Gerard
Sabenya
a,
Ilaria
Gamba
a,
Laura
Gómez
a,
Martin
Clémancey
b,
Jonathan R.
Frisch
c,
Eric J.
Klinker
c,
Geneviève
Blondin
b,
Stéphane
Torelli
b,
Lawrence
Que
Jr
*c,
Vlad
Martin-Diaconescu
*ad,
Jean-Marc
Latour
*b,
Julio
Lloret-Fillol
*ade and
Miquel
Costas
*a
aInstitut de Química Computacional i Catàlisi (IQCC), Departament de Química, Universitat de Girona, Campus Montilivi, E17071 Girona, Spain. E-mail: miquel.costas@udg.edu
bUniv. Grenoble-Alpes, CNRS, CEA, IRIG, DIESE, CBM, Grenoble 38000, France
cDepartment of Chemistry, University of Minnesota, Pleasant Str 207, Minneapolis, Minnesota, USA
dInstitute of Chemical Research of Catalonia (ICIQ), The Barcelona Institute of Science and Technology, Avinguda Països Catalans 16, 43007 Tarragona, Spain. E-mail: jlloret@iciq.es
eCatalan Institution for Research and Advanced Studies (ICREA), Passeig Lluïs Companys, 23, 08010, Barcelona, Spain
First published on 20th August 2019
High valent iron species are very reactive molecules involved in oxidation reactions of relevance to biology and chemical synthesis. Herein we describe iron(IV)–tosylimido complexes [FeIV(NTs)(MePy2tacn)](OTf)2 (1(IV)NTs) and [FeIV(NTs)(Me2(CHPy2)tacn)](OTf)2 (2(IV)NTs), (MePy2tacn = N-methyl-N,N-bis(2-picolyl)-1,4,7-triazacyclononane, and Me2(CHPy2)tacn = 1-(di(2-pyridyl)methyl)-4,7-dimethyl-1,4,7-triazacyclononane, Ts = Tosyl). 1(IV)NTs and 2(IV)NTs are rare examples of octahedral iron(IV)–imido complexes and are isoelectronic analogues of the recently described iron(IV)–oxo complexes [FeIV(O)(L)]2+ (L = MePy2tacn and Me2(CHPy2)tacn, respectively). 1(IV)NTs and 2(IV)NTs are metastable and have been spectroscopically characterized by HR-MS, UV-vis, 1H-NMR, resonance Raman, Mössbauer, and X-ray absorption (XAS) spectroscopy as well as by DFT computational methods. Ferric complexes [FeIII(HNTs)(L)]2+, 1(III)–NHTs (L = MePy2tacn) and 2(III)–NHTs (L = Me2(CHPy2)tacn) have been isolated after the decay of 1(IV)NTs and 2(IV)NTs in solution, spectroscopically characterized, and the molecular structure of [FeIII(HNTs)(MePy2tacn)](SbF6)2 determined by single crystal X-ray diffraction. Reaction of 1(IV)NTs and 2(IV)NTs with different p-substituted thioanisoles results in the transfer of the tosylimido moiety to the sulphur atom producing sulfilimine products. In these reactions, 1(IV)NTs and 2(IV)NTs behave as single electron oxidants and Hammett analyses of reaction rates evidence that tosylimido transfer is more sensitive than oxo transfer to charge effects. In addition, reaction of 1(IV)NTs and 2(IV)NTs with hydrocarbons containing weak C–H bonds results in the formation of 1(III)–NHTs and 2(III)–NHTs respectively, along with the oxidized substrate. Kinetic analyses indicate that reactions proceed via a mechanistically unusual HAT reaction, where an association complex precedes hydrogen abstraction.
Herein we describe two novel octahedral iron(IV)–tosylimido complexes using pentadentate ligands based on the 1,4,7-triazacyclononane macrocycle. The reactivity of these complexes in model N-transfer and HAT reactions has been explored, putting forward novel mechanistic scenarios, distinct from those previously established for analogous iron(IV)–oxo complexes and 3(IV)NTs complexes. Overall, the current study discloses unprecedented single electron mechanistic scenarios for high-valent iron compounds.
Scheme 1 Chemical strategy for the generation of 1(IV)NTs and 2(IV)NTs from iron(II) precursors. 1(II) and 2(II) (S = MeCN, CF3SO3). |
High resolution mass spectra (HR-MS) evidence the proposed formulation of 1(IV)NTs and 2(IV)NTs, and both show a major peak at m/z = 275.09, with an isotopic pattern that corresponds to [FeIV(NTs)(L)]2+, (L = MePy2tacn and Me2(CHPy2)tacn) (Fig. 1 and S1†). The spectra also show a lower intensity peak at m/z = 699.13, which corresponds to monocationic [[FeIV(NTs)(L)](OTf)]+. The later peaks shift by one mass unit upon formation of 1(IV)NTs and 2(IV)NTs with PhI(15N)Ts. Close inspection of these peaks reveals small contributions of the respective [FeIII(HNTs)(L)]2+ species (∼10%). Finally, a minor peak at m/z = 550.18 and its isotopic pattern are consistent with a [FeIII(NTs)(L)]+ formulation.
The 1H-NMR spectra of 1(IV)NTs and 2(IV)NTs in CD3CN or D6-acetone show paramagnetically shifted signals between −75 and 50 ppm, with the signals of the pyridine moiety as the most distinctive features due to their relative sharpness (Fig. 2). These could be identified by comparison with the spectra of the parent iron(IV)–oxo complexes, and also with that of [FeIV(X)(N4Py)]2+3(IV)X (X = O, NTs) and [FeIV(O)(Bn-TPEN)]2+, 4(IV)O (Bn-TPEN, N-benzyl-N,N′,N′-tris(2-pyridylmethyl)-1,2-diaminoethane).30,51 Moreover, the presence of broad signals for both 1(IV)NTs and 2(IV)NTs in the high-field region of the spectra, which are not present in the spectrum of 3(IV)NTs, suggests that they belong to the N-methyl and N-methylene protons on the triazacyclononane macrocycle. A bidimensional COSY spectrum of 2(IV)NTs (inset of Fig. 2), allowed the assignment of pyridine hydrogens to the signals at 20.87 (β), 9.70 (γ) and −6.56 (β′) ppm. Of note, the signal pattern closely resembles that of 3(IV)X (X = O, NTs), where all the pyridine rings are also parallel to the FeO/N axis. On the other hand, 1(IV)NTs presents a more complex spectrum, in accordance with its reduced symmetry, where one pyridine is parallel to the FeN axis and the other pyridine ring perpendicular to the FeN axis of this complex. The 1H-NMR spectrum and corresponding COSY data identify two distinct sets of pyridine protons. Subset “a” is assigned to the pyridine parallel to the FeN axis, showing the same pattern as observed for 2(IV)NTs and 3(IV)NTs, with the βα proton signal downfield shifted to 41.0 ppm, the proton signal upfield shifted to −13.3 ppm, and the γa signal at 12.7 ppm. However, subset “b” belongs to the pyridine perpendicular to the FeN axis and shows a distinct pattern, with both βb and proton signals upshifted to −5.6 and −6.9 ppm and the γb proton at 6.0 ppm, consistent with the pattern observed for the perpendicular pyridine previously reported for 4(IV)O.51
Fig. 2 1H-NMR and COSY spectra of 1(IV)NTs (top) in acetone-d6 at 0 °C and 2(IV)NTs (bottom) in acetone-d6 at 25 °C; small peaks at 32 and −9 ppm belong to <5% of 2(IV)O. |
The range of chemical shifts observed for the pyridine protons in this pair of complexes reflects differences in the orientation of the pyridine ring relative to the FeNTs axis. These shifts can be understood within the context of trends recently identified by Rasheed et al. in a survey of NMR data for a group of nonheme iron(IV)–oxo complexes, in which the paramagnetic shifts observed are found to be inversely dependent on the size of the torsion angle between the pyridine plane and the FeIVO unit (Table 1).52
Py ring to FeX angle | β | γ | β′ | Ref. | |
---|---|---|---|---|---|
a Torsion angles based on the crystal structure of the corresponding iron(II) complex.53 b Torsion angles obtained from the crystal structures of [FeIV(O)(N4Py)]2+ and [FeIV(O)(Py5Me2)]2+.51,52 c Torsion angles in parentheses from the crystal structure of [FeIV(O)(5Me2N4Py)]2+.52 | |||||
1(IV)NTs | 41.0 (34) | 12.7 (5.7) | −13.3 (−20.3) | This work | |
−5.6 (−12.6) | 6.0 (−1.0) | −6.9 (−13.9) | |||
2(IV)NTs | 20.9 (13.9) | 9.7 (2.7) | −6.6 (−13.6) | This work | |
[FeIV(O)-(BnTPEN)]2+a | 6° | 43 (36) | 10.6 (3.6) | −15 (−22) | 51 |
16° | 40 (33) | 10.0 (3.0) | −15.5 (−22.5) | ||
91° | −0.3 (−7.3) | 8.2 (1.2) | −1.2 (−8.2) | ||
[FeIV(O)-(N4Py)]2+b | 13.9° (6.8°) c | 44 (37) | 9.5 (2.5) | −17 (−24) | 51 |
21.1° (30.4°) c | 30 (23) | 8.3 (1.3) | −11 (−18) | ||
[FeIV(O)-(Py5Me2)]2+b | Ave 36.5° | 27 (20) | 3.5 (−3.5) | −12 (−19) | 52 |
The largest shifts are found for the pyridines parallel to the FeO unit, while the smallest shifts are found for the pyridines perpendicular to the FeO unit (Table 1). A perusal of this table indicates that this pattern can also be used to interpret the NMR data for 1(IV)NTs and 2(IV)NTs. There are two sets of pyridine peaks in the spectrum of 1(IV)NTs, because one pyridine is nearly parallel to the FeN bond while the other is almost perpendicular to the FeN bond. On the other hand, 2(IV)NTs only exhibits one set of pyridine-derived signals because of the plane of symmetry relating the two pyridines and the paramagnetic shifts observed for the pyridine protons have intermediate values that are between the two extremes found in 1(IV)NTs because of the intermediate angle found for 2(IV)NTs.
Furthermore, complex 1(IV)NTs exhibits a broad resonance enhanced Raman feature at ca. 984 cm−1 (Fig. 3) obtained with 488.0 nm laser excitation, which can be assigned to a vibration arising from the FeN moiety on the basis of the 24 cm−1 downshift observed upon formation of 1(IV)NTs with PhI15NTs. Similarly, the Raman spectrum of compound 2(IV)NTs shows a broad feature around 1016 cm−1 that shifts ca. 22 cm−1 upon formation of 2(IV)NTs with PhI15NTs. The observed downshifts approximate the 27 cm−1 shift calculated for a diatomic FeN oscillator, but the frequencies observed are 200 cm−1 higher than found for corresponding FeIVO units, which have shorter FeO distances than the FeN distances found for FeNTs complexes.2
Fig. 3 Resonance enhanced Raman spectra for complex 1(IV)NTs (top) and 2(IV)NTs (bottom) upon excitation at 488 nm. |
A rationale for the higher frequencies observed for 1(IV)NTs and 2(IV)NTs complexes reported here can be found from a comparison of the vibrational spectra of other FeNR complexes. For example, analogous six-coordinate FeIV(NTs) complexes supported by other pentadentate N5 ligands, 3(IV)NTs and 4(IV)NTs, where N5 = N4Py and BnTPEN, respectively, also exhibit vibrations at 998 and 984 cm−1 that respectively downshift to 975 and 957 cm−1 upon 15N labelling.54 Related tetrahedral FeNR complexes give rise to 15N-sensitive vibrations with frequencies in the 1100 cm−1 range.55 The higher frequencies observed for these complexes are rationalized by the coupling of the FeN stretching mode to the higher-frequency N–S or N–C stretching mode of the NTs or NR moiety to account for the upshift in vibrational frequency. This coupling interaction increases as the FeN and N–(S or C) bonds approach linearity. For example, DFT calculations on 6-coordinate [FeIV(NTs)(N4Py)]2+ predict an FeN–S angle of 161°,54 while an FeN–C angle of 179° is found crystallographically for the tetrahedral [(PhBP3)FeIII(NtBu)] (PhBP3 = [PhB(CH2PPh2)3]−) complex.
One exception to the vibrational trends observed in the FeNTs complexes discussed above is the recently reported [FeV(NTs)(TAML)]− complex (TAML = TetraAmido Macrocyclic Ligand), which exhibits a ν(FeN) mode at 817 cm−1,31 which falls in the range of values reported for FeO vibrations.2 Indeed it is 45 cm−1 lower in frequency than the ν(FeO) mode reported for the [FeV(O)(bTAML)]− (bTAML = biuret TetraAmido Macrocyclic Ligand) complex,56 commensurate with the 0.07 Å longer FeNTs distance found for [FeV(NTs)(TAML)]− by EXAFS analysis.31 DFT calculations on this complex find a more acute FeN–S angle at 125°, thereby essentially uncoupling the two vibrations and allowing the actual FeN stretching frequency to be observed.
X-ray diffraction quality crystals could not be obtained for 1(IV)NTs and 2(IV)NTs, but insight into their structures could be derived from EXAFS analysis of frozen samples in acetone (Fig. 4) and supported by comparison to the corresponding amido derivatives 1(III)–NHTs and 2(III)–NHTs which will be described in the next section.
Fig. 4 Iron K-edge XANES region (left) for 1(III)–NHTs, 2(III)–NHTs, 1(IV)NTs and 2(IV)NTs together with (right) non-phase shift corrected Fourier transforms of EXAFS data for 1(IV)NTs and 2(IV)NTs with insets showing the k-space data and fits. Experimental data are shown in black and fits in red (see fit Tables S4–S7†). |
Iron K-edge X-ray absorption spectroscopy (XAS) of 1(IV)NTs and 2(IV)NTs (Fig. 4) show rising edges centred at 7123.4 eV and 7123.3 eV, having pre-edge features corresponding to 1s → 3d transitions at 7113.2 eV and 7113.3 eV similar to those previously reported iron(IV) centered complexes (Table 2, ESI Table S8†).30,49,57,58 Inspection of the iron K-edge pre-edge and rising energies of the 1(III)–NHTs (1s → 3d 7112.5 eV, 7122.5 eV) and 2(III)–NHTs (1s → 3d 7112.4 eV, 7122.3 eV) derivatives as well as the recently reported analogous 1(V)–N complex ([FeV(N)(MePy2tacn)]+2, (1s → 3d 7114.2 eV, 7123.8 eV))59 show a systematic shift of the XANES profile to higher energy strongly suggesting metal centred oxidation for the series.60–63 In particular, the pre-edge energy shifts in steps of about 0.9 eV across the series, consistent with a +1 change in formal oxidation state at the metal centre (Table 2 and S8†). Additionally, the pre-edge intensities increase from about 0.11 normalized units for X(III)–NHTs to 0.18 for X(IV)–NTs and 0.48 for X(V)–N due to both a change in oxidation as well as enhanced p–d mixing due to more covalent, shorter Fe–N/(NTs) bonds, causing ever increasing tetragonal distortions of the pseudo-octahedral environment.58 This is supported by EXAFS analysis (Tables S4–S7†) showing six-coordinate environments for both X(III)–NHTs and X(IV)–NTs species consisting of 5 N/O scattering atoms at ∼2.0 Å and a shorter Fe–N bond, presumably from the tosylimido moiety. For X(IV)–NTs complexes this bond, centred at 1.72 Å, is ∼0.12–0.17 Å shorter than that determined for X(III)–NTs species (∼1.84 Å–1.89 Å) and similar to that of the previously reported 3(IV)NTs (1.73 Å).30
Mössbauer spectroscopy | Raman v14N/15N (cm−1) | X-ray absorption | |||||
---|---|---|---|---|---|---|---|
δ (mm s−1) | ΔEQ (mm s−1) | 1s → 3d area (×102) | 1s → 3d (eV) | (eV) | Fe–N (Å) | ||
a DFT calculated values in parenthesis. b Data from ref. 30 and 54. c Data from complex prepared in acetonitrile. | |||||||
1(IV)NTs | 0.05 (0.03) | 1.09 (1.01) | 984/960 (1029) | 18.5 | 7113.2 | 7123.4 | 1.71 (1.73) |
2(IV)NTs | 0.06 (0.04) | 0.73 (0.77) | 1016/994c (1063) | 18.0 | 7113.3 | 7123.3 | 1.72 (1.72) |
3(IV)NTs | 0.02 | 0.98 | 998/975 | 18.0 | 7113.1 | 7123.0 | 1.73 |
To further our understanding of X(IV)–NTs complexes theoretical models of both 1(IV)NTs and 2(IV)NTs were explored with density functional theory. Geometry optimized structures at the B3LYP/6-311+g** level together with electronic energies at the B3LYP/cc-pVTZ level favor a triplet structure for X(IV)–NTs complexes with predicted Fe–NTs bond distances of 1.73 Å and 1.72 Å respectively with a second shell of 5 N scattering atoms at ∼2.03 Å, comparable to experimentally determined EXAFS metrics (Table 2). In the quintet X(IV)–NTs models on the other hand, the Fe–NTs distance is still centred around 1.72 Å, but the remaining ligands are split into two scattering shells of 3 N at 2.10 Å and 2 N at 2.20 Å, well outside the EXAFS experimental fit parameters. In addition, Mössbauer parameters calculated for the triplet structures are in good agreement with experimentally derived values (Table 2), suggesting that in addition to geometry the electronic structure is also well described by the theoretical models. It is important to highlight that similar to the iron(III) analogues (vide infra) the quadrupole splitting for 1(IV)NTs is larger than that of 2(IV)NTs, suggesting a more centrosymmetric environment for 2(IV)NTs. Finally, a description of the electronic structure in terms of a molecular orbital picture was carried out at the B3LYP/def2-TZVP level to facilitate comparison with the previously described electronic structure of 3(IV)NTs data30 (Fig. 5). Similar to 3(IV)NTs we find a highly covalent Fe–NTs interaction in the X(IV)–NTs species with spin density shared by both the iron center (∼1.11) and the N of the tosylimido ligand (∼0.7). As such the question of a iron(IV)–N imido or iron(III)–N− imido radical pair may arise, the latter situation being recently described by Neidig et al. for a pyrazolyl derived high valent iron–imido complex.38 It is important to note however, that the Mössbauer isomer shift for the pyrazolyl complex is 0.20 mm s−1 with a quadrupole splitting of 1.96 mm s−1, parameters that may indeed be attributed to an iron(III) center, and which indeed is quite similar to that of the corresponding iron(III)–amido derivative (0.26 mm s−1) and to values for X(III)–NHTs complexes reported herein. Secondly, the experimentally derived XAS rising edges for our family of complexes consistently show metal based oxidation which when correlated with their corresponding Mössbauer isomer shifts and rising edges of similar Fe–imido complexes fall in a range best attributable to formally iron(IV) centers (Fig. S15†).30,59,64 Lastly, it is important to note that spin density alone may not be a good measure for inferring oxidation state for FeE (E = O, N) type centers. Experimental and theoretical descriptions of FeO centers, that are isoelectronic to FeN, carried out by Que and Solomon show that the highly covalent nature of the Fe–O bond increases the unpaired spin-density on the O-atom.65 Systematic theoretical descriptions by Neese et al. further show that the unpaired spin-density is expected to increase with metal oxidation state and is further favoured in the π* manifold in the case of N due to the smaller effective nuclear charge of N over O resulting in a more covalent π-interaction for iron–imido complexes.66 FeO bonds on the other hand are predicted to have a stronger σ-interaction which helps rationalize the shorter FeE bonds in iron–oxo moieties (1.63 Å).30,49,66 In the FeNTs case the weaker σ-interaction results in nearly degenerate Fe dx2−y2 and Fe dz2 energies as opposed to FeO where the Fe dx2−y2 and Fe dz2 are well separated.65 Therefore the X(IV)–NTs complexes are best described as having a formally +4 oxidation spin 1 iron center similar to the previously reported 3(IV)NTs complex.30 Employing both unrestricted corresponding orbitals (Fig. S16†), which highlight doubly, singly or spin paired molecular orbitals based on the spatial overlap of spin up and down pairs, as well as quasi-restricted orbitals, which allow for the “familiar” picture of doubly and singly occupied as well as empty orbitals,67 a highly mixed and covalent Fe d-orbital manifold is evidenced. Yet, the primary contribution to the SOMO orbitals is the iron metal center yielding a π(Op)4(Fedxy)2π*(Fedxz,yz)2σ*(Fedx2−y2)0σ*(Fedz2)0 arrangement giving an effective bond order of 2 for the FeN moiety (Fig. 5).
To date only a few iron(IV)–imido species have been prepared and structurally characterised. X-ray diffraction examples include three41 and four coordinate iron complexes.6,23,43,45,68,69 Most of these contain a distorted tetrahedral geometry,6,23,45,68,69 but a complex with a cis-divacant octahedron geometry has also been recently described.43 Notably, all of them exhibit significantly shorter FeN bonds of 1.61–1.64 Å in comparison to octahedral systems. The aforementioned 3(IV)NTs (Fe=N 1.73 Å from EXAFS data) is the only example structurally and electronically comparable to 1(IV)NTs and 2(IV)NTs. On the other hand, examples of ferric–imido radical complexes have been described,38–40 including an example of an octahedral complex.38 FeN distances determined for 1(IV)NTs and 2(IV)NTs fall between those of high-spin three and four coordinate ferric–imido radical complexes (1.76–1.77 Å)18,40 and the computed distance for the single example of the low-spin octahedral complex (1.696 Å) recently described by Neidig and Maron.38 Therefore, the distinctive Mössbauer parameters but not the Fe–N distance allow the differentiation between the two electronic structures.
S | g x | δ (mm s−1) | ΔEQ (mm s−1) | Γ (mm s−1) | η | A x /gnμn | % | E o (Epre-edge) (eV) | |
---|---|---|---|---|---|---|---|---|---|
g y | A y /gnμn | ||||||||
g z | A z /gnμn (T) | ||||||||
[1(III)–NHTs](OTf)2 | 1/2 | 2.23 | 0.28 | 2.40 | 0.57 | 0 | −36.2 | 100 | 7122.5 (7112.5) |
2.23 | 12.8 | ||||||||
1.93 | 9.1 | ||||||||
[2(III)–NHTs](OTf)2 | 1/2 | 2.24 | 0.27 | −2.03 | 0.25 | 0.84 | −23 | 100 | 7122.3 (7112.4) |
2.24 | 4.6 | ||||||||
1.95 | 3.8 |
Single crystals suitable for X-ray diffraction were obtained for [1(III)–NHTs](SbF6)2. An ORTEP diagram corresponding to its molecular structure along with selected bond distances are shown in Fig. 6. The complex contains an iron centre in a distorted octahedral geometry. Five coordination sites are occupied by nitrogen atoms from the pentadentate ligand with Fe–N distances typical for iron(III) in low-spin (1.98–2.03 Å).70–72 The sixth site is occupied by the tosylamido N atom. The two pyridine rings are perpendicular one to each other, maintaining the structure of 1(II), with one of them parallel to the Fe–N(H) axis. Fe–Py bonds are 1.990 and 1.988 Å and Fe–Nalkyl are 2.024, 2.027 and 1.986 Å. Fe–Py bonds are slightly shorter likely due to the π-acceptor character of pyridine. The Fe–N(H) distance is relatively short (1.89 Å). For comparison the Fe–N(H) bond is 0.18 Å longer than FeN in 1(IV)NTs.
The Fe–N(H) distance in 1(III)–NHTs is very similar to the values reported by Spasyuk et al. for octahedral complexes [FeIII(NHR)(B2Pz4Py)] (R = C6H4C(CH3)3 or adamantane, 1.869 and 1.854 Å, respectively).38 This is in line with the low-spin character of the iron center in all three complexes. By contrast, longer Fe–N(H) distances in the range 1.95–2.00 Å are noted for the high-spin five coordinate ferric amido complexes [FeIII(NHTol)H22] and [FeIII(NH–SO2C6H4R)TPA2C(O)NHtBu](OTf)2 reported respectively by Borovik73 and Chang.74
EPR studies of 1(III)–NHTs and 2(III)–NHTs at 2 K (Fig. 7) show axial spectra (geff,⊥ = 2.23 and geff,// = 1.93) and (geff,⊥ = 2.21 and geff,// = 1.95) indicative of S = 1/2 systems, characteristic of low-spin ferric centres. In both samples a minor high-spin signal is observed. Mössbauer parameters obtained are congruent with the presence of low-spin ferric centers for both complexes. At 80 K in the absence of a magnetic field, both show asymmetric doublets with δ ≈ 0.28 mm s−1, but differ only in ΔEQ = 2.40 and 2.03 mm s−1, respectively (Fig. 7). When recorded at 4.2 K under a field of 7 T applied parallel to the γ-rays, the spectra of both complexes split into a multiplet spread over a limited velocity range −3 to +3.5 mm s−1 which is consistent with an S = 1/2 center. The data could indeed be fitted under this assumption with the parameters listed in Table 3.
Fig. 7 Spectroscopic data of isolated 1(III)–NHTs (left) and 2(III)–NHTs (right). (Top) EPR spectra of 1(III)–NHTs (left) and 2(III)–NHTs (right). (Middle) Mössbauer spectra of 1(III)–NHTs (left) and 2(III)–NHTs (right) recorded at 80 K in absence of field (middle) and at 4.2 K under a field of 7 T applied parallel to the γ-ray (bottom). Experimental spectrum: hatched bars, theoretical spectrum: solid line (see Table 3 for parameters). |
X | XPhS(NTs)Me yields (%) | |||
---|---|---|---|---|
Complex | 1(IV)NTs | 2(IV)NTs | ||
Solvent | Acetonitrile | Acetone | Acetonitrile | |
MeO | 57 | 42 | 48 | |
Me | 46 | 43 | 44 | |
H | 49 | 53 | 33 | |
Cl | 44 | 43 | 29 |
Frozen samples were prepared immediately following the first rapid phase of the reaction of 1(IV)NTs and 2(IV)NTs with MeOPhSMe, and analysed by Mössbauer and EPR spectroscopy. These analyses (see ESI† for details) revealed the presence of monoiron species with an approximate relative ratio of Fe(II):Fe(III) of 1:6 for the decay of 1(IV)NTs and 1:3 for 2(IV)NTs. The major ferric species formed after reaction of 1(IV)NTs and 2(IV)NTs with the sulphide can be identified as 1(III)–NHTs and 2(III)–NHTs/2(III)–OH, respectively on the basis of the agreement with the EPR and Mössbauer parameters of the independently prepared complexes. On the other hand, the parameters of the minor ferrous component observed in the Mössbauer spectra (Fig. S7 and Table S1†) are characteristic of a low-spin ferrous centre and are indicative of the formation of 1(II) and 2(II), respectively.49 Of interest, no reaction is observed when 1(IV)NTs is mixed with 1(II) or when 2(IV)NTs is mixed with 2(II). Therefore, it can be concluded that 1(III)–NHTs and 2(III)–NHTs do not originate from subsequent comproportionation reactions. The sum of the spectroscopic data and the stoichiometry of the reactions of 1(IV)NTs and 2(IV)NTs with sulphides indicates that transfer of the tosylimido moiety to the sulphide does not correspond to a 2e− N-transfer where the iron(IV) centre becomes reduced to iron(II), as it has been previously documented for 3(IV)NTs,75 but instead, a competitive and most dominant path involves formation of iron(III) species.
d[FeIV]/dt = −kobs[FeIV] | (1) |
The respective pseudo-first order rate constants (kobs) were then calculated from this fitting.
In the first place, reactions of 1(IV)NTs and 2(IV)NTs at a fixed concentration (0.25 mM) against different concentrations of thioanisole substrate, used in large excess (30–220 mM) were analysed (Fig. 9, panels A and B). Observed reaction rates (kobs) were found to be linearly dependent on sulphide concentration, indicating that they are pseudo-first order rate constants with the expression kobs = k′[sulphide].
The k′ values for 1(IV)NTs and 2(IV)NTs were determined for para-substituted thioanisole substrates (XPhSMe, X = MeO, Me, H, Cl) and are shown in Table 5. A Hammett plot was represented by plotting against Hammett parameters (σp). The obtained plots (Fig. 9, panels C and D) show linearity with a slope of ρ = −3.12 and ρ = −3.89 for 1(IV)NTs (kobs values were determined in reactions with a common concentration of 1(IV)NTs and 2(IV)NTs respectively). These values are relatively large in magnitude compared to those described for OAT reactivity examples with iron(IV)–oxo compounds, which range between ρ = −1 to −2,76–79 but are in good agreement with that associated with N atom transfer with 3(IV)NTs (−3.35).47 Moreover, when k′ values were plotted against the oxidation potentials of corresponding thioanisole substrates, a linear plot was obtained with slopes of −5.9 and −7.8 V−1 for 1(IV)NTs and 2(IV)NTs respectively (Fig. S8†). For comparison, for OAT with iron(IV)–oxo processes those values range between −2 and −3 V−1. The relatively large slopes exhibited by 1(IV)NTs and 2(IV)NTs in comparison with 1(IV)O and 2(IV)O indicate that the former reactions proceed with larger charge transfer in the transition states. In fact, the herein determined values are reminiscent of reactions where a single electron transfer process is the rate determining step.47,79 However, as will be discussed in the following lines, thermodynamic considerations discard this scenario.
Comparison between reaction rates determined for 1(IV)NTs and 2(IV)NTs with respect to those previously described for 3(IV)NTs deserve some comment (Table 5). Of note, even when performed at a higher temperature (293 K vs. 273 K), reactions of 1(IV)NTs and 2(IV)NTs are nearly two orders of magnitude slower than those of 3(IV)NTs. These observations suggest that 1(IV)NTs and 2(IV)NTs are milder oxidants. Comparison with reaction rates for oxo-transfer reactions to sulphides by the corresponding oxo complexes 1(IV)O and 2(IV)O is also pertinent. Comparison between 3(IV)NTs and 3(IV)O revealed the former to react with sulphide five times faster than the latter.75 In contrast, oxo transfer reactions by 1(IV)O and 2(IV)O are systematically faster than tosylimido transfer by 1(IV)NTs and 2(IV)NTs, with reaction rates between two to twenty times larger when comparing complexes with the same pentadentate ligand.
Scheme 3 Mechanism proposal for the first phase of the reactions of iron(IV)–tosylimido complexes with thioanisole substrates. Solv. stands for solvent molecule. |
Alternatively, reaction of 1(IV)NTs and 2(IV)NTs with p-X-ArSMe substrates is proposed to take place via a two-step process and involves two molecules of the iron(IV)–tosylimido complex (1(IV)NTs or 2(IV)NTs) for each molecule of sulphide. In the first, rate determining step, the sulphide reacts via a nucleophilic attack on the tosylimido N atom of a first molecule of iron(IV)–tosylimido complex. This attack results in tosylimido transfer to the substrate, forming a putative [FeII(p-X-ArS(NTs)Me)(L)], 1–2(II)ArSNTs intermediate that may undergo sulfilimine dissociation and acetonitrile binding to yield [FeII(CH3CN)(L)]2+, as earlier documented for 3(IV)NTs.47 However, in the case of 1(IV)NTs and 2(IV)NTs, the latter appears to be a minor path (k3 < k2), and instead 1–2(II)ArSNTs can be oxidized by a second molecule of 1(IV)NTs or 2(IV)NTs, to form one equivalent of sulfilimine, and two equivalents of ferric species. One of the two retains a tosylimido ligand and converts to 1(III)–NHTs or 2(III)–NHTs presumably by proton transfer from adventitious water. Formation of additional amounts of 1(III)–NHTs or 2(III)–NHTs/2(III)–OH from the five coordinate ferric species [FeIII(L)] may originate from reaction with residual H2NTs and adventitious water.
Therefore, while the three iron(IV)–imido complexes 1–3(IV)NTs react with sulphides via an initial two-electron tosylimido transfer step, the reactions that follow this step are dependent on the nature of L. Structural/steric factors depending on the specific complex may account for the different reactivity, by limiting bimolecular reactions or favouring dissociation. Specifically, the N4Py ligand may provide a sterically more demanding scaffold than the tacn based ligands (MePy2tacn and Me2(CHPy2)tacn). This factor presumably favors rapid sulfilimine dissociation from 3(II)ArSNTs and disfavors its reaction with 3(IV)NTs.
Reactions were studied kinetically. The decay of 1(IV)NTs and 2(IV)NTs exhibits an exponential behavior upon addition of different substrate concentrations. Unexpectedly, the observed rates (kobs, s−1), extracted from the exponential fits, do not vary linearly with the substrate concentrations and instead saturation profiles are observed (Fig. 10).
Conditions that lead to rate saturation profiles involve the specific mechanistic feature of presenting a fast reversible reaction step that precedes a slower irreversible step. Thus, by fitting the data with the pre-equilibrium approximation model:
kobs = (Keq × k × [CH])/(1 + Keq × [CH]) | (2) |
The values of k (s−1) and Keq (M−1) reported in Table 6 could be derived.
BDE, kcal mol−1 | Substrate | 1(IV)NTs | 2(IV)NTs | ||
---|---|---|---|---|---|
k, s−1 | K eq, M−1 | k, s−1 | K eq, M−1 | ||
75.5 | Xanthene | 3.5 ± 0.7 (10−2) | 254 ± 118 | 1.3 ± 0.9 (10−3) | 112 ± 19 |
77 | DHA | 2.2 ± 0.2 (10−2) | 315 ± 81 | 6.0 ± 0.6 (10−3) | 227 ± 67 |
DHA-D4 | 5.2 ± 0.6 (10−3) | 390 ± 115 | 8.3 ± 0.4 (10−4) | 314 ± 74 | |
78 | CHD | 1.8 ± 0.2 (10−2) | 301 ± 154 | 4.4 ± 0.2 (10−3) | 203 ± 31 |
80 | Fluorene | 3.4 ± 0.3 (10−3) | 329 ± 118 | 7.0 ± 0.9 (10−4) | 415 ± 301 |
The logarithms of the extrapolated rates k (s−1) values corrected by the number of CH (nH) contained in each substrate (log(k/nH) = log(k′)) exhibit a linear dependence versus the BDE of the studied substrates (Fig. S11†), with slopes of −0.21 and −0.26 kcal−1 mol for 1(IV)NTs and 2(IV)NTs complexes, respectively. Notably, these slopes are in good agreement with those determined for HAT reactions of iron(IV)–oxo complexes with the same substrates (between −0.1 and −0.4 kcal−1 mol),48,77,82,83 reinforcing the credibility of the kinetic analysis. This suggests the contribution of a H-atom abstraction event in the rate-determining step of substrates oxidation reaction.84 Consistently, reactions with deuterated DHA (DHA-D4) give rise to k′ values that reveal a primary KIE of 4 and 7 for 1(IV)NTs and 2(IV)NTs, respectively.
The values observed for Keq are all in the range, within experimental error, of 150–500 M−1 for both 1(IV)NTs and 2(IV)NTs, and they are barely related to the BDE of the substrate involved in the reaction. Furthermore, reaction with deuterated DHA (DHA-D4) produce k′ values that reveal a primary KIE of 4 and 7 for 1(IV)NTs and 2(IV)NTs, respectively, while equilibrium constants (Keq) remain basically constant (Table 6 and Fig. 10). These results strongly suggest that a hydrogen atom abstraction is not involved in the first reversible reaction step, but instead contribute to the second, irreversible, rate-determining step of the oxidation.
The reaction of 1(IV)NTs with DHA has also been studied in acetone as solvent. Again, rate constants show saturation profiles with increasing substrate concentration. The overall process is accelerated in acetone, with a kacetone/kCH3CN ratio of 5 at 20 °C, while equilibrium constants Keq are very similar, with a Keq acetone/Keq CH3CN ratio of 0.8 (Table S3 and Fig. S12†). Unfortunately, direct comparison with the reaction rates exhibited by 3(IV)NTs and oxo–iron complexes 1(IV)O and 2(IV)O, is not possible because of the different rate laws.
Scheme 4 Mechanism proposal for the reactions of iron(IV)–tosylimido complexes 1(IV)NTs and 2(IV)NTs with hydrocarbons. |
The exact nature of this pre-equilibrium step cannot be determined, because the intermediate species that form do not show distinctive spectroscopic features. However, kinetic evidence for the formation of association complexes preceding hydrogen atom transfer by metal oxo species85–87 and free radicals88,89 have precedents in the literature. With few exceptions,85 these reactions involve substrates with polarized hydrogen atoms that can engage in hydrogen donor interactions. In the current reactions, reversible hydrogen bond interaction of the substrates with the FeIVNTs can be reasonably discarded because of the apolar nature of the hydrocarbons, and because acetonitrile and acetone are polar solvents that should limit weak hydrogen-bond interactions. Instead, we suggest that the substrate may engage in a series of weak interactions with the electron poor aromatic rings of the tosylimido and ligand pyridine moieties.
Notably, the saturation profiles observed for the reaction of 1(IV)NTs and 2(IV)NTs with hydrocarbons are in contrast with the kinetic behavior of the analogous reactions performed by the related Fe(IV)O species,48,81 by 3(IV)NTs (ref. 75) and by ferric–imido/imidyl radicals,11,38,90,91 for which rate constants vary linearly with substrates concentration. With the single exception of 3(IV)NTs, these reactions are kinetically described as single step HAT processes. In the case of 3(IV)NTs, reversible ET is proposed to precede HAT,47 formally resulting in a hydride transfer reaction. A thermochemical analysis indicates that an analogous ET can not be proposed to account for the pre-equilibrium observed in the reactions of 1(IV)NTs and 2(IV)NTs; the DFT computed single electron oxidation of cyclohexadiene to the cyclohexadienyl radical is 1.61 V vs. SHE, rendering outer sphere ET oxidation by 1(IV)NTs and 2(IV)NTs highly endergonic.
Important differences arise also when comparing KIE's; the values determined for 1(IV)NTs and 2(IV)NTs are much smaller than the remarkably large value (kH/kD = 567) obtained for the octahedral ferric imidyl radical complex described by Spasyuk et al.38 but instead is closer to that of 3(IV)NTs,47 (kH/kD = 7) and to Betley's low-coordinate ferric-amidyl radical complex (kH/kD = 13).11
Rationalization of the HAT reactivity of 1(IV)NTs and 2(IV)NTs may be done by estimating the BDE of the N–H bond in the resulting 1(III)–NHTs and 2(III)–NHTs. As earlier indicated, the pKa and E1/2(FeIV/FeIII) values necessary for determining these BDE's can not be obtained experimentally, and were then derived from DFT methods (B3LYP, see ESI† for details). From this analysis, the deduced BDE's for N–H bonds in 1(III)–NHTs and 2(III)–NHTs are 91.6 and 94.0 kcal mol−1, which rationalizes the HAT ability of 1(IV)NTs and 2(IV)NTs. A cautionary note is that the analysis predicts that 2(IV)NTs should be much more reactive than 1(IV)NTs, but this is not observed. This suggests that steric factors may also contribute in a significant extent to modulate the relative reactivities of the two complexes.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1918211. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c9sc02526j |
This journal is © The Royal Society of Chemistry 2019 |