, gz′(max) = 8.50, which suggests this compound displays a negative D. The CASSCF/NEVPT2 calculations result in gz′ = 9.94, D = −122.5 cm−1, |E/D|= 0.04 that corresponds to a zero-field gap of 245 cm−1. The difference in the sign of D between CoCl and CoCH3 can be rationalized by examining how the excited states couple through the spin-orbit interaction into the ground state. This is shown in Figure 11 (right), which reveals that there is a substantial reordering of the excited state energies between the two compounds. This energetic reordering is responsible for the change in the sign of D and can be explained qualitatively by the difference between the methyl and chlorido ligands. The increase in energy of both the 4B1 and, more importantly, 4A2 excited states upon going from CoCH3 to CoCl (Figure 11, right; also Figure 9) arises from the increase in energy of the b1 () and a2 () orbitals as result of adding π-donation from Cl− (along x) that is absent in the σ-only donor CH3−. There is a counteracting decrease in energy of the 4A1 excited state upon going from CoCH3 to CoCl caused by the a1 () orbital decreasing in energy as the stronger σ-donor methyl is replaced by chloride. But this reordering has a lesser effect on the zfs as 4A1 is more weakly coupled to the 4B2 ground state than either 4B1 and 4A2 are. Additionally, Table S6 lists the major SOC contributions from excited states to the zfs of CoX.
The AILFT values for ζ, B, and C can be input into the LFT software (DDN program) along with the single electron d orbital energies (Table 2) to yield spin Hamiltonian parameters for modelling MX as a pure dn system. For the FeX complexes, this demonstrates that there is a single g′ that is large, consistent with experiment. Use of an applied field of 50 mT (corresponding roughly to the experimental X-band resonant field) gives g′ ≈ 11.3(1) for both FeCl and FeCH3, in good agreement with the observed values (10.9 for FeCl and 11.4 for FeCH3).4 Moreover, the direction is correct in that this is g′ along x – the Fe-X bond. The splitting of the ground state quintet gives −50 cm−1 > D > −70 cm−1 for both FeX complexes, the negative sign and large magnitude as seen experimentally. In the case of CoCl, this procedure gives a splitting 2D* ≈ 170 cm−1. Ignoring the rhombic splitting, which cannot be extracted from this calculation and is in any case small, gives D ≈ +85 cm−1, ~50% off from experiment (Table 1). The sign can be readily determined since the lower doublet has spin magnitude ±0.31 and is thus mS = ±1/2, while the higher doublet has spin magnitude ±1.19 and is mS = ±3/2. These values were calculated using an applied field B0 = 300 mT – a typical X-band resonant field. This calculation also yielded g′ = [2.38, 8.55, 1.32] (given as gx, gy, gz, rather than the observed gmax, gmid, gmin). This g′ is quite different from that observed for CoCl, and requires a rather peculiar set of intrinsic parameters: E/D = 1/3 with g = [2.38, 3.13, 1.80]. For CoCH3, the spin magnitudes are more ambiguous (lower doublet ±0.06; higher doublet ±0.44), but the higher spin doublet clearly corresponds to mS = ±1/2, hence the negative sign given here for D. The calculated 2D* ≈ 278 cm−1, so D ≈ −139 cm−1, larger magnitude that what is found experimentally, and g′ = [0.36, 10.08, 0.24] (again as gx, gy, gz, rather than the observed gmax, gmid, gmin), which is consistent with the limited experimental information. Overall, AILFT reproduces well the low-field X-band EPR resonances observed for both FeX and CoCH3, although not the “conventional” EPR signature of CoCl, with the g anisotropy and zfs of both CoX complexes being overstated.
QCT Calculated Mössbauer Parameters and 57Fe, 59Co, and 14N Hyperfine Couplings.
Thanks to the use of variable applied magnetic fields, along with the standard Mössbauer parameters, namely isomer shift, δ, and quadrupole splitting ΔEQ, the study by Andres et al. reported an internal magnetic field, Bint, along the x axis (i.e., C2 axis, see Figure 1) that was quite large, +62 T for FeCl and +82 T for FeCH3.4 Subsequently, in a study of N2-binding involving related FeII and FeI β-diketiminate complexes,121 DFT calculations were performed on FeCl and FeCH3 as well as on the novel complexes in that work. The relevant results are summarized in Table S5. This was a pioneering study given the state of DFT calculations at that time and Stoian et al. were able to reproduce the δ and ΔEQ values reasonably well, even given the near orbital degeneracy of the FeX system. Indeed, we obtain essentially the same results for these parameters, despite using more highly developed software and having the benefit of extensive computational benchmarking of 57Fe Mössbauer data.79, 122 We note that although ab initio (CASSCF) methods are suitable for the purely electronic parameters (i.e., zfs as well as orbital energies), the use of DFT for nuclear-electronic parameters (i.e., hyperfine coupling) we believe is still the optimal approach.123–126 The calculated quadrupole splitting is reasonably close to experiment (see Table S5) although the calculated asymmetry parameter (η = Vmid − Vmin)/Vmax) is not, being too high for FeCl and too low for FeCH3. In D3h (or C3h) symmetry as in the idealized complex in Figure 1 (left), η = 0, so the results here demonstrate the difficulty in quantifying the in-plane bonding (electron distribution) in FeX albeit not the out-of-plane behavior. The situation with respect to hyperfine coupling is more complicated as this depends on the difference between α and β spin densities at the nucleus as opposed to their sum, which determines the isomer shift. As noted above, the complete A(57Fe) tensor was not determined for FeX, only the component along the Fe-X bond. We find here that the largest magnitude calculated component is quite far off from experiment (see Table S5) indicating the challenge of such calculations even with current computational power. Lacking the extreme orbital near-degeneracy of the FeX complexes, the CoX complexes present a potentially more fruitful area for hyperfine coupling calculations. In principle, the quadrupole coupling (i.e., yielding the electric field gradient, Vii) of 59Co could also be determined from EPR/ENDOR, as is possible for (excited state) 57Fe from Mössbauer, but this was not possible and rarely is, although it has been determined for 51V in vanadyl complexes.127 We also observe only one component of the hyperfine coupling in CoCH3. The calculated A(59Co) for CoCl appears to underestimate the overall coupling as well as being much more anisotropic than observed. We have no explanation for this discrepancy other than the difficulty in quantifying the small difference between α and β spin populations, which affected the FeX calculations as well. Despite this, the calculated A(14N) for the nacnac ligands in CoCl agrees reasonably well with experiment (Table S5), with Aiso(14N) differing by only ~1 MHz (~15 – 20%). The calculation also supports the two 14N being essentially magnetically equivalent. The calculated 14N quadrupole coupling for CoCl also matches experiment with each component agreeing within ~0.1 MHz (Table S5).
Conclusions
The β-diketiminate (nacnac) ligand is widely used in coordination chemistry and can support complexes with low coordination numbers. The electronic structure of four pairwise related nacnac-supported three-coordinate complexes, two each of FeII and CoII and two each with chlorido and methyl ancillary ligands, is examined in detail here. The two iron(II) complexes had been previously studied by applied-field Mössbauer spectroscopy,4 but QCT calculations of the type performed here were impossible 20 years ago. The zfs in these FeII complexes was inferred earlier only from the Mössbauer measurements. Here we directly observe their zfs thanks to the use of far-infrared magnetic spectroscopy (FIRMS). These measurements definitively show the large magnitude negative zfs in both FeCl and FeCH3 (respectively, D = −38 and −37 cm−1), with very low rhombicity (|E/D| = 0.05 and 0.03, respectively). It is notable that despite the differences between a methyl and chlorido ligand, the zfs is essentially the same in the two FeX complexes. This finding has implications for the design of SMMs in that the overall geometry rather than the identities of the coordinating ligands may be the key factor, at least in S = 2 systems. In the case of the two cobalt(II) complexes, no theoretical analysis of electronic structure or advanced spectroscopic measurements had been previously performed. In these Kramers (half-integer spin) ions, conventional EPR and ENDOR spectroscopy provided information on metal (via 59Co hyperfine coupling from EPR) and ligand (via 14N hyperfine coupling from ENDOR) spin delocalization. The results for both metal and ligand are consistent with those for related CoII complexes, and the present study provided the first measurement of a 14N hyperfine coupling for a β-diketiminate complex. The CoX complexes exhibit large magnitude zfs (respectively, D = +55 and −49 cm−1), but with a sign change from positive in CoCl to negative in CoCH3 that comes from a rearrangement of excited-state energies due to the donor properties of Cl− versus CH3−. Thus, the nature of the third donor may be crucial in determining the details of electronic structure in S = 3/2 systems. Calculation of spectroscopic parameters obtained from Mössbauer and EPR spectroscopy in its various forms still represents a challenge even using ab initio methods – at least for the low-coordinate MX systems studied here. We hope that theoreticians will take up this challenge so that a better understanding of the origin of the parameters can be obtained, which will assist in the design of such complexes with desired magnetic properties.
Supplementary Material
Synopsis.
High-spin (S > 1/2), low coordination number complexes are of interest as single molecule magnets (SMMs). The β-ketiminate (nacnac) ligand can support such complexes, here of general formula (nacnac)MX, where M = FeII (S = 2) and CoII (S = 3/2), and X = Cl− and CH3−. Advanced paramagnetic resonance spectroscopy reveals the zero-field splitting (zfs) in these complexes and theory explains the origin of zfs and the complexes’ overall electronic structure.
Acknowledgements
We thank the National Institutes of Health for funding (GM-065313 to P. L. H.; F32-GM136179 to M. S. F.) Part of this work was performed at the National High Magnetic Field Laboratory which is supported by NSF Cooperative Agreement No. DMR-1644779 and the State of Florida. We thank Prof. Brian M. Hoffman (Northwestern University) for use of X- and Q-band EPR (ENDOR) spectrometers, which are supported by funding from the U. S. Department of Energy (DOE) Office of Science, Basic Energy Sciences (BES) under contract DE-SC0019342 (to B. M. H.) and the NSF (grant MCB-1908587 to B. M. H.). Part of this work was supported by the Laboratory Directed Research and Development (LDRD) program at Los Alamos National Laboratory (20230399ER). We also thank Dr. Hao Yang, Northwestern University, for assistance with the pulsed ENDOR measurements and Profs. David L. Tierney, Miami University, and Sebastian A. Stoian, University of Idaho, for helpful comments. We thank Prof. Jesper Bendix (Copenhagen U., Denmark) for the LFT program Ligfield.
Footnotes
The authors declare no competing financial interests.
Supporting Information
Additional EPR, ENDOR, and FIRMS spectra. Tables of LFT fits and QCT results. This material is available free of charge at http://pubs.acs.org/.
References
- 1.Alvarez S, Bonding and stereochemistry of three-coordinated transition metal compounds. Coord. Chem. Rev. 1999, 193–195, 13–41. [Google Scholar]
- 2.Panda A; Stender M; Wright RJ; Olmstead MM; Klavins P; Power PP, Synthesis and Characterization of Three-Coordinate and Related β-Diketiminate Derivatives of Manganese, Iron, and Cobalt. Inorg. Chem. 2002, 41, 3909–3916. [DOI] [PubMed] [Google Scholar]
- 3.Sanakis Y; Power PP; Stubna A; Münck E, Mössbauer Study of the Three-Coordinate Planar FeII Thiolate Complex [Fe(SR)3]− (R = C6H2-2,4,6-tBu3): Model for the Trigonal Iron Sites of the MoFe7S9:Homocitrate Cofactor of Nitrogenase. Inorg. Chem. 2002, 41, 2690–2696. [DOI] [PubMed] [Google Scholar]
- 4.Andres H; Bominaar EL; Smith JM; Eckert NA; Holland PL; Münck E, Planar Three-Coordinate High-Spin FeII Complexes with Large Orbital Angular Momentum: Mössbauer, Electron Paramagnetic Resonance, and Electronic Structure Studies. J. Am. Chem. Soc. 2002, 124, 3012–3025. [DOI] [PubMed] [Google Scholar]
- 5.Holland PL; Cundari TR; Perez LL; Eckert NA; Lachicotte RJ, Electronically Unsaturated Three-Coordinate Chloride and Methyl Complexes of Iron, Cobalt, and Nickel [J. Am. Chem. Soc. 2002, 124 (48), 14416−14424]. J. Am. Chem. Soc. 2003, 125, 11772–11772. [DOI] [PubMed] [Google Scholar]
- 6.Dai X; Warren TH, Discrete Bridging and Terminal Copper Carbenes in Copper-Catalyzed Cyclopropanation. J. Am. Chem. Soc. 2004, 126, 10085–10094. [DOI] [PubMed] [Google Scholar]
- 7.Melzer MM; Jarchow-Choy S; Kogut E; Warren TH, Reductive Cleavage of O-, S-, and N-Organonitroso Compounds by Nickel(I) β-Diketiminates. Inorg. Chem. 2008, 47, 10187–10189. [DOI] [PubMed] [Google Scholar]
- 8.Rose RP; Jones C; Schulten C; Aldridge S; Stasch A, Synthesis and Characterization of Amidinate–Iron(I) Complexes: Analogies with β-Diketiminate Chemistry. Chem. Eur. J. 2008, 14, 8477–8480. [DOI] [PubMed] [Google Scholar]
- 9.Power PP, Stable Two-Coordinate, Open-Shell (d1–d9) Transition Metal Complexes. Chem. Rev. 2012, 112, 3482–3507. [DOI] [PubMed] [Google Scholar]
- 10.Chandrasekaran P; Chiang KP; Nordlund D; Bergmann U; Holland PL; DeBeer S, Sensitivity of X-ray Core Spectroscopy to Changes in Metal Ligation: A Systematic Study of Low-Coordinate, High-Spin Ferrous Complexes. Inorg. Chem. 2013, 52, 6286–6298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Danopoulos AA; Braunstein P; Wesolek M; Monakhov KY; Rabu P; Robert V, Three-Coordinate Iron(II) N-Heterocyclic Carbene Alkyl Complexes. Organometallics 2012, 31, 4102–4105. [Google Scholar]
- 12.Davis TL; Watts JL; Brown KJ; Hewage JS; Treleven AR; Lindeman SV; Gardinier JR, Structural classification of metal complexes with three-coordinate centres. Dalton Trans. 2015, 44, 15408–15412. [DOI] [PubMed] [Google Scholar]
- 13.Pratt J; Bryan AM; Faust M; Boynton JN; Vasko P; Rekken BD; Mansikkamäki A; Fettinger JC; Tuononen HM; Power PP, Effects of Remote Ligand Substituents on the Structures, Spectroscopic, and Magnetic Properties of Two-Coordinate Transition-Metal Thiolate Complexes. Inorg. Chem. 2018, 57, 6491–6502. [DOI] [PubMed] [Google Scholar]
- 14.Rau IG; Baumann S; Rusponi S; Donati F; Stepanow S; Gragnaniello L; Dreiser J; Piamonteze C; Nolting F; Gangopadhyay S; Albertini OR; Macfarlane RM; Lutz CP; Jones BA; Gambardella P; Heinrich AJ; Brune H, Reaching the magnetic anisotropy limit of a 3d metal atom. Science 2014, 344, 988–992. [DOI] [PubMed] [Google Scholar]
- 15.Layfield RA, Organometallic Single-Molecule Magnets. Organometallics 2014, 33, 1084–1099. [Google Scholar]
- 16.Gómez-Coca S; Aravena D; Morales R; Ruiz E, Large magnetic anisotropy in mononuclear metal complexes. Coord. Chem. Rev. 2015, 289–290, 379–392. [Google Scholar]
- 17.Deng Y-F; Han T; Yin B; Zheng Y-Z, On balancing the QTM and the direct relaxation processes in single-ion magnets – the importance of symmetry control. Inorg. Chem. Front. 2017, 4, 1141–1148. [Google Scholar]
- 18.Reiff WM; LaPointe AM; Witten EH, Virtual Free Ion Magnetism and the Absence of Jahn−Teller Distortion in a Linear Two-Coordinate Complex of High-Spin Iron(II). J. Am. Chem. Soc. 2004, 126, 10206–10207. [DOI] [PubMed] [Google Scholar]
- 19.Reiff WM; Schulz CE; Whangbo M-H; Seo JI; Lee YS; Potratz GR; Spicer CW; Girolami GS, Consequences of a Linear Two-Coordinate Geometry for the Orbital Magnetism and Jahn−Teller Distortion Behavior of the High Spin Iron(II) Complex Fe[N(t-Bu)2]2. J. Am. Chem. Soc. 2009, 131, 404–405. [DOI] [PubMed] [Google Scholar]
- 20.Merrill WA; Stich TA; Brynda M; Yeagle GJ; Fettinger JC; Hont RD; Reiff WM; Schulz CE; Britt RD; Power PP, Direct Spectroscopic Observation of Large Quenching of First-Order Orbital Angular Momentum with Bending in Monomeric, Two-Coordinate Fe(II) Primary Amido Complexes and the Profound Magnetic Effects of the Absence of Jahn− and Renner−Teller Distortions in Rigorously Linear Coordination. J. Am. Chem. Soc. 2009, 131, 12693–12702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Freedman DE; Harman WH; Harris TD; Long GJ; Chang CJ; Long JR, Slow Magnetic Relaxation in a High-Spin Iron(II) Complex. J. Am. Chem. Soc. 2010, 132, 1224–1225. [DOI] [PubMed] [Google Scholar]
- 22.Lin P-H; Smythe NC; Gorelsky SI; Maguire S; Henson NJ; Korobkov I; Scott BL; Gordon JC; Baker RT; Murugesu M, Importance of Out-of-State Spin–Orbit Coupling for Slow Magnetic Relaxation in Mononuclear FeII Complexes. J. Am. Chem. Soc. 2011, 133, 15806–15809. [DOI] [PubMed] [Google Scholar]
- 23.Bryan AM; Merrill WA; Reiff WM; Fettinger JC; Power PP, Synthesis, Structural, and Magnetic Characterization of Linear and Bent Geometry Cobalt(II) and Nickel(II) Amido Complexes: Evidence of Very Large Spin–Orbit Coupling Effects in Rigorously Linear Coordinated Co2+. Inorg. Chem. 2012, 51, 3366–3373. [DOI] [PubMed] [Google Scholar]
- 24.Zadrozny JM; Atanasov M; Bryan AM; Lin C-Y; Rekken BD; Power PP; Neese F; Long JR, Slow magnetization dynamics in a series of two-coordinate iron(II) complexes. Chem. Sci. 2013, 4, 125–138. [Google Scholar]
- 25.Atanasov M; Zadrozny JM; Long JR; Neese F, A theoretical analysis of chemical bonding, vibronic coupling, and magnetic anisotropy in linear iron(II) complexes with single-molecule magnet behavior. Chem. Sci. 2013, 4, 139–156. [Google Scholar]
- 26.Zadrozny JM; Xiao DJ; Atanasov M; Long GJ; Grandjean F; Neese F; Long JR, Magnetic blocking in a linear iron(I) complex. Nat Chem 2013, 5, 577–581. [DOI] [PubMed] [Google Scholar]
- 27.Zadrozny JM; Xiao DJ; Long JR; Atanasov M; Neese F; Grandjean F; Long GJ, Mössbauer Spectroscopy as a Probe of Magnetization Dynamics in the Linear Iron(I) and Iron(II) Complexes [Fe(C(SiMe3)3)2]1–/0. Inorg. Chem. 2013, 52, 13123–13131. [DOI] [PubMed] [Google Scholar]
- 28.Samuel PP; Mondal KC; Amin Sk N; Roesky HW; Carl E; Neufeld R; Stalke D; Demeshko S; Meyer F; Ungur L; Chibotaru LF; Christian J; Ramachandran V; van Tol J; Dalal NS, Electronic Structure and Slow Magnetic Relaxation of Low-Coordinate Cyclic Alkyl(amino) Carbene Stabilized Iron(I) Complexes. J. Am. Chem. Soc. 2014, 136, 11964–11971. [DOI] [PubMed] [Google Scholar]
- 29.Bryan AM; Long GJ; Grandjean F; Power PP, Synthesis, Structural, Spectroscopic, and Magnetic Characterization of Two-Coordinate Cobalt(II) Aryloxides with Bent or Linear Coordination. Inorg. Chem. 2014, 53, 2692–2698. [DOI] [PubMed] [Google Scholar]
- 30.Eichhöfer A; Lan Y; Mereacre V; Bodenstein T; Weigend F, Slow Magnetic Relaxation in Trigonal-Planar Mononuclear Fe(II) and Co(II) Bis(trimethylsilyl)amido Complexes—A Comparative Study. Inorg. Chem. 2014, 53, 1962–1974. [DOI] [PubMed] [Google Scholar]
- 31.Schulte KA; Vignesh KR; Dunbar KR, Effects of coordination sphere on unusually large zero field splitting and slow magnetic relaxation in trigonally symmetric molecules. Chem. Sci. 2018, 9, 9018–9026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Bunting PC; Atanasov M; Damgaard-Møller E; Perfetti M; Crassee I; Orlita M; Overgaard J; van Slageren J; Neese F; Long JR, A linear cobalt(II) complex with maximal orbital angular momentum from a non-Aufbau ground state. Science 2018, 362, eaat7319. [DOI] [PubMed] [Google Scholar]
- 33.Thomsen MK; Nyvang A; Walsh JPS; Bunting PC; Long JR; Neese F; Atanasov M; Genoni A; Overgaard J, Insights into Single-Molecule-Magnet Behavior from the Experimental Electron Density of Linear Two-Coordinate Iron Complexes. Inorg. Chem. 2019, 58, 3211–3218. [DOI] [PubMed] [Google Scholar]
- 34.Sarkar A; Dey S; Rajaraman G, Role of Coordination Number and Geometry in Controlling the Magnetic Anisotropy in FeII, CoII, and NiII Single-Ion Magnets. Chem. Eur. J. 2020, 26, 14036–14058. [DOI] [PubMed] [Google Scholar]
- 35.Saber MR; Przyojski JA; Tonzetich ZJ; Dunbar KR, Slow magnetic relaxation in cobalt N-heterocyclic carbene complexes. Dalton Trans. 2020, 49, 11577–11582. [DOI] [PubMed] [Google Scholar]
- 36.Jung J; Legendre CM; Demeshko S; Herbst-Irmer R; Stalke D, Trigonal Planar Iron(II) and Cobalt(II) Complexes Containing [RS(NtBu)3]n− (R = NtBu, n = 2; CH2PPh2, n = 1) as Acute Bite-Angle Chelating Ligands: Soft P Donor Proves Beneficial to Magnetic Co Species. Inorg. Chem. 2021, 60, 9580–9588. [DOI] [PubMed] [Google Scholar]
- 37.Blackaby WJM; Harriman KLM; Greer SM; Folli A; Hill S; Krewald V; Mahon MF; Murphy DM; Murugesu M; Richards E; Suturina E; Whittlesey MK, Extreme g-Tensor Anisotropy and Its Insensitivity to Structural Distortions in a Family of Linear Two-Coordinate Ni(I) Bis-N-heterocyclic Carbene Complexes. Inorg. Chem. 2022, 61, 1308–1315. [DOI] [PubMed] [Google Scholar]
- 38.Münster K; Baabe D; Kintzel B; Böhme M; Plass W; Raeder J; Walter MD, Low-Coordinate Iron(II) Amido Half-Sandwich Complexes with Large Internal Magnetic Hyperfine Fields. Inorg. Chem. 2022, 61, 18883–18898. [DOI] [PubMed] [Google Scholar]
- 39.Khurana R; Ali ME, Single-Molecule Magnetism in Linear Fe(I) Complexes with Aufbau and Non-Aufbau Ground States. Inorg. Chem. 2022, 61, 15335–15345. [DOI] [PubMed] [Google Scholar]
- 40.Gatteschi D; Sessoli R, Quantum Tunneling of Magnetization and Related Phenomena in Molecular Materials. Angew. Chem. Int. Ed. 2003, 42, 268–297. [DOI] [PubMed] [Google Scholar]
- 41.Waldmann O, A Criterion for the Anisotropy Barrier in Single-Molecule Magnets. Inorg. Chem. 2007, 46, 10035–10037. [DOI] [PubMed] [Google Scholar]
- 42.Zadrozny JM; Liu J; Piro NA; Chang CJ; Hill S; Long JR, Slow magnetic relaxation in a pseudotetrahedral cobalt(II) complex with easy-plane anisotropy. Chem. Commun. 2012, 48, 3927–3929. [DOI] [PubMed] [Google Scholar]
- 43.Bourget-Merle L; Lappert MF; Severn JR, The Chemistry of β-Diketiminatometal Complexes. Chem. Rev. 2002, 102, 3031–3066. [DOI] [PubMed] [Google Scholar]
- 44.Mindiola DJ; Holland PL; Warren TH, Complexes of Bulky β-Diketiminate Ligands. In Inorg. Synth, 2010; pp 1–55. [Google Scholar]
- 45.Chen C; Bellows SM; Holland PL, Tuning steric and electronic effects in transition-metal β-diketiminate complexes. Dalton Trans. 2015, 44, 16654–16670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Webster RL, β-Diketiminate complexes of the first row transition metals: applications in catalysis. Dalton Trans. 2017, 46, 4483–4498. [DOI] [PubMed] [Google Scholar]
- 47.We shall employ the nomenclature nacnacR,R( wherein R = the N-substituent and R( = the “backbone” aliphatic group, e.g., Me for the simplest ligand, 2,4-bis([R]imido)-3-pentyl anion, and tBu for that used herein, 2,2,6,6-tetramethyl-3,5-bis([R]imido)-4-heptyl anion.
- 48.Arslan E; Lalancette RA; Bernal I, An historic and scientific study of the properties of metal(III) tris-acetylacetonates. Structural Chemistry 2017, 28, 201–212. [Google Scholar]
- 49.Latreche S; Schaper F, Chromium(III) Bis(diketiminate) Complexes. Organometallics 2010, 29, 2180–2185. [Google Scholar]
- 50.Lazarov BB; Hampel F; Hultzsch KC, Synthesis and Structural Characterization of β-Diketiminato Yttrium Complexes and their Application in Epoxide/CO2-Copolymerization. Z. Anorg. Allg. Chem. 2007, 633, 2367–2373. [Google Scholar]
- 51.C. Andrews P; L. Atwood J; J. Barbour L; D. Croucher P; J. Nichols P; O. Smith N; W. Skelton B; H. White A; L. Raston C, Supramolecular confinement of C60, S8, P4Se3 and toluene by metal(II) macrocyclic complexes. J. Chem. Soc., Dalton Trans. 1999, 2927–2932. [Google Scholar]
- 52.Kumar Sahu P; Kharel R; Shome S; Goswami S; Konar S, Understanding the unceasing evolution of Co(II) based single-ion magnets. Coord. Chem. Rev. 2023, 475, 214871. [Google Scholar]
- 53.Smith JM; Lachicotte RJ; Holland PL, Tuning metal coordination number by ancillary ligand steric effects: synthesis of a three-coordinate iron(II)complex. Chem. Commun. 2001, 1542–1543. [DOI] [PubMed] [Google Scholar]
- 54.Holland PL; Cundari TR; Perez LL; Eckert NA; Lachicotte RJ, Electronically Unsaturated Three-Coordinate Chloride and Methyl Complexes of Iron, Cobalt, and Nickel. J. Am. Chem. Soc. 2002, 124, 14416–14424. [DOI] [PubMed] [Google Scholar]
- 55.Brackett GC; Richards PL; Caughey WS, Far-infrared magnetic resonance in Fe(III) and Mn(III) porphyrins, myoglobin, hemoglobin, ferrichrome A, and Fe(III) dithiocarbamates. J. Chem. Phys. 1971, 54, 4383–4401. [Google Scholar]
- 56.Champion PM; Sievers AJ, Far infrared magnetic resonance in FeSiF6.6H2O and Fe(SPh)42−. J. Chem. Phys. 1977, 66, 1819–1825. [Google Scholar]
- 57.Chiang KP; Barrett PM; Ding F; Smith JM; Kingsley S; Brennessel WW; Clark MM; Lachicotte RJ; Holland PL, Ligand Dependence of Binding to Three-Coordinate Fe(II) Complexes. Inorg. Chem. 2009, 48, 5106–5116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Werst MM; Davoust CE; Hoffman BM, Ligand Spin Densities in Blue Copper Proteins by Q-band 1H and 14N ENDOR Spectroscopy. J. Am. Chem. Soc. 1991, 113, 1533–1538. [Google Scholar]
- 59.Davoust CE; Doan PE; Hoffman BM, Q-Band Pulsed Electron Spin-Echo Spectrometer and Its Application to ENDOR and ESEEM. J. Magn. Reson., Ser. A 1996, 119, 38–44. [Google Scholar]
- 60.Davies ER, A New Pulse ENDOR Technique. Physics Letters A 1974, 47, 1–2. [Google Scholar]
- 61.Mailer C; Hoffman BM, Tumbling of an adsorbed nitroxide using rapid adiabatic passage. J. Phys. Chem. 1976, 80, 842–846. [Google Scholar]
- 62.Mailer C; Taylor CPS, Rapid adiabatic passage EPR of ferricytochrome c. Signal enhancement and determination of the spin-lattice relaxation time. Biochimica et Biophysica Acta 1973, 322, 195–203. [DOI] [PubMed] [Google Scholar]
- 63.Belford RL; Nilges MJ In Computer Simulations of Powder Spectra, EPR Symposium, 21st Rocky Mountain Conference, Denver, Colorado, Denver, Colorado, August,1979. [Google Scholar]
- 64.Belford RL; Belford GG, Eigenfield expansion technique for efficient computation of field-swept fixed-frequency spectra from relaxation master equations. J. Chem. Phys. 1973, 59, 853–854. [Google Scholar]
- 65.Bain GA; Berry JF, Diamagnetic Corrections and Pascal’s Constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar]
- 66.Azuah RT; Kneller LR; Qiu Y; Tregenna-Piggott PLW; Brown CM; Copley JRD; Dimeo RM, DAVE: A Comprehensive Software Suite for the Reduction, Visualization, and Analysis of Low Energy Neutron Spectroscopic Data. Journal of Research of the National Institute of Standards and Technology; 2009, 114, 341–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Nehrkorn J; Telser J; Holldack K; Stoll S; Schnegg A, Simulating Frequency-Domain Electron Paramagnetic Resonance: Bridging the Gap between Experiment and Magnetic Parameters for High-Spin Transition-Metal Ion Complexes. J. Phys. Chem. B 2015, 119, 13816–13824. [DOI] [PubMed] [Google Scholar]
- 68.Stoll S EasySpin. http://www.easyspin.org/.
- 69.Abragam A; Bleaney B, Electron Paramagnetic Resonance of Transition Ions (Oxford Classic Texts in the Physical Sciences). Oxford University Press: Oxford, UK, 2012. [Google Scholar]
- 70.Bendix J, Ligfield. In Comprehensive Coordination Chemistry II, Volume 2: Fundamentals: Physical Methods, Theoretical Analysis, and Case Studies, Lever ABP, Ed. Elsevier: Amsterdam, 2003; Vol. 2, pp 673–676. [Google Scholar]
- 71.Schönherr T; Atanasov M; Adamsky H, Angular Overlap Model. In Comprehensive Coordination Chemistry II, McCleverty JA; Meyer TJ, Eds. Pergamon: Oxford, 2003; pp 443–455. [Google Scholar]
- 72.Schäffer CE; Yamatera H, Ligand field of conjugated bidentate ligands parametrized by the angular overlap model. Inorg. Chem. 1991, 30, 2840–2853. [Google Scholar]
- 73.Miessler GL; Fischer PJ; Tarr DA, Inorganic Chemistry. Pearson: Upper Saddle River, NJ, 2014. [Google Scholar]
- 74.Neese F, Software update: the ORCA program system, version 4.0. Wiley Interdisciplinary Reviews: Computational Molecular Science; 2018, 8, e1327. [Google Scholar]
- 75.Perdew JP, Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33, 8822–8824. [DOI] [PubMed] [Google Scholar]
- 76.Perdew JP, Erratum: Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 34, 7406. [PubMed] [Google Scholar]
- 77.Becke AD, Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [DOI] [PubMed] [Google Scholar]
- 78.Weigend F; Ahlrichs R, Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [DOI] [PubMed] [Google Scholar]
- 79.Römelt M; Ye S; Neese F, Calibration of Modern Density Functional Theory Methods for the Prediction of 57Fe Mössbauer Isomer Shifts: Meta-GGA and Double-Hybrid Functionals. Inorg. Chem. 2009, 48, 784–785. [DOI] [PubMed] [Google Scholar]
- 80.Stoychev GL; Auer AA; Neese F, Automatic Generation of Auxiliary Basis Sets. J. Chem. Theory Comput. 2017, 13, 554–562. [DOI] [PubMed] [Google Scholar]
- 81.Hess BA, Relativistic electronic-structure calculations employing a two-component no-pair formalism with external-field projection operators. Phys. Rev. A 1986, 33, 3742–3748. [DOI] [PubMed] [Google Scholar]
- 82.Angeli C; Bories B; Cavallini A; Cimiraglia R, Third-order multireference perturbation theory: The n-electron valence state perturbation-theory approach. J. Chem. Phys. 2006, 124, 054108. [DOI] [PubMed] [Google Scholar]
- 83.Xu S; Bucinsky L; Breza M; Krzystek J; Chen C-H; Pink M; Telser J; Smith JM, Ligand Substituent Effects in Manganese Pyridinophane Complexes: Implications for Oxygen-Evolving Catalysis. Inorg. Chem. 2017, 56, 14315–14325. [DOI] [PubMed] [Google Scholar]
- 84.A hypothetical MX3 (e.g., X = Cl) would have D3h symmetry, while C3h retains all of the symmetry relevant to the present discussion and allows for inclusion of a hypothetical M(XR)3 (e.g., XR = OR, SR) complex.
- 85.McNaughton RL; Roemelt M; Chin JM; Schrock RR; Neese F; Hoffman BM, Experimental and Theoretical EPR Study of Jahn−Teller-Active [HIPTN3N]MoL Complexes (L = N2, CO, NH3). J. Am. Chem. Soc. 2010, 132, 8645–8656. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Hendrich MP; Debrunner PG, Integer-spin electron paramagnetic resonance of iron proteins. Biophys. J. 1989, 56, 489–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Surerus KK; Hendrich MP; Christie PD; Rottgardt D; Orme-Johnson WH; Münck E, Moessbauer and integer-spin EPR of the oxidized P-clusters of nitrogenase: POX is a non-Kramers system with a nearly degenerate ground doublet. J. Am. Chem. Soc. 1992, 114, 8579–8590. [Google Scholar]
- 88.Münck E; Surerus KK; Hendrich MP, Combining Mössbauer Spectroscopy with Integer Spin Electron Paramagnetic Resonance. Methods in Enzymology 1993, 227, 463–479. [DOI] [PubMed] [Google Scholar]
- 89.Gupta R; Lacy DC; Bominaar EL; Borovik AS; Hendrich MP, Electron Paramagnetic Resonance and Mössbauer Spectroscopy and Density Functional Theory Analysis of a High-Spin FeIV–Oxo Complex. J. Am. Chem. Soc. 2012, 134, 9775–9784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Gupta R; Taguchi T; Borovik AS; Hendrich MP, Characterization of Monomeric MnII/III/IV–Hydroxo Complexes from X- and Q-Band Dual Mode Electron Paramagnetic Resonance (EPR) Spectroscopy. Inorg. Chem. 2013, 52, 12568–12575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Telser J, EPR Interactions – Zero-Field Splittings. In eMagRes, John Wiley & Sons, Ltd: 2017; Vol. 6, pp 207–233. [Google Scholar]
- 92.Telser J, Linewidth, field, and frequency in electron paramagnetic resonance (EPR) spectroscopy. J. Biol. Inorg. Chem. 2022, 27, 605–609. [DOI] [PubMed] [Google Scholar]
- 93.Krzystek J; Ozarowski A; Telser J, Multi-frequency, high-field EPR as a powerful tool to accurately determine zero-field splitting in high-spin transition metal coordination complexes. Coord. Chem. Rev. 2006, 250, 2308–2324. [Google Scholar]
- 94.Krzystek J; Telser J, Measuring giant anisotropy in paramagnetic transition metal complexes with relevance to single-ion magnetism. Dalton Trans. 2016, 45, 16751–16763. [DOI] [PubMed] [Google Scholar]
- 95.Banci L; Bencini A; Benelli C; Gatteschi D; Zanchini C, Spectral-Structural Correlations in High-Spin Cobalt(II) Complexes. Struct. Bonding 1982, 52, 37–86. [Google Scholar]
- 96.Telser J; Drago RS, Correction: Reinvestigation of the Electronic and Magnetic Properties of Ruthenium Butyrate Chloride. Inorg. Chem. 1985, 24, 4765–4765. [Google Scholar]
- 97.Walsby CJ; Krepkiy D; Petering DH; Hoffman BM, Cobalt-Substituted Zinc Finger 3 of Transcription Factor IIIA: Interactions with Cognate DNA Detected by 31P ENDOR Spectroscopy. J. Am. Chem. Soc. 2003, 125, 7502–7503. [DOI] [PubMed] [Google Scholar]
- 98.Makinen MW; Kuo LC; Yim MB; Wells GB; Fukuyama JM; Kim JE, Ground term splitting of high-spin Co2+ as a probe of coordination structure. 1. Dependence of the splitting on coordination geometry. J. Am. Chem. Soc. 1985, 107, 5245–5255. [Google Scholar]
- 99.Baum RR; Myers WK; Greer SM; Breece RM; Tierney DL, The Original CoII Heteroscorpionates Revisited: On the EPR of Pseudotetrahedral CoII. Eur. J. Inorg. Chem. 2016, 2016, 2641–2647. [Google Scholar]
- 100.Converting from g’ for Bp2Co yields g = [2.35, 2.35, 2.03] with E/D = 0.004, so Az = A′z.
- 101.Van Doorslaer S; Jeschke G; Epel B; Goldfarb D; Eichel R-A; Kräutler B; Schweiger A, Axial solvent coordination in “base-off” Cob(II)alamin and related Co(II)-corrinates revealed by 2D-EPR. J. Am. Chem. Soc. 2003, 125, 5915–5927. [DOI] [PubMed] [Google Scholar]
- 102.Przyojski JA; Arman HD; Tonzetich ZJ, NHC Complexes of Cobalt(II) Relevant to Catalytic C–C Coupling Reactions. Organometallics 2013, 32, 723–732. [Google Scholar]
- 103.Tripathi S; Vaidya S; Ansari KU; Ahmed N; Rivière E; Spillecke L; Koo C; Klingeler R; Mallah T; Rajaraman G; Shanmugam M, Influence of a Counteranion on the Zero-Field Splitting of Tetrahedral Cobalt(II) Thiourea Complexes. Inorg. Chem. 2019, 58, 9085–9100. [DOI] [PubMed] [Google Scholar]
- 104.Scholes CP; Lapidot A; Mascarenhas R; Inubushi T; Isaacson RA; Feher G, Electron nuclear double resonance (ENDOR) from heme and histidine nitrogens in single crystals of aquometmyoglobin. J. Am. Chem. Soc. 1982, 104, 2724–2735. [Google Scholar]
- 105.Scholes CP; Falkowski KM; Chen S; Bank J, Electron nuclear double resonance (ENDOR) of bis(imidazole) ligated low-spin ferric heme systems. J. Am. Chem. Soc. 1986, 108, 1660–1671. [Google Scholar]
- 106.Hoffman BM; Venters RA; Martinsen J, General Theory of Polycrystalline ENDOR Patterns. Effects of Finite EPR and ENDOR Component Linewidths. J. Magn. Reson. 1985, 62, 537–542. [Google Scholar]
- 107.Hoffman BM; Gurbiel RJ, Polycrystalline ENDOR Patterns from Centers with Axial EPR Spectra. General Formulas and Simple Analytic Expressions for Deriving Geometric Information from Dipolar Couplings. J. Magn. Reson. 1989, 82, 309–317. [Google Scholar]
- 108.Holland PL; Tolman WB, Three-Coordinate Cu(II) Complexes: Structural Models of Trigonal-Planar Type 1 Copper Protein Active Sites. J. Am. Chem. Soc. 1999, 121, 7270–7271. [Google Scholar]
- 109.Eckert NA; Dinescu A; Cundari TR; Holland PL, A T-Shaped Three-Coordinate Nickel(I) Carbonyl Complex and the Geometric Preferences of Three-Coordinate d9 Complexes. Inorg. Chem. 2005, 44, 7702–7704. [DOI] [PubMed] [Google Scholar]
- 110.Myers WK; Scholes CP; Tierney DL, Anisotropic Fermi Couplings Due to Large Unquenched Orbital Angular Momentum: Q-Band 1H, 14N, and 11B ENDOR of Bis(trispyrazolylborate) Cobalt(II). J. Am. Chem. Soc. 2009, 131, 10421–10429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Lloret F; Julve M; Cano J; Ruiz-García R; Pardo E, Magnetic properties of six-coordinated high-spin cobalt(II) complexes: Theoretical background and its application. Inorg. Chim. Acta 2008, 361, 3432–3445. [Google Scholar]
- 112.Tierney DL, Jahn–Teller Dynamics in a Series of High-Symmetry Co(II) Chelates Determine Paramagnetic Relaxation Enhancements. J. Phys. Chem. A 2012, 116, 10959–10972. [DOI] [PubMed] [Google Scholar]
- 113.Myers WK; Duesler EN; Tierney DL, Integrated paramagnetic resonance of high-spin Co(II) in axial symmetry: Chemical separation of dipolar and contact electron−nuclear couplings. Inorg. Chem. 2008, 47, 6701–6710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Brown TG; Hoffman BM, 14N, 1H, and metal ENDOR of single crystal Ag(II)(TPP) and Cu(II)(TPP). Mol. Phys. 1980, 39, 1073–1109. [Google Scholar]
- 115.For Ag(TPP), P(14N) = [−0.692, +0.915, (−0.222)] MHz. Even in this single-crystal study, the third (smallest) nqc component was not directly measured in Cu,Ag(TPP), but was obtained from the requirement that .
- 116.Kragskow JGC; Marbey J; Buch CD; Nehrkorn J; Ozerov M; Piligkos S; Hill S; Chilton NF, Analysis of vibronic coupling in a 4f molecular magnet with FIRMS. Nat. Commun. 2022, 13, 825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Brorson M; Schäffer CE, Orthonormal interelectronic repulsion operators in the parametrical dq model. Application of the model to gaseous ions. Inorg. Chem. 1988, 27, 2522–2530. [Google Scholar]
- 118.Use of ζ = 425 cm−1 (~80% of free-ion value) gives g′ = [5.93, 2.88, 1.76] and 2D* = 71 cm−1.
- 119.Atanasov M; Ganyushin D; Sivalingam K; Neese F, A Modern First-Principles View on Ligand Field Theory Through the Eyes of Correlated Multireference Wavefunctions. In Molecular Electronic Structures of Transition Metal Complexes II, Mingos DMP; Day P; Dahl JP, Eds. Springer Berlin Heidelberg: Berlin, Heidelberg, 2012; pp 149–220. [Google Scholar]
- 120.Bendix J; Brorson M; Schäffer CE, Accurate empirical spin orbit coupling parameters ζnd for gaseous ndq transition metal ions. The parametrical multiplet term model. Inorg. Chem. 1993, 32, 2838–2849. [Google Scholar]
- 121.Stoian SA; Vela J; Smith JM; Sadique AR; Holland PL; Münck E; Bominaar EL, Mössbauer and Computational Study of an N2-Bridged Diiron Diketiminate Complex: Parallel Alignment of the Iron Spins by Direct Antiferromagnetic Exchange with Activated Dinitrogen. J. Am. Chem. Soc. 2006, 128, 10181–10192. [DOI] [PubMed] [Google Scholar]
- 122.Sinnecker S; Slep LD; Bill E; Neese F, Performance of Nonrelativistic and Quasi-Relativistic Hybrid DFT for the Prediction of Electric and Magnetic Hyperfine Parameters in 57Fe Mössbauer Spectra. Inorg. Chem. 2005, 44, 2245–2254. [DOI] [PubMed] [Google Scholar]
- 123.Greer SM; Gramigna KM; Thomas CM; Stoian SA; Hill S, Insights into Molecular Magnetism in Metal–Metal Bonded Systems as Revealed by a Spectroscopic and Computational Analysis of Diiron Complexes. Inorg. Chem. 2020, 59, 18141–18155. [DOI] [PubMed] [Google Scholar]
- 124.Greer SM; Üngor Ö; Beattie RJ; Kiplinger JL; Scott BL; Stein BW; Goodwin CAP, Low-spin 1,1′-diphosphametallocenates of chromium and iron. Chem. Commun. 2021, 57, 595–598. [DOI] [PubMed] [Google Scholar]
- 125.Goodwin CAP; Giansiracusa MJ; Greer SM; Nicholas HM; Evans P; Vonci M; Hill S; Chilton NF; Mills DP, Isolation and electronic structures of derivatized manganocene, ferrocene and cobaltocene anions. Nature Chemistry 2021, 13, 243–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Greer SM; McKay J; Gramigna KM; Thomas CM; Stoian SA; Hill S, Probing Fe–V Bonding in a C3-Symmetric Heterobimetallic Complex. Inorg. Chem. 2018, 57, 5870–5878. [DOI] [PubMed] [Google Scholar]
- 127.Grant CV; Ball JA; Hamstra BJ; Pecoraro VL; Britt RD, 51V ESE-ENDOR Studies of Oxovanadium(IV) Complexes: Investigation of the Nuclear Quadrupole Interaction. J. Phys. Chem. B 1998, 102, 8145–8150. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
