Dimitra K.
Gioftsidou
a,
Michael G.
Kallitsakis
b,
Konstantina
Kavaratzi
a,
Antonios G.
Hatzidimitriou
a,
Michael A.
Terzidis
c,
Ioannis N.
Lykakis
*b and
Panagiotis A.
Angaridis
*a
aLaboratory of Inorganic Chemistry, Department of Chemistry, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece. E-mail: panosangaridis@chem.auth.gr
bLaboratory of Organic Chemistry, Department of Chemistry, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece. E-mail: lykakis@chem.auth.gr
cLaboratory of Chemical Biology, Department of Nutritional Sciences and Dietetics, International Hellenic University, Sindos, 57400 Thessaloniki, Greece
First published on 21st December 2023
Reduction of nitro-compounds to amines is one of the most often employed and challenging catalytic processes in the fine and bulk chemical industry. Herein, we present two series of mononuclear homoleptic and heteroleptic Co(III) complexes, i.e., [Co(LNS)3] and [Co(LNS)2L1L2]x+, respectively (x = 0 or 1, LNS = pyrimidine- or pyridine-thioamidato, L1/L2 = thioamidato, phosphine or pyridine), which successfully catalyze the transformation of nitroarenes to anilines by methylhydrazine. The catalytic reaction can be accomplished for a range of electronically and sterically diverse nitroarenes, using mild experimental conditions and low catalyst loadings, resulting in the corresponding anilines in high yields, with high chemoselectivity, and no side-products. Electronic and steric properties of the ligands play pivotal role in the catalytic efficacy of the respective complexes. In particular, complexes bearing ligands of high hemilability/lability and being capable of stabilizing lower metal oxidation-states exhibit the highest catalytic activity. Mechanistic investigations suggest the participation of the Co(III) complexes in two parallel reaction pathways: (a) coordination-induced activation of methylhydrazine and (b) reduction of nitroarenes to anilines by methylhydrazine, through the formation of Co(I) and Co-hydride intermediates.
For the above reasons, the quest for novel homogeneous catalytic systems which rely on the low-cost and low-environmental impact, earth-abundant 3d-metals that can mediate the reduction of nitro-compounds to amines selectively and under mild conditions is gaining growing interest.10 In this context, a method successfully and widely used for this transformation is Béchamp's reduction, which involves the use of Fe powder in concentrated HCl(aq). However, the utilization of harsh reaction conditions and the generation of toxic waste, in combination with the lack of functional group selectivity, suggests the imperative need for more economical and efficient methods.
To that end, efficient molecular catalysts based on 3d-metal ions have recently been developed. In general, structures and physicochemical properties of transition metal molecular catalysts, and, therefore, their catalytic activity, can be tuned via selection of ligands with appropriate stereochemical and electronic properties or the metal itself,11 so that they can catalyze efficiently a particular catalytic transformation, such as the selective reduction of nitroarenes (ArNO2) to anilines (ArNH2). A number of transition metal complexes that catalyze this transformation via catalytic hydrogenation has been reported. However, they mainly rely on noble metals,12,13 and only a few examples of 3d-metal molecular catalysts, such as Fe,14–16 Mn,17 and Co18 complexes, have been reported. Catalytic transfer hydrogenation reactions, utilizing hydrazine derivatives, acids, or solvents as hydrogen donor molecules,19,20 in combination with Fe,21–23 Mn,24 and Co25 molecular catalysts have also been developed. In addition, NaBH4 has also been used for the reduction of nitro-compounds in the presence of Fe-based catalysts.26,27 Finally, another method that is also attracting increasing attention is catalytic hydrosilylation, due to the mild reaction conditions employed and the selectivity exhibited occasionally towards other functional groups. The most commonly used 3d-metal molecular catalysts in these reactions involve Fe,28–32 Co,33 Ni,34,35 and more recently, Mn complexes.36
Among the least studied molecular catalysts of earth-abundant metals for the reduction of nitroarenes are those of Co. Nevertheless, owing to the ability of this metal to easily undergo three consecutive one-electron redox transformations Co(I)/Co(II)/Co(III) and to adopt both high- and low-spin electronic states, its complexes are expected to function as competent catalysts for this multi-electron redox reaction. In one of the earliest reports, Co(II) phthalocyanine (Scheme 1) was found to successfully catalyze the reduction of nitroarenes by NH2NH2·H2O.37 More recently, Gupta and coworkers described the reduction of a range of different nitro-compounds by mononuclear Co(II) or Co(III) complexes bearing a tetradentate N-donor ligand, in the presence of NH2NH2.38,39
![]() | ||
Scheme 1 Examples of molecular cobalt catalysts for the reduction of nitroarenes to anilines reported in the literature. |
Recently, our group reported the transfer hydrogenation of nitroarenes to anilines by CH3NHNH2 using a tris-chelating pyrimidine-thioamidato Co(III) complex as catalyst.40 Aiming to develop more efficient catalysts for this transformation, we extended our research endeavours in the field. In particular: (a) we synthesized new mononuclear thioamidato Co(III) complexes bearing different heteroaromatic thioamidato ligands with electron-withdrawing (EW) or electron-donating (ED) group substituents, (b) we studied the catalytic potential of these Co(III) complexes in the reduction of a range of different nitroarenes, and (c) we investigated their catalytic reaction mechanism. Generally, heteroaromatic thioamide (LNHS) ligands offer a range of attractive features. Specifically, they have the capacity to: (a) affect the redox properties of their complexes (through their stereochemical and electronic properties), (b) exhibit hemilabile/labile coordination behaviour (acting either as terminal monodentate κS- or κN- or bidentate chelating κS,N-ligands), and (c) participate in proton transfer reactions (owing to the presence of N atoms in their heterocyclic ring that can act as protonation sites).41–43 Therefore, the combination of these ligands with a redox-active metal center, such as Co, offer an ideal scaffold for the synthesis of robust, versatile and efficient molecular transition metal catalysts for the reduction of nitroarenes.44,45 All complexes synthesized herein were found to be efficient catalysts for the reduction of a variety of nitroarenes to the corresponding anilines in high yields and chemoselectivity, without formation of azo and azoxy intermediates, under mild experimental conditions and low catalyst loading. Interesting structure–catalytic activity relationships were established. Finally, mechanistic investigations, using a range of different physicochemical methods, allowed us to suggest plausible reaction pathways for the catalytic transformation.
The first series of complexes comprises homoleptic tris-pyrimidine-thioamidato Co(III) complexes: [Co(pymt)3] (1) bearing unsubstituted thioamidato ligands, [Co(tfmp2S)3] (2) having CF3-group substituted thioamidato ligands, and [Co(dmp2S)3] (3) and [Co(damp2S)3] (4) with CH3- or NH2-group substituted thioamidato ligands, respectively. As shown in Scheme 2, these were synthesized from reactions of CoCl2·6H2O with the corresponding pyrimidine–thioamide in 1:
3 molar ratio, in the presence of (CH3CH2)2NH, in CH3OH, at 50–70 °C and under atmospheric conditions (Section S1.2 in the ESI†).
Views of the molecular structures of 1–4, determined by single-crystal X-ray diffraction analysis (Table S1†), are presented in Fig. 1.40 In all cases, an octahedrally coordinated Co(III) ion is surrounded by three chelating pyrimidine-thioamidato ligands bound through their exocyclic S and heterocyclic N donor atoms (κS,N-LNS). Their Co–N and Co–S bond lengths fall within the ranges of 1.90–1.99 and 2.25–2.34 Å, respectively (Table S2†), indicative of low-spin, octahedrally coordinated d6 Co(III) ions. Noticeably, 3 and 4 exhibit slightly shorter Co–S bond lengths, compared to 1, a fact that can be attributed to the ED group substituents (CH3 and NH2) of their ligands, which increase the basicity of their exocyclic S atom and strengthen its interaction with the metal center. In all cases, distortions from the ideal octahedral coordination geometry of the metal centers are observed (Table S2†). These are due to the small chelate ring angles of the thioamidato ligands, which result in very narrow bite angles around the metal centers. However, the sensitivity of bond angles to crystal packing forces should also not be neglected. Interestingly, the relative dispositions of the thioamidato ligands of 1, 2 and 4 display a meridional arrangement of their respective S and N donor atoms, while in 3 a facial arrangement is observed. This variation is in accordance with examples of analogous thioamidato Co(III) complexes reported in the literature.46–48
![]() | ||
Fig. 1 Molecular structures of 1–4 as determined by single crystal X-ray diffraction. Thermal ellipsoids are drawn at the 35% probability level. All H atoms are omitted for clarity. |
FTIR spectra of 1–4 in the 4000–400 cm−1 region are dominated by characteristic bands of their respective thioamidato ligands. For example, 3 exhibits bands at 1580 and 1267 cm−1, attributed to the C–N and the C–S bond stretch vibrations the dmp2S− ligands, respectively. 1H NMR spectra of the complexes display resonances which correspond to the protons of their respective ligands (Fig. S1 and S2†). For example, 3 features one singlet for the aromatic protons of the heterocyclic rings of its dmp2S− ligands at +6.42 ppm and two singlets in the region +2.42 to +1.78 ppm attributed to the two types of CH3-group substituents present in the complex, due to the facial arrangement of its chelating ligands. Interestingly, 1H NMR spectra of all complexes remain unchanged for extended periods of time, providing an indication for their long-term stability in organic solvents.
Apart from 1–4, heteroleptic complex [Co(pymt)3(PPh3)] (5) was also synthesized (Scheme 2). This was obtained by adding to a solution of CoCl2·6H2O in CH3OH a mixture of 2 equiv. pymtH and 1 equiv. PPh3, in the presence of 2 equiv. KOH, at 50 °C and under atmospheric conditions (Section S1.2 in the ESI†). Single-crystal X-ray diffraction shows a six-coordinate Co(III) ion surrounded by two chelating κS,N-pymt, a monodentate κS-pymt, and a PPh3 ligand (Fig. 2). The metal center exhibits distorted octahedral coordination geometry due to the steric bulk of PPh3 ligand and the small chelate ring angles of pymt− ligands. Noticeably, Co–S bond lengths (Table S3†) are longer than those of 1–4 (average values: 2.310 vs. 2.280 Å, respectively). 1H NMR spectrum of 5 suggests its stability in organic solvents, displaying resonances in the region +8.81 to +6.84 ppm, which correspond to the aromatic protons of thioamidato and phosphine ligands (Fig. S3†).
![]() | ||
Fig. 2 Molecular structure of 5 determined by single crystal X-ray diffraction. Thermal ellipsoids are drawn at the 35% probability level. All H atoms are omitted for clarity. |
A series of pyridine-thioamidato Co(III) complexes 6–8 was also synthesized (Scheme 2). [Co(2-mpy)3] (6) was obtained in a similar manner to its homologous complexes 1–4. Its X-ray crystal structure was reported in the past and, therefore, is not discussed herein.49–51 [Co(2-mpy)2(dppe)]PF6 (7) was obtained from the reaction of CoCl2·6H2O with 2 equiv. 2-mpyH and 1 equiv. dppe in CH3OH, in the presence of (CH3CH2)2NH, at 60 °C and under atmospheric conditions, followed by addition of NH4PF6 (Section S1.2 in the ESI†). Single-crystal X-ray diffraction shows a cationic complex with its metal center chelated by two κN,S-2-mpy− and a κP,P′-dppe ligand (Fig. 3). The coordination geometry of the metal center deviates from the ideal octahedral geometry due to the combined effect of the steric bulk of dppe ligand and the small chelate ring angles of 2-mpy− ligands. Average Co–S and Co–N bond lengths are equal to 2.29 and 1.96 Å, respectively (Table S4†).52 Finally, [Co(2-mpy)2(py)2]PF6 (8) was synthesized by adding to a solution of CoCl2·6H2O in CH3OH 2 equiv. Na+(2-mpy)− and 1 equiv. pyridine, at room temperature and under atmospheric conditions, followed by addition of NH4PF6 (Section S1.2 in the ESI†). The cation of 8 consists of a Co(III) ion in a pseudo-octahedral coordination environment surrounded by two chelating κN,S-2-mpy and two py ligands at cis positions (Fig. 3). Average Co–S and Co–N bond lengths of thioamidato ligands are equal to 2.30 and 1.94 Å, respectively (Table S4†).
FTIR spectra of 7 and 8 confirm the presence of PF6− ions, as revealed by high intensity peaks at 840 cm−1, which are characteristic of the ν(P–F) stretching vibrations. In the 1H NMR spectrum of 7, signals of the aromatic protons of the thioamidato and diphosphine ligands appear as complex multiples in the +8.39 to +6.40 ppm region, while the CH2–CH2 backbone of dppe appears as triplet at +1.30 ppm (Fig. S5†). For 8, the appearance of a broad singlet resonance at +8.27 ppm is assigned to the ortho-protons of the pyridine ligands (Fig. S6†). Resonances of the protons of its thioamidato ligands appear in the +7.50 to +6.58 ppm range; these are shifted up-field compared to the corresponding signals of the thioamidato ligands of 7, which bears the same thioamidato ligands and a similar structural motif.
Solventa | Conversionb (%) | 1a (%) | 1b (%) | 1c (%) | 1d (%) |
---|---|---|---|---|---|
a Reaction conditions: 1a (0.2 mmol), CH3NHNH2 (5 equiv.), 2 (1% mol), in various solvents, at 70 °C and for 4 h in a 5 mL sealed tube. b Conversion percentages (%) were measured by 1H NMR spectroscopy after addition of a specific amount of 1,3-dimethoxybenzene as internal standard. c Product yields were measured by 1H NMR spectroscopy based on the integration of the appropriate proton peaks. | |||||
CH3OH | 95 | 5 | 95 | — | — |
CH3OH (2 h) | 43 | 57 | 43 | — | — |
CH3OH (r. t.) | — | 100 | — | — | — |
CH3OH (50 °C) | 63 | 37 | 63 | — | — |
CH3OH (3 equiv. CH3NHNH2) | 48 | 52 | 48 | — | — |
CH3CH2OH | 54 | 46 | 54 | — | — |
CH3CN | <5 | >95 | — | — | — |
CH3COOCH2CH3 | <5 | >95 | — | — | — |
CH2Cl2 | <5 | >95 | — | — | — |
OC(OCH3)2 | <5 | >95 | — | — | — |
THF | <5 | >95 | — | — | — |
Toluene | <5 | >95 | — | — | — |
H2O | <5 | >95 | — | — | — |
2 (mol%) | Conversionb (%) | 1b (%) | TON | TOF (h−1) |
---|---|---|---|---|
a 1a (0.2 mmol), CH3NHNH2 (5 equiv.), in CH3OH at 70 °C and for 4 h. b Conversion percentages (%) were measured by 1H NMR spectroscopy after addition of a specific amount of 1,3-dimethoxybenzene as internal standard. c Product yields were measured by 1H NMR spectroscopy based on the integration of the appropriate proton peaks. | ||||
1 | 95 | 95 | 95 | 23 |
0.5 | 50 | 50 | 100 | 25 |
0.2 | 12 | 12 | 60 | 15 |
No catalyst | <10 | <10 | — | — |
We next compared the catalytic efficiency of 1–8 in the reduction of 1a under the optimized reaction conditions, i.e., 1 mol% catalyst, 5 equiv. CH3NHNH2, in CH3OH at 70 °C and for 4 h. In all cases, p-toluidine (1b) was formed as the only product (Fig. S7–S14†). Among homoleptic pyrimidine-thioamidato Co(III) complexes 1–4, 2 bearing EW CF3-group substituents on its ligands was found to be the most efficient catalyst, giving almost quantitative conversion to 1b (95% yield) (Fig. 4 and Table S5†). 1 bearing unsubstituted thioamidato ligands exhibited lower catalytic activity (65% yield), while 3 and 4, having ligands with ED group substituents (CH3 and NH2, respectively), were less efficient (55% and 40% yields, respectively). These results clearly indicate the beneficial effect of EW group substituents on the thioamidato ligands in the catalytic activity of the Co(III) complexes. Interestingly, [Co(pymt)3(PPh3)] (5), bearing an additional PPh3 ligand compared to [Co(pymt)3] (1), displayed catalytic activity that is comparable to that of 2, giving conversion of 1a to 1b in 85% yield.
Pyridine-thioamidato Co(III) complexes 6–8 were also found to be efficient catalysts, with homoleptic complex 6 exhibiting the highest catalytic activity (giving 1b in 70% yield). Replacement of a chelating thioamidato ligand in 6 with a bulky bidentate chelating dppe ligand to give 7 resulted in significant decrease of catalytic activity (40% yield). Heteroleptic complex 8 exhibited moderate catalytic activity similar to 6 (55% yield).
It should be mentioned that under the employed reaction conditions but in the absence of any catalyst formation of 1b was also observed, but in a negligible amount (yield < 10%, Table S5†). Also, the use of the tris-chelating Co(III) complex [Co(acac)3] as catalyst also resulted in the formation of 1b but with very low yield (Table S5†). Therefore, overall, these results clearly demonstrate the capacity of 1–8 to efficiently catalyze the reduction of 1a by CH3NHNH2 giving 1b in high yield and selectivity.
In an effort to further explore the applicability of our catalytic reaction protocol, 2 was also tested in the reduction of aliphatic nitroalkane 15a and methyl (E)-3-(4-nitrophenyl)acrylate 16a (Scheme 3). Nitroalkane 15a gave a mixture of unidentified products in ca. 10% conversion, in which the corresponding aliphatic amine was not observed (based on the 1H NMR spectrum of the crude reaction mixture). Interestingly, 2 displayed exceptional catalytic activity towards the chemoselective nitro-group reduction of 16a, yielding the desired amine 16b as the major product (68% isolated yield, Fig. S29†). Remarkably, no CC reduction was observed (based on the 1H NMR spectrum of the crude reaction mixture). Therefore, these results strongly confirm the key role that 2 can play in the chemoselective synthesis of valuable amine derivatives.
Finally, we attempted the catalytic reduction of azomycin (17a), an antibiotic used to treat bacterial infections. Surprisingly, as revealed by 1H NMR spectroscopy, the corresponding N-hydroxylamine derivative (17c) was observed as the only product (Fig. S30†). This was found to be stable under the employed reduction conditions, not getting reduced further to the desired amine even after longer reaction times. This is in agreement with our previous study on the photoreduction of nitroarenes into the corresponding N-hydroxylamines.53
Finally, to test our catalytic reaction protocol in lab-scale conditions, the reductions of 1a and 6a were attempted in a 1 mmol-scale, using 2 mol% 2, in the presence of 6 equiv. CH3NHNH2, in 10 mL of CH3OH at 70 °C in a sealed reaction tube. Upon completion of each reaction, the solvent was removed under reduced pressure, and, after addition of CH3COOCH2CH3 (∼15 mL), the mixture was filtered over a short silica gel pad. Following solvent evaporation, the corresponding anilines 1b and 6b were purified by column chromatography and isolated in yields of 71% and 75%, respectively (Scheme 3).
Among the pyridine-thioamidato Co(III) complexes 6–8, the diphosphine-substituted complex 7 was found to be less efficient than 6 and 8. At first glance, this comes in contrast to the enhancement of catalytic activity observed with the introduction a phosphine ligand in 5. Possibly, the steric bulk of chelating dppe in 7 creates a sterically congested metal coordination environment that hinders the binding of CH3NHNH2 and/or the substrate. Interestingly, 8 bearing two pyridine ligands appears to exhibit similar catalytic activity to homoleptic complex 6.
![]() | ||
Fig. 5 UV-vis absorption spectra of 1–5 in CH3OH (2 × 10−4 M). Spectra in the lower energy region are given in the inset. |
We followed the changes of the spectral characteristics of 2 in CH3OH upon addition of increasing amounts of CH3NHNH2. Colour changes of the solution were observed (from dark brown to light brown), with concomitant changes in its d–d absorption band. Upon addition of 0.25 equiv. CH3NHNH2, the intensity of the d–d absorption band was reduced, while addition of 0.5 equiv. CH3NHNH2 resulted in its complete disappearance (Fig. 6). No further changes were observed after addition of more than 0.5 equiv. CH3NHNH2. Considering that 1 equiv. CH3NHNH2 can generate 4e− and 4H+ before its conversion to N2, the 1:
0.5 stoichiometry of 2
:
CH3NHNH2 suggests the reduction of the Co(III) complex to a lower oxidation-state Co species (possibly Co(I)). Such a species is also suggested to be formed in the catalytic reaction and participate in the reduction of the nitroarenes. An analogous spectrophotometric titration was conducted for 3, a complex that showed lower catalytic activity. In this case, we did not discern analogous spectral changes in its d–d absorption band (Fig. S36†), a fact that might be related to the difficulty of 3 to form a reduced Co species (see below in Cyclic voltammetry).
Complex |
E
III/IIpa![]() |
E
III/IIpc![]() |
E
III/II1/2![]() |
ΔEIII/II | E II/Ipa |
---|---|---|---|---|---|
a E pa is anodic peak potential (V). b E pc is cathodic peak potential (V). c E 1/2 = (Epa + Epc)/2. All potentials are referenced vs. the Fc/Fc+ couple with E1/2(Fc/Fc+) = 0.409 V. | |||||
1 | −0.862 | −0.775 | −0.819 | −0.087 | −2.254 |
2 | −0.629 | −0.536 | −0.583 | −0.093 | −1.790 |
3 | −0.955 | −0.717 | −0.836 | −0.238 | −2.376 |
5 | −0.860 | −0.786 | −0.823 | −0.074 | −2.219 |
6 | −1.002 | −0.894 | −0.948 | −0.108 | −2.348 |
7 | −0.971 | −0.898 | −0.935 | −0.073 | −1.778, −2.204 |
8 | −0.998 | −0.899 | −0.949 | −0.099 | −2.451 |
Among all complexes, 2 displays the most easily accessible Co(II)/Co(I) reduction potential (Fig. 7). This can be attributed to the EW effect of the CF3-group substituents of its ligands, which tend to stabilize better the lower metal oxidation states.59 Therefore, its high catalytic activity can be related to its easier access to a Co(I) species during the catalytic reaction. However, although the ease of Co(II)-to-Co(I) reduction is a factor of great importance for enhanced catalytic activity, other factors should also be favorable, such as hemilability/lability of the ligands. Indeed, 7 comes to verify this scenario (Fig. S37†): although displaying a similar redox profile with 2, its catalytic activity is diminished.
A variable scan-rate cyclic voltammetry study was also conducted for 2 in CH3CN (Section S2.3.3 Cyclic voltammetry†) which confirmed the homogeneous nature of the Co complexes during the redox transformations, also suggested to take place in the catalytic reduction of nitroarenes, excluding the formation of metal-based nanoparticles.
The redox behavior of 2 in CH3CN upon addition of progressively increasing amounts of CH3NHNH2 was also studied. Upon addition of 1 equiv. CH3NHNH2, the initial redox characteristics of the complex essentially remain unaffected but two additional redox peaks also appear at −1.2 and −2.4 V (Fig. 8). The peak at −1.2 V can be attributed to the Co(III)/Co(II) redox couple of a new Co species generated from the interaction of 2 with CH3NHNH2 which, based on the UV-vis spectroscopy and HRMS data, possibly contains a dangling thioamidato ligand and/or a coordinated CH3NHNH2. The addition of more equivalents of CH3NHNH2 results in an increase of the height of this peak, consistent with the increase of the concentration of this new species in solution. The occurrence of the Co(III)/Co(II) reduction wave of this new species at a more negative potential than the corresponding redox wave of 2 is anticipated, considering the strong σ-donating nature of the coordinated CH3NHNH2 ligands. The addition of 1 equiv. CH3NHNH2 also resulted in the increase of the reversibility of the Co(III)/Co(II) reduction wave of 2, bringing the Ia/Ic ratio closer to 1 (with respect to the Ia/Ic ratio of 2 in the absence of CH3NHNH2). The same observation was made when the experiment was conducted in CH3OH, the solvent of the catalytic reaction. The second new redox peak at −2.4 V appearing upon addition of CH3NHNH2 can be assigned to the Co(II)/Co(I) redox couple of the new species. In addition, a shift of the initial Co(II)/Co(I) reduction peak of 2 to a more accessible potential is also observed. This shift becomes greater upon increased amounts of CH3NHNH2 (reaching ∼0.04 V at 4 equiv. CH3NHNH2), a result that can be attributed to the ability of CH3NHNH2 to provide suitable conditions in the solution (e.g., solvation, polarity) to facilitate the Co(II)/Co(I) reduction.
To examine the possibility of the potential formation of a Co-hydride species and the occurrence of metal-mediated proton transfer during the catalytic reduction of nitroarenes, a CV study was conducted with 2 in CH3CN upon addition of progressively increasing amounts of CH3COOH as H+ source. A new irreversible reduction wave is observed at a less negative potential than that of the original Co(II)/Co(I) redox couple (Fig. S38†).60,61 Also, an irreversible reduction wave started growing at more negative potentials, with progressive increase of the current of the wave with increasing amounts of H+, which is tentatively assigned to the electrocatalytically formed H2. This behavior is typical for hydride-containing transition metal complexes that participate in proton transfer processes. Therefore, 2 in the catalytic reduction of nitroarenes to anilines has the capacity to participate in proton transfer processes (provided either by CH3OH or CH3NHNH2), such as protonation of the heterocyclic N atom a thioamidato ligand or formation of a Co-hydride species, potentially key intermediates for the progress of the catalytic reaction.
The catalytic reduction of 1a by 2 in the presence of CH3NHNH2 was also monitored by GC analysis. Analysis of the supernatant volume of gas above the reaction mixture revealed the presence of CH4 and N2 (Fig. S40†). The formation of these species is an immediate result of the interaction of CH3NHNH2 with 2, as no evolution of gases was observed when CH3NHNH2 was tested under the same experimental conditions and in the absence of 2. These results, in combination with our NMR study which showed that CH3NHNH2 under the same conditions is fast transformed to CH2NNH2, suggest the occurrence of a CH3NHNH2-coordination activation pathway along with the catalytic transformation of nitroarenes (Fig. S31†). Analogous diazoalkane species were reported previously for other transition metal catalysts.64
![]() | ||
Scheme 5 Suggested mechanism of the reduction of nitroarenes to anilines by CH3NHNH2 catalyzed by 2. |
In the first step of the coordination-induced activation of CH3NHNH2 pathway (Scheme 4), transformation of a chelating thioamidato (κS,N-LNS) ligand of 2 to a monodentate ligand is suggested to take place. The transformation may be assisted by the inherent hemilability of the thioamidato ligands as well as the protonation of the uncoordinated donor atom of the dangling thioamidato ligand forming a five-coordinate Co(III) intermediate ([CoIII]+). The monodentate thioamidato ligand is suggested to remain S-bound, with the most favorable protonation site being its heterocyclic N atom (κS-LNHS+).51 This is in agreement with the fact that the lower oxidation-state Co species, formed in a subsequent step of the mechanism of the reduction of nitroarene by CH3NHNH2 (discussed below), favors the S-coordination of the ligand. Subsequently, CH3NHNH2 is suggested to coordinate to the [CoIII]+ intermediate ([CoIII–CH3NHNH2]+). This CH3NHNH2-adduct can mediate the decomposition of CH3NHNH2 with concomitant liberation of H2 and CH4. Formation of H2 might lead to a Co–(CH3NNH) species, from which the diazene can uncoordinate restoring the [CoIII]+ intermediate. The Co–(HN
NH) species, formed upon evolution of CH4, decomposes to N2 resulting to a Co(III)-hydride intermediate ([CoIII–H]+) which, upon protonation, might regenerate the [CoIII]+ intermediate. The suggestion of the coordination-induced activation of CH3NHNH2 pathway is based on the detection of the pertinent gases and the observation of CH2
NNH2 in the reaction mixture. Also, it agrees with recent reports on the activation of NH2NH2 by transition metal complexes, proposed to take place by weakening of the N–H bonds of NH2NH2 upon its coordination to the metal center, enabling the release of H2 and the formation of metal-diazene and metal-hydride intermediates.65,66
A Co-hydride species mentioned in the abovementioned suggested pathway is suggested to be involved in the reduction of nitroarene (ArNO2) to aniline (ArNH2) by 2 in the presence of CH3NHNH2. In addition, cyclic voltammetry and catalytic activity studies suggest the importance of easy access to a Co(I)-based species for nitroarene reduction. Both species possibly play important roles in the catalytic transformation.51,52
Therefore, in our suggested mechanism (Scheme 5), initially, 2 ([CoIII]0) undergoes 2e− reduction by CH3NHNH2 and protonation, leading to a five-coordinate Co(I) intermediate ([CoI]−). This negatively charged species is expected to be stabilized by the EW CF3-group substituted thioamidato ligands of 2, but not by ED group substituted ligands. Considering the higher catalytic activity of 2 compared to 3 and 4 (which comprise CH3- and NH2-group substituents in the ligands), the particular step might have a high impact on the efficiency of the catalytic reduction. Subsequent reaction of the resulting Co(I) species with protons results in the formation of a Co(III)-hydride ([CoIII–H]). This intermediate is suggested to be formed via direct protonation of the Co(I) center and not via intramolecular proton transfer from the protonated thioamidato ligand to the Co center (the latter could lead to the κN,S-coordination of the thioamidato ligand which would destabilize the low oxidation state Co(I) center). Considering that hydride formation through direct protonation is expected to have high activation energy,51 the particular step might also have an impact on the rate and efficiency of the transformation. The [CoIII–H] intermediate might promote hydride transfer to the incoming substrate ArNO2 (by insertion) and form a new intermediate which, upon intramolecular proton transfer from the protonated thioamidato ligand and subsequent protonation, releases H2O. After an additional protonation, ArNHOH is released (observed experimentally as intermediate of the transformation), regenerating [CoIII]0. Intramolecular proton transfer and subsequent κN,S-coordination of the thioamidato ligand in [CoIII]0 is consistent with the stabilization of the amidato form of the ligand coordinated to a high oxidation state Co(III) center. The product of the transformation (ArNHOH) is suggested to be inserted once more in the catalytic cycle and be converted to the aniline (ArNH2). It should be mentioned that the [CoIII–H] intermediate, except form the abovementioned catalytic cycle, might also participate in a competitive reaction, as its direct protonation would release H2 (Scheme 5 in the center).
Generally, in the reduction of nitroarenes to anilines, the dimers azoxyarene (1c), and a mixture of 1,2-diphenylhydrazine (1d) and azobenzene may also be formed (Table 1), which are subsequently transformed to the aniline (indirect route). Although we did not detect these species in the reaction mixtures, in order to exclude their participation in the mechanism of the reaction, they were used as substrates under the employed experimental conditions. No conversion into the corresponding amines was observed, a result that supports the proposed direct route, with the reduction of nitroarenes into anilines through the corresponding observed N-arylhydroxylamine (ArNHOH) intermediates. Finally, nitrosoarene (ArNO) intermediates could also be formed during the catalytic reaction, which could be further reduced to the final product under either catalytic or non-catalytic conditions.40,67 Herein, their existence was not experimentally verified; however, such reduction pathway cannot be excluded. Overall, based on the above discussion, it becomes clear that for the herein presented catalytic system of thioamidato Co(III) complexes with CH3NHNH2, the combination of (a) thioamidato ligands exhibiting high hemilability or lability and (b) Co(III) complexes exhibiting easily accessible Co(II)/Co(I) reduction potentials, are both essential for achieving high catalytic efficiency. These properties are more favorable in 2 (bearing EW CF3-group substituents on its thioamidato ligands) and their synergy results in a very efficient catalyst, while in other complexes their synergy is not so efficient resulting in less effective catalysts.
Overall, the herein-presented results highlight the competency of the thioamidato/Co(III) scaffold in the highly efficient and chemoselective reduction of nitroarenes to anilines, and lay the grounds for its utilization in the synthesis of fine and specialty chemicals or pharmaceuticals, as well as in other multi-electron redox transformations.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1995929 (1), 1995931 (3), 1995932 (4), 1995933 (5), 1995934 (7) and 1995935 (8). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt02923a |
This journal is © The Royal Society of Chemistry 2024 |