Jin
Xiong
a,
Qing
Liu
b,
Barbara
Lavina
cd,
Michael Y.
Hu
c,
Jiyong
Zhao
c,
Esen E.
Alp
c,
Liang
Deng
*b,
Shengfa
Ye
*e and
Yisong
Guo
*a
aDepartment of Chemistry, Carnegie Mellon University, Pittsburgh, Pennsylvania 15213, USA. E-mail: ysguo@andrew.cmu.edu
bState Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai, 200032, P. R. China. E-mail: deng@sioc.ac.cn
cAdvanced Photon Source, Argonne National Laboratory, Argonne, Illinois 60439, USA
dCenter for Advanced Radiation Sources, University of Chicago, Chicago, Illinois 60439, USA
eState Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, P. R. China. E-mail: shengfa.ye@dicp.ac.cn
First published on 15th February 2023
High valent iron terminal imido species (FeNR) have been shown to be key reactive intermediates in C–H functionalization. However, the detailed structure–reactivity relationship in FeNR species derived from studies of structurally well-characterized high-valent FeNR complexes are still scarce, and the impact of imido N-substituents (electron-donating vs. electron-withdrawing) on their electronic structures and reactivities has not been thoroughly explored. In this study, we report spectroscopic and computational studies on a rare S = 1 iron(IV)–bisimido complex featuring trifluoromethyl groups on the imido N-substituents, [(IPr)Fe(NC(CF3)2Ph)2] (2), and two closely related S = 0 congeners bearing alkyl and aryl substituents, [(IPr)Fe(NC(CMe3)2Ph)2] (3) and [(IPr)Fe(NDipp)2] (1), respectively. Compared with 1 and 3, 2 exhibits a decreased FeNR bond covalency due to the electron-withdrawing and the steric effect of the N-substituents, which further leads to a pseudo doubly degenerate ground electronic structure and spin polarization induced β spin density on the imido nitrogens. This unique electronic structure, which differs from those of the well-studied Fe(IV)–oxido complexes and many previously reported Fe(IV)–imido complexes, provides both kinetic and thermodynamic advantages for facile C–H activation, compared to the S = 0 counterparts.
Compared to the progress in research on Fe(IV)O species, the progress on other FeE species is much less developed. Among these species, high valent FeNR species have garnered significant attention in fields such as enzyme bioengineering, organometallic chemistry and bioinorganic chemistry, as they are thought to be the key reactive intermediates in N2 reduction, C–H amination, olefin aziridation, and aminohydroxylation.1,5,27–31 Numerous FeNR complexes, including Fe(IV)NR, Fe(V)NR, and even Fe(VI)(NR)2 complexes, have been synthesized.32–60 However, understanding of the structure–reactivity relationship of iron imido species, derived from studies of structurally well-characterized high-valent FeNR complexes, is still limited.
Holland and coworkers have reported crystallographically characterized three-coordinated LFe(III)NAd complexes (L = a bulky β-diketiminate ligand; Ad = 1-adamantyl).35–38 Their studies on the HAT reactivity of these Fe(III)–NR have suggested that more exposed imido nitrogen (less steric effect) with high basicity may enhance HAT reactivity.35,36 Smith and coworkers have also suggested the importance of imido nitrogen basicity on HAT reactivity through studies on a crystallographically characterized four-coordinated Fe(IV)NAd complex supported by a tris(carbene)-borate ligand.61 Betley and coworkers have crystallographically characterized three-coordinated S = 5/2 dipyrrin supported iron(III)–imido (dipyrrin-Fe(III)NR (R = Ad, Mes)) complexes and one-electron oxidized four-coordinated iron(III)-iminyl (dipyrrin-FeCl(˙NR)) complexes that are capable of nitrene transfer and C–H amination.55,56 Comparative studies on the iron–imido/iminyl redox isomers suggested that a high-spin iron center with strong radical character on the imido nitrogen may improve reaction efficacy in the C–H activation and subsequent amination.56 These studies suggest that the HAT reactivity of iron–imido species can be promoted by imido nitrogen basicity, iron spin state, and imido nitrogen radical character, thus drawing parallels with iron–oxido species.
Yet, different from the iron–oxido species, iron–imido species display a unique structural feature, namely the presence of N-substituents of the imido ligand. However, the impact of these N-substituents on the electronic structures and reactivities of iron–imido species has not been systemically studied. Based on general chemical principles, electron donating substituents strengthen the FeNR bonding interactions, while aromatic or electron withdrawing substituents weaken these interactions by stabilizing the electron lone pair on the imido nitrogen. Such weakening would likely reduce the iron–imido ligand field strength and promote a higher spin state, which may enhance the HAT reactivity. So far, most reported iron–imido complexes bear electron donating or aromatic groups as N-substituents, such as tert-butyl, 1-adamantyl, 2,4,6-trimethylphenyl, 4-tert-butylphenyl, and 2,6-diisopropylphenyl (Dipp) substituents (some examples are shown in Scheme 1).32–60 Isolable iron–imido species bearing strong electron withdrawing N-substituents are rare.42,43,45,62–65 Recently, the first structurally characterized Fe(IV)–bisimido complex bearing a strong electron-withdrawing N-substituent, α,α-bis(trifluoromethyl)benzyl, was reported by some of us (Scheme 1).53 The complex, [(IPr)Fe(NC(CF3)2Ph)2] (2, IPr = 1,3-bis(2′,6′-diisopropylphenyl)imidazol-2-ylidene), exhibits an S = 1 ground spin state and readily undergoes intramolecular C–H bond dehydrogenation at room temperature via HAT. Interestingly, the replacement of CF3 groups in 2 to electron-donating CH3 groups on the N-substituents resulted in an S = 0 Fe(IV)–bisimido complex, [(IPr)Fe(NC(CH3)2Ph)2] (3), which was stable and showed no chemical reactivity even under elevated temperatures (30–70 °C). Additionally, the use of aromatic N-substituents in another similar Fe(IV)–bisimido complex, [(IPr)Fe(Ndipp)2] (1), which would in principle also weaken the FeNR bond and promote higher spin state, still yielded an S = 0 spin ground state.48 Thus, complexes 1–3 provide a great opportunity to study the effect of the imido N-substituents on the structure and chemical reactivity of Fe(IV)–imido complexes.
Scheme 1 The Fe(IV)–bisimido complexes studied in this work and examples of the previously reported S = 1 Fe(IV)–imido complexes. |
Herein, we present a comprehensive spectroscopic and computational analysis of the novel S = 1 complex 2 and its S = 0 congeners, 1 and 3. Results from 57Fe nuclear resonance vibrational spectroscopy (NRVS), 57Fe Mössbauer spectroscopy, and direct current (dc) magnetic measurements showed that complex 2 exhibits strongly reduced Fe(IV)NR bond covalency compared with those of 1 and 3, which leads to distinctive magnetic properties due to its large unquenched orbital angular momentum. We also used the complete-active-space self-consistent field (CASSCF)66 method combined with N-electron valence state perturbation theory (NEVPT2)67,68 and DFT calculations to understand the origin of these unique electronic and magnetic properties in 2 and compare them to complexes 1 and 3. Our findings indicate that the reduction of the Fe(IV)NR covalency in 2 is not only due to the electron-withdrawing nature of the N-substituents but also due to their steric effects. These two features are not simultaneously present in 1 and 3. Furthermore, the reduced Fe(IV)NR covalency in 2 promotes spin polarization that induces β spin density on the imido nitrogens, which facilitates an α spin transfer in the HAT process and is fundamentally different from those in Fe(IV)O species as well as other Fe(IV)NR species in a pseudo-C4 ligand field. This unique electronic structure of 2 provides both kinetic and thermodynamic advantages to enable facile C–H activation.
Complex 1 | Complex 2 | Complex 3 | |
---|---|---|---|
Bond length (Å) | |||
FeN(imido) | 1.638, 1.638 | 1.705, 1.712 | 1.614, 1.616 |
Fe–C(NHC) | 1.916 | 2.049 | 1.928 |
N–C(R) | 1.376, 1.376 | 1.438, 1.435 | 1.440, 1.455 |
Bond angle ( o ) | |||
FeN–C(R) | 172.68, 172.68 | 144.05, 140.84 | 160.16, 161.13 |
NFeN | 142.98 | 125.28 | 127.73 |
NFe–C(NHC) | 108.51, 108.51 | 121.22, 113.38 | 115.59, 116.67 |
Torsion angle ( o ) | |||
N(imido)Fe–C(NHC)–N(NHC) | 35.03, 35.03 | 78.03, 83.62 | −5.79, −2.54 |
C19–N2⋯N3–C23 | −25.41 | −69.69 | 0.98 |
Ref. | This work | Ref. 53 | Ref. 53 |
Fig. 2 NRVS-derived 57Fe PVDOS spectra of complexes 1 and 2 and selected normal modes of vibrations. Experimental (black) and DFT-calculated (red) 57Fe PVDOS spectra of 1 (A) and 2 (B) and the symmetric and asymmetric stretching modes of RNFeNR moiety (C) are shown. The frequencies of major spectral features and the mode assignments based on a simplified MX(Y2)2-type molecule are indicated in the figure (see Fig. S4† for mode pictures). The 57Fe PVDOS intensity is shown on the left Y axis while the intensity of calculated mode decomposition factor, e2, is shown on the right Y axis. The mode pictures in (C) are derived from the DFT frequency calculations of the full structure of complex 1. The red arrows represent the direction and the magnitude of the atom displacements (>0.07 Å) in the corresponding normal mode of vibration. |
To gain a better understanding of the 57Fe PVDOS spectra, we performed DFT calculations. Several functional-basis set combinations were used in the calculations, and all of them reproduced the experimental data reasonably well (Fig. S2 and S3†). Here, the results from the BP86/TZVP72,73 with GD3 correction74 are presented in Fig. 2. The DFT optimized structure for complex 1 in an S = 0 state and for complex 2 in an S = 1 state matched well with their crystal structures (Table S2 and Fig. S1†). Based on the DFT optimized structures, the predicted 57Fe PVDOS spectra of complexes 1 (S = 0) and 2 (S = 1) derived from frequency calculations also aligned well with the experimental data, allowing us to confidently assign the vibrational modes (see Tables 2 and S4† for a list of vibrational mode assignments and the ESI† for additional analysis and discussion). One key feature revealed by the DFT analysis is that the 57Fe PVDOS features at ∼990 cm−1 and ∼1285 cm−1 belong to the vibrations of the MN–R moieties. The very weak 57Fe PVDOS features at 1285 cm−1 and 1322 cm−1 (DFT predicted at 1302 and 1350 cm−1) can be assigned to the asymmetric N–R stretching modes between the N and C atoms of the imido ligand (νas(FeN–R)), which are coupled with different imido substituent vibrations (in this case Dipp vibrations). Due to the short Fe–imido bonds, the asymmetric N–R stretching mode also mixes slightly with the asymmetric FeNR stretching mode (νas(FeNR)) to cause the movement of the iron center, thereby rendering their detection. The main νas(FeNR) is located at 988 cm−1 (995 cm−1 in DFT). The lack of appreciable iron movement in the symmetric FeNR stretching mode (νs(FeNR)) results a weak shoulder at ∼970 cm−1 (968 cm−1 in DFT) for this vibrational mode. In addition, both νas(FeNR) and νs(FeNR) show strong coupling with aromatic ring vibration from the Dipp substituent (Fig. 3). Overall, the vibrational features of the MN–R moieties in complex 1 are consistent with many reported metal–imido complexes studied by IR and resonance Raman (Table 2).75 Furthermore, the NRVS data provided a definitive confirmation for the assignment of the FeNR stretching mode and the FeN–R stretching mode, as the former mode leads to stronger iron movement thus higher 57Fe PVDOS intensity than the latter. This is clearly observed in the current NRVS data, showing a high intensity 988 cm−1 (νas(FeNR)) and a low intensity 1285 cm−1 (νas(FeNR)) features.
Complex | d(FeNR) | ν(FeNR) | ν as(N–R) | Ref. |
---|---|---|---|---|
a The italic numbers are derived from DFT calculations. Distances are in Å, and vibrational frequencies are in cm−1. | ||||
1 | 1.638 | 968 (968) s | 1285, 1322 (1302, 1350) as | This work |
988 (995) as | ||||
2 | 1.709 | 814 (826) s | 1139, 1173 (1060, 1070, 1155, 1180) | This work |
858 (861) as | ||||
[FeIV(NTs)(MePy2tacn)]2+ | 1.71 | 984 | — | 46 |
[FeIV(NTs)(Me2CHPy2tacn)]2+ | 1.72 | 1061 | — | 46 |
[FeIV(NTs)(N4py)]2+ | 1.73 | 998 | — | 43 |
[(TAML)FeV(NTs)]− | 1.65 | 817 | — | 42 |
[PhBP3]FeIIINtBu | 1.635 | 1104 | 1233 | 75 |
[PhBP3]FeIIIN(1-Ad) | 1.641 | 1097 | 1225 | 75 |
[PhBP3]FeIIINPh | — | 958 | 1292/1309 | 75 |
[PhBP3]FeIIIN(p-tolyl) | 1.659 | 962 | 1281/1305 | 75 |
[Fe4(μ3-NtBu)4(NtBu)Cl3] | 1.635 | 1111 | 1214 | 75 |
For complex 2, all the vibrational modes identified in complex 1 are shifted to lower frequencies due to the elongation of all the iron–ligand distances in going from 1 to 2 (Fig. 2). In particular, the νas(FeNR) is now located at 858 cm−1 (861 cm−1 in DFT), the νs(FeNR) is located at 814 cm−1 (826 cm−1 in DFT). Thus, the FeNR stretching frequency observed for 2 is one of the lowest reported in the literature (Table 2), which is consistent with the relatively long FeNR bond length in 2 (1.709 Å). In addition, we compared the displacement of the Fe atom and the N atoms from the bisimido ligands in different normal modes and found that they exhibit a very strong correlation (Fig. S5†). As such, the 57Fe PVDOS spectra of complexes 1 and 2 are dominated by the vibrations associated with Fe–bisimido moiety, which is consistent with the strong FeNR bonding interactions in the Fe–bisimido moiety of these two complexes, particularly for complex 1. Overall, the significantly red-shifted vibrational frequencies of the FeNR and FeN–R stretching modes observed in 2 than those in 1 correlate well with the long FeNR bonds in 2, which further suggests the strongly reduced FeNR covalency in this complex.
The VTVH Mössbauer spectra of complex 2 are shown in Fig. 3. The magnetization of the 4.2 K Mössbauer spectra under various applied field is typical of a paramagnetic species with an integer spin ground state, a large negative D (D ≪ 0), and an easy magnetization direction (an internal field only rapidly develops along the z direction defined by the D tensor, the relative orientations of the D, g and electric field gradient (EFG) tensor (V) with respect to the molecular frame will be discussed in the electronic structure calculation section below).76,77 For such a uni-axial magnetic system, the Mössbauer spectra measured at 4.2 K only reflect the properties of the ground non-Kramers doublet of the spin manifold, which can be approximated as an effective Seff = 1/2 spin Hamiltonian with geff,x = geff,y ≈ 0, geff,z ≫ 0 and an off-diagonal matrix element of the effective spin Hamiltonian Δ/2 (to account for the zero field splitting (Δ) of the magnetic sublevels, see Fig. S7†).76 With this effective spin Hamiltonian, the VTVH Mössbauer spectra of complex 2 can be satisfactorily simulated (Fig. S8†), which affirms that the ground state electronic structure of 2 indeed features an isolated pseudo-doublet and the excited spin sublevels are not significantly populated even at 100 K.
Overall, the spectral simulations by using the effective spin Hamiltonian revealed two major features: (1) the Mössbauer spectra reveals that a large positive internal field exists in 2 (Bint = −Aeff,z〈Seff,z〉/gnβn ∼ +28 T at 7 T applied field), which indicates the internal field is dominated by the orbital contribution due to the unquenched orbital angular momentum; (2) the largest principal component of the EFG tensor (Vz′′z′′) is perpendicular to the direction of the internal field. The obtained effective spin Hamiltonian parameters are Δ = 13.6 cm−1, geff,z = 5.6, Aeff,z/gnβn = +71 T, ΔEQ = −1.02 mm s−1, η = 0, Vz′′z′′⊥geff,z, and δ = −0.11 mm s−1. These parameters can be mapped to a canonical S = 1 spin Hamiltonian with a large negative D (see Fig. S7†). We can use the |±1′〉 doublet of an S = 1 spin Hamiltonian to mimic the two-level system described by the effective Seff = 1/2 spin Hamiltonian. Thus, Δ, geff,z, and Aeff,z from the effective Hamiltonian correspond to E, gz′z′ and Az′z′ of the S = 1 Hamiltonian, giving E = −6.8 cm−1, gz′z′ = 2.80, Az′z′/gnβn = +35.5 T (note that x′, y′, and z′ are used to label the x, y and z direction of the D, g and A tensors, and to distinguish them from those of the molecular frame coordination, as explained in the electronic structure calculation section). Using these parameters together with the assumption that the principal axes of A and D are colinear (Az′z′∥Dz′z′), we were able to obtain the Mössbauer parameters that reproduce all the spectroscopic data of complex 2 with an S = 1 spin Hamiltonian (by fixing D = −79 cm−1 and g = [2, 2, 2.80]): δ = −0.11 mm s−1, ΔEQ = −1.02 mm s−1, η = 0 with Vz′′z′′ ⊥ Dz′z′, E/D = 0.085, Az′z′/gnβn = +35.5 T with Az′z′∥Dz′z′ (Fig. 3).
Dc magnetic measurements were also carried out to further investigate the magnetic anisotropy of 2 (see the analysis in the ESI and Fig. S9† for details). The effective magnetic moment (μeff) of ∼3.1μB measured under 300 K and 1 T is larger than spin-only value for a triplet system (2.832μB). The VTVH susceptibilities and magnetizations can be well reproduced with the following parameters: S = 1, D = −79 cm−1, E/D = 0.077, g⊥ = 1.91, g∥ = 2.73, TIP = 1188 × 10−6 emu (TIP refers to temperature-independent paramagnetism). The large D and TIP values are indicative of low-lying excited states and unquenched angular momentum.
Taken together, the magnetic susceptibility/magnetization measurements and the Mössbauer analysis reveal an orbitally nearly degenerate electronic ground state in 2. In the literature, there are several reported iron complexes exhibiting such an orbitally nearly degenerate electronic structure, but they are either iron(I) or iron(II) species.47,78–81 To the best of our knowledge, high-valent iron complexes, such as complex 2, featuring such a unique electronic structure have not been reported yet.
The ab initio calculations show that 1 has a diamagnetic ground state with a leading electron configuration of , which only accounts for ∼73% of the wavefunction, indicating the multi-reference character of the ground state (as shown in Fig. 4). In the leading configuration, dz2 is the doubly occupied non-bonding orbital, while the other four Fe d orbitals participate in π bonding and antibonding interactions with py and pz orbitals from the two imido ligands with as the other doubly occupied orbital. The calculations yield ΔEQ = −1.75 mm s−1, η = 0.49, which are consistent with the Mössbauer simulation for 1, and the EFG tensor frame is well compatible to the pseudo C2v symmetry, with Vz′′z′′ aligning with the z-axis of the molecular frame defined in Fig. 4, Vx′′x′′ aligning with y-axis, and Vy′′y′′ with x-axis (Fig. S11†).
Multi-reference calculations were also performed on 3, revealing a similar electron ground state as that of complex 1. Specifically, the leading configuration is the same as 1, contributing ∼75% of the ground state wavefunction. The bonding orbitals πxy and πx2−y2 are heavily mixed and undistinguishable, but the corresponding anti-bonding pairs are well-defined (Fig. S10†). The EFG tensor principal frame also follows the same orientation as that in 1 (Fig. S11†), with the calculated ΔEQ and η to be −1.81 mm s−1 and 0.94, respectively.
The wavefunction of the triplet ground state of 2 exhibits even more substantial multi-reference character, with the leading configuration contributing only 59% of the total wavefunction (Fig. 4). This configuration differs greatly from the ground state electron configuration of 1 in that not only a single electron is promoted from to to lead to a triplet ground state for 2, more importantly, instead of exhibiting antibonding nature for and in 1, these two corresponding orbitals are essentially non-bonding Fe dyz and dxz orbitals in 2 based on orbital decomposition (Fig. 4). This reflects the significant reduction of covalent interactions between iron and the imido ligands in 2. With three non-bonding orbitals (dz2, dyz, and dxz) and four 3d electrons, a triplet ground state is favored for 2 based on the first Hund's rule. In addition, the energies of these three non-bonding orbitals are likely to be close, thus leading to an orbital (pseudo)degeneracy. This is confirmed by the excited state calculations. The first triplet excited state with a leading configuration (46%) of lies only 1020 cm−1 above the ground state. As such, the first excited state primarily corresponds to a single electron excitation from the dz2 non-bonding orbital of the ground state to the dyz non-bonding orbital, thereby leading to a sizeable unquenched orbital angular momentum along the x-direction. Higher excited states have much larger energy separations from the ground state (Table S4†), thus their spin orbit coupling (SOC) with the ground state is much less pronounced. The calculations give the value of the largest principal component of the g tensor as 2.78, which is much larger than 2, and a large negative D value of −79.2 cm−1. The electronic structure revealed by CASSCF calculations is in agreement with the one reflected by the spectral simulation model: the triplet ground state of complex 2 exhibits a large negative zero-field splitting and strong easy axis magnetic anisotropy. More importantly, the estimated energy separation of 16.2 cm−1 for the lowest-energy doublet from the calculations is in a reasonable agreement with 2|E| of 13.6 cm−1 derived from the spectroscopic simulations. Furthermore, based on the calculations, the direction of the largest principal components (Dz′z′ and gz′z′) of the D and the g tensors, which reflects the direction of unquenched orbital angular momentum, is along the x-direction (Fig. 5). This is consistent with SOC generated by the dyz/dz2 pair as described above. Remarkably, the calculations also successfully reproduce the relative orientation between V and D principal axes, with angles of 172° between Vz′′z′′ and Dy′y′, and 97° between Vz′′z′′ and Dz′z′, which is consistent with the Mössbauer analysis of 2.
Finally, the spin density plot shown in Fig. 5 reveals a sizeable negative spin density primarily located in the equatorial plane on two imido N atoms (i.e. the RNFeNR coordination plane). As elaborated in the previous studies for related high-valent iron–nitrido and –oxido complexes,83,84 the negative spin density is the result of the spin polarization of the formally doubly occupied πxy, and πx2−y2 orbitals induced by the two unpaired electrons resided in the non-bonding Fe dxz and dyz orbitals. As a manifestation, the occupation numbers of formally doubly occupied πxy, and πx2−y2 are considerably less than 2, while those of their correspond antibonding orbitals ( and ) are substantially deviated from 0. The drastically reduced covalent interactions between Fe and two imido ligands render the two singly occupied orbitals (Fe dxz and dyz) essentially non-bonding, which leads to a more pronounced spin polarization effect compared to well-studied Fe(IV)–oxido complexes. In the latter complexes Fe is involved in strong covalent interactions with the oxido ligand so that the singly occupied of the Fe dxz parentage and of the Fe dyz parentage contains substantial contributions from the O-px and O-py atomic orbitals, thus leading to sizable positive spin population in the O-px and O-py orbitals and a positive spin density on the oxido ligand. Thus the overall bonding interactions of Fe(IV)NR moieties in 2 are fundamentally distinct from the well-studied Fe(IV)O species and the other reported Fe(IV)NR complexes having a pseudo C4 symmetry. The differences of frontier molecular orbitals (MOs) in complex 1 (or 3) and 2 derived from CASSCF calculations that show strong covalent interactions between the iron and the imido ligand, as well as those from the well-studied Fe(IV)–oxido complexes are shown in Scheme 2.
Overall, the CASSCF calculations for complexes 1–3 successfully reproduced the experimentally determined magnetic spectroscopic parameters. However, since such type of calculations cannot predict Mössbauer isomer shifts, so DFT calculations were performed on the full complex structure to predict this spectroscopic parameter (see the ESI† for detailed computational method). The DFT optimized structures closely resembled the crystal structures of the three complexes (see Table S2 and Fig. S1†). Based on these, isomer shifts were further calculated. Although, the calculated isomer shift showed a systematic negative shift of ∼0.1 mm s−1 from the experimental values based on a published isomer shift calibration method,85 the overall trend is consistent with the experimental observations, giving a highest isomer shift for complex 2 (−0.19 mm s−1, DFT vs. −0.10 mm s−1, exp), a lowest isomer shift for complex 3 (−0.53 mm s−1, DFT vs. −0.41 mm s−1, exp), and an isomer shift of −0.42 mm s−1 (DFT, vs. −0.28 mm s−1, exp) for complex 1 (see Table S5† for the comparison of the calculated and the experimental Mössbauer parameters).
According to the CASSCF calculations, the orbital overlap of natural orbitals (NOs) between Fe and the two N(imido) atoms in complexes 1 and 3 is generally excellent as shown in Fig. 4 and S10,† and Scheme 2A. However, in complex 2, the NOs overlap among these three atoms is significantly reduced due to the unfavorable relative orientation of the Fe 3d and the N(imido) p orbitals (Fig. 4 and Scheme 2B). The Fe 3d orbitals have a good π overlap with the p orbital from only one of the N atoms, leaving a poor σ overlap with the p orbital from the other N atom, thus leading to essentially non-bonding Fe dxz and dyz orbitals (see the and the pair in Fig. 4). Therefore, structural factors that reduce the effective orbital overlap between Fe and the two N atoms are key to stabilize an S = 1 state.
Two structural differences between complexes 1 (or 3) and 2 most likely affect the orbital overlap between Fe and the two N atoms, namely the FeN(imido) bond length (1.638 Å in 1, 1.615 Å in 3, and 1.709 Å in 2) and the dihedral angle between the two planes formed by the two Fe–N–C(imido) moieties (C19–N2⋯N3–C23 dihedral angle, −25° in 1, 1° in 3, and −70° in 2). The tendency to form longer FeN bonds in 2 was suggested to be originated from the use of an electron-withdrawing N-substituent, α,α-bis(trifluoromethyl)benzyl. Our DFT calculations support this hypothesis. When using different N-substituents to carry out DFT geometry optimizations on the truncated model, it is evident that alkyl electron-withdrawing N-substituents tend to elongate the FeN bonds, regardless of the spin state used, when compared with N–Me (Fig. 6, Table S2†). At the same time, the aryl N-substituents tend to elongate the FeN bonds in the singlet state and to a lesser extend in the triplet state. These observations are consistent with the understanding that both electron-withdrawing alkyl and aryl substituents can stabilize the lone pair electrons on the p orbitals of the imido N, thereby weakening the FeN bonds. However, based on the DFT calculations on the truncated models, this effect only elongates the FeN bonds by a small amount (<0.03 Å), and may not promote the spin state conversion on its own.
Fig. 6 Imido-substituent-dependent deviation of average Fe–N(imido) distance relative to that of the methyl substituent obtained from DFT optimization. The substituents on NHC side are two methyl groups in all cases presented here. Full data can be found in Table S2.† Abbreviations: Ad = adamantyl, tBu = tert-butyl, Cyh = cyclohexyl, RFC(CF3)3, Ms = methylsulfonyl, Vn = vinyl, Ph = phenyl. |
The second structural feature that affects the orbital overlap between Fe and the two N atoms is the C19–N2⋯N3–C23 dihedral angle (see Fig. 1 for atom labeling). For complex 2, this angle is −70°, which leads to overlap mismatch between the Fe 3d orbitals and N(imido) p orbitals, particularly between dxz and pz of N(imido), and between dxy and py of N(imido), thereby weakening the covalent interactions between Fe and the imido ligands (Scheme 2). This angle is likely to be controlled by steric interactions of the substituents. To provide evidence, we performed geometric scans along the C19–N2⋯N3–C23 dihedral angle on three truncated models, Me_Me, Me_tBu, and Me_CF3, in the S = 1 state. During scans, the values of other selected key bond lengths and angles were fixed at the crystal structure values from complex 2 (see the ESI† for details). The use of N–Me led to the identification of two local energy minima, with the dihedral angle of ±70° (Fig. 7), which is close to the angle observed in complex 2. Compared with a 0 degree dihedral angle, the 70 degree angle was favored by ∼2 kcal mol−1. When the bulkiness of the N-substituent was increased from N–Me to N–tBu, 65 degree dihedral angle was favored, with stronger stabilization (by ∼8 kcal mol−1). For N–CF3 substituent, 85 degree dihedral angle was favored, with stabilization of ∼7 kcal mol−1. Therefore, regardless of the nature of N-substituent (electron withdrawing vs. electron donating), its bulkiness favored a large C19–N2⋯N3–C23 dihedral angle, thus contributing to the mismatch of the orbital overlap between Fe and the two N atoms. Based on this analysis, we reason that the bulkier –CF3 moieties in 2 compared to the –CH3 moieties in 3 likely provide the needed steric effect to favor the large C19–N2⋯N3–C23 dihedral angle (−70° in 2vs. 1° in 3, Fig. 1). The aryl N-substituents (Dipp) in complex 1, although bulky, form good π–π interactions with the Dipp substituents on the NHC ligand, thus preventing the formation of a large C19–N2⋯N3–C23 dihedral angle (−25° in 1, Fig. 1).
Overall, the DFT results suggest that the combination of the N-substituent's steric properties and electron-withdrawing nature reduce the covalent interactions between Fe and the imido ligands in 2. As also revealed by the CASSCF calculations, this reduction in covalent interactions results in three close lying non-bonding Fe 3d based orbitals (dz2, dyz, and dxz), which stabilize an S = 1 ground spin state and lead to an orbital pseudo-degeneracy for 2.
According to Bell–Evans–Polanyi principle,86,87 the driving force for C–H activation via HAT is determined by the difference in the bond dissociation free energies (BDFEs) between the C–H bond activated and the N–H bond formed in the resulting Fe(III)–NHR species in our case. In the Fe(IV)–bisimido complexes studied here, since the target C–H bond is from the same Dipp substituent, the increased driving force of HAT for complex 2 compared to the S = 0 complexes is likely only due to the stronger N–H BDFE in the resulting Fe(III)–NHR species derived from 2. This is supported by our DFT calculations. The N–H BDFE is highest for the Fe(III)–NHR species derived from complex 2, which is ∼15 kcal mol−1 and ∼10 kcal mol−1 greater than the N–H BDFEs of the Fe(III)–NHR species derived from complexes 1 and 3, respectively (Fig. S16†). To understand the differences in N–H BDFEs, we used the classical thermodynamic cycle, where the diagonal direction represents the single HAT step and the edges represent two stepwise processes: a sequential proton-transfer (PT) to the Fe(IV)–(NR)2 species followed by an electron transfer (ET) to form the final Fe(III)–NHR species or an ET step followed by a PT step to form the same final state (Fig. S15†). The free energies of the PT and the ET steps can be described by pKa values and reduction potentials (E0), respectively. By adopting a thermodynamic analysis proposed by Srnec and coworkers,88 we performed DFT analysis (see the ESI, Fig. S16, S17, and Tables S6, S7† for details). By defining an effective pKa (pKa,eff, an averaged pKa value from the pKa values of the two PT steps in the thermodynamic scheme) and an effective reduction potential (E0eff, an averaged E0 value from the two ET steps), we showed that although complex 2 had a reduced pKa,eff compared to those of complexes 1 and 3 due to the use of the electron-withdrawing substituent (∼5 and ∼7 units reduction calculated in the benzene solvent or 266 mV and 440 mV reduction when converting to electric potential unit), this withdrawing effect led to a significant increase of E0eff in 2 compared to the other two complexes (613 mV and 882 mV increase respectively). Therefore, the higher reduction potential of complex 2 is likely the dominant factor determining the stronger N–H BDFE of the resulting Fe(III)–NH species than those derived from 1 and 3, thus making 2 more reactive than the S = 0 complexes.
In addition to the HAT driving force, the kinetic barrier of the HAT reaction is equally important. Our CASSCF and DFT calculations suggest that the unique ground state electronic and geometric structures of complex 2 may directly contribute to the lowering of the reaction barrier by mimicking the features of the transition state in the HAT step. The ground state and the transition state structures of complex 2 are highly similar, featuring a large β spin density on the p orbital of imido nitrogen in the RNFeNR coordination plane (Fig. 5, 8, and S18†). This p orbital, which is involved in bonding/antibonding interactions with Fe dxy/dx2−y2 orbitals (see the and the pair in Fig. 4), points toward the C–H bond to be activated on the Dipp substituent of NHC ligand, providing the necessary orbital overlap to facilitate the initial electron transfer from the C–H bond to the Fe(IV)NR moiety in the HAT process (Fig. 5 and 8). In addition, the transition state spin density distribution reveals that the initial electron transfer may occur via an α spin transfer since the carbon of the C–H bond being activated bears a β spin density (Fig. 8). This α spin transfer may be facilitated by the β spin density on the p orbital of imido nitrogen (Fig. 8). The α spin transfer channel of HAT in 2 resembles the σ pathway in the well-studied Fe(IV)O complexes, where due to the positioning of the C–H bond relative to the Fe(IV)O bond vector an α spin transfer to the σ* MO with Fe dz2 parentage occurs in the so-called σ pathway. Additionally, in Fe(IV)O complexes, a β spin transfer to the singly occupied π* MO with Fe dxz/dyz parentage is also possible, which occurs in the π pathway (Schemes 2 and S1†). This dual-channel reactivity does not likely to exist in 2, since the singly occupied frontier MOs in 2 are essentially non-bonding dxz/dyz orbitals (Fig. 4), which exhibit minimal orbital overlap with the C–H bond, thus preventing an efficient β spin transfer during the HAT process. Thus, 2 likely carries out HAT via a single α spin transfer channel (Fig. 8).
In summary, the facile C–H activation reactivity of S = 1 complex 2 correlates well with its electronic structure, which features a β spin density on the imido nitrogen due to the reduced covalent interactions between Fe and the imido ligands and the strengthened spin polarization effect, an exposed N p orbital of the imido ligand towards the C–H bond due to bent and twisted FeN–R angle and C–N⋯N–C dihedral angle, and an electron-withdrawing N-substituent.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2172442. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2sc06273a |
This journal is © The Royal Society of Chemistry 2023 |