Next Article in Journal
Study on Structure Activity Relationship of Natural Flavonoids against Thrombin by Molecular Docking Virtual Screening Combined with Activity Evaluation In Vitro
Next Article in Special Issue
Fe-catalyzed Decarbonylative Alkylative Spirocyclization of N-Arylcinnamamides: Access to Alkylated 1-Azaspirocyclohexadienones
Previous Article in Journal
Understanding Olive Oil Stability Using Filtration and High Hydrostatic Pressure
Previous Article in Special Issue
Iron-Catalyzed C(sp2)–C(sp3) Cross-Coupling of Aryl Chlorobenzoates with Alkyl Grignard Reagents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Cyclopentadienone Iron Tricarbonyl Complexes-Catalyzed Hydrogen Transfer in Water

by
Daouda Ndiaye
1,2,
Sébastien Coufourier
1,
Mbaye Diagne Mbaye
2,
Sylvain Gaillard
1 and
Jean-Luc Renaud
1,*
1
Normandie Univ., LCMT, ENSICAEN, UNICAEN, CNRS, 6 boulevard du Maréchal Juin, 14050 Caen, France
2
Université Assane Seck de Ziguinchor, BP 523, Ziguinchor, Senegal
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(2), 421; https://doi.org/10.3390/molecules25020421
Submission received: 30 December 2019 / Revised: 15 January 2020 / Accepted: 17 January 2020 / Published: 20 January 2020
(This article belongs to the Special Issue Recent Advances in Iron Catalysis)

Abstract

:
The development of efficient and low-cost catalytic systems is important for the replacement of robust noble metal complexes. The synthesis and application of a stable, phosphine-free, water-soluble cyclopentadienone iron tricarbonyl complex in the reduction of polarized double bonds in pure water is reported. In the presence of cationic bifunctional iron complexes, a variety of alcohols and amines were prepared in good yields under mild reaction conditions.

Graphical Abstract

1. Introduction

The reduction of polarized C=X bonds is an important process, both in industry and in academia, for the synthesis of fine chemicals, perfumes, agrochemicals, and pharmaceuticals [1,2,3,4,5,6,7,8]. To avoid the use of stoichiometric amounts of borohydrides or aluminium hydrides, metal-catalyzed pathways to amines and alcohols have been introduced [9,10,11,12,13,14]. These procedures consist of hydrogenation, hydrosilylation, and transfer hydrogenation (with iso-propanol or formic acid) and involve mainly platinum complexes [9,10], but recent contributions highlighted the rise of Earth-abundant metals for such reductions [11,12,13,14]. Hydrogenation is the most atom economical approach, but requires hydrogen gas handling and consequently implies some safety issues. Hydrogen transfer (TH) is an alternative pathway and a more practical tool. Alcohols and formic acid (or formates) are among the most advantageous hydride donors.
Water is a non-toxic, non-flammable, non-explosive and also an economically relevant solvent [15,16]. Water-soluble organometallic complexes have attracted some interest because of the environmentally acceptable process, the simple product separation and, in some reactions, the possibility to control the selectivity by adjusting the pH [15,16]. Despite these advantages, the use of water in catalysis, and more specifically, in reduction, still constitutes a challenge and is underexplored compared to organic solvent [17,18]. Hydrogenation of ketones and imines [19], and reductive amination [20] in water have been reported with few iron complexes. As an example; our group has disclosed the first water-soluble and well-defined cyclopentadienone iron complex able to catalyze the reduction of aldehydes, ketones, and 2-substituted dihydroisoquinolines in pure water at 85–100 °C under hydrogen pressure (Figure 1, for the first synthesis of a water-soluble cyclopentadienone iron complex, see [19]). Little is known on the transfer hydrogenation with Earth-abundant complexes, while formates are used by enzymes for enantioselective reduction. To the best of our knowledge, excepted the hydrogen transfer reduction of heterocyclic compounds with formic acid catalyzed by a cobalt-phosphines complex [20], no reduction of polarized C=X bonds (aldehydes, ketones, and imines) with formic acid derivatives has been yet reported. Toward this objective, we thought of developing new water-soluble non-phosphine ligand iron complexes for the reduction of polarized bonds in the presence of formates or formic acid in pure water.
In 2015, we introduced in catalysis the tricarbonyl iron complex Fe2 bearing a diaminocyclopentadienone ligand [21]. Compared to other cyclopentadienone iron carbonyl complexes, this phosphine-free iron complex has, to the best of our knowledge, the highest catalytic activities to date in reductive amination [21], in chemoselective reduction of α,β-unsaturated ketones [22], in the hydrogenation of carbon dioxide [23], in alkylation of ketones [24,25,26,27], amines [25,28], oxindoles [29], indoles [25,30] and alcohols [31,32]. In our ongoing interest in reduction and alkylation, we thought that a water-soluble analog of Fe2 would be more active than our previous water-soluble cyclopentadienone iron complex Fe3 (Figure 1). In this work, we report on the synthesis and application of two water-soluble cyclopentadienone iron complexes in the reduction of aldehydes and in reductive amination in pure water.

2. Results and Discussion

2.1. Synthesis of Complexes

To develop water-soluble iron complexes, we selected a diaminocyclopentadienone ligand bearing ammonium functionalities [33]. The tetraamines 2 and 4 were prepared from diethyloxalate and N,N-dimethylpropylenediamine and N-aminopropylenemorpholine via an amidation followed by a reduction in good overall yield (93 and 95%, respectively). The corresponding aminocyclopentadienone ligands 1 and 2 were then prepared by reacting the amines 2 and 4 with the cyclopentatrienone in refluxing methanol for 16 h and were isolated in moderate yield (62% and 88%, respectively, Scheme 1). The complexes Fe6 and Fe7 were synthesized in 48% and 76% yield by simple heating of the corresponding amino ligand with [Fe2(CO)9] in refluxing toluene (Scheme 1). Finally, the water-soluble bifunctional iron complexes Fe4 and Fe5 bearing ionic frameworks were obtained in almost quantitative yields after a subsequent alkylation of the pendant amines with iodomethane (Scheme 1) [33]. These complexes were fully characterized by 1H-, 13C-NMR, and IR spectroscopies (see Supplementary Materials). These analyses showed that complexes Fe2 and Fe4-Fe7 have similar features. The back donation from the metal center to the CO ligands is more significant than in the Knölker’s complex Fe1 [34,35]. Thus, the CO stretching frequencies were at 2032, 1961, and 1919 cm−1 and at 2015 and 1967 cm−1 in the neutral complexes Fe6 and Fe7, respectively, at 2038 and 1955 cm−1 and at 2034 and 1957 cm−1 in the ionic complexes Fe4 and Fe5, respectively. These frequencies are comparable to those of the analog Fe2 (2027, 1962 and 1947 cm−1) and lower than those of the Knölker’s complex Fe1 (2061, 2053, and 1987 cm−1) or its water-soluble analog Fe3 (2066, 2016, and 1996 cm−1) [19].

2.2. Iron-Catalyzed Reduction of Carbonyl Compounds

With these complexes in hands, we evaluated their catalytic activities in the reduction of 4-methoxybenzaldehyde as a benchmark reaction. Various methods of activation can be used with the cyclopentadienone iron carbonyl complexes [34,35,36,37,38,39,40,41,42,43,44]. We applied in this work the activation with Me3NO as oxidant [34,35,36,37,38,39,40,41,42]. Our first attempt with formic acid, in the presence of 2 mol % of Fe4 and 2.5 mol % of Me3NO at 100 °C for 24 h in 2 mL of pure water (concentration of 0.5 M), was unsuccessful as no reduction was noticed (entry 1, Table 1). In sharp contrast, in the same reaction conditions, complete conversions were obtained with different formate salts (entries 2–5, Table 1). Without a hydride donor or iron complex, no reduction occurred (entries 6–7, Table 1). Decreasing the reaction time (entries 8 and 11, Table 1), the catalyst loading (entry 13) and the amount of formate (entry 14) led to a drop in the conversion. No variation of the conversion was noticed by lowering the temperature to 80 °C (entries 3 and 9, Table 1), while, at 60 °C, the conversion was only 75% (entry 12). To our surprise, the first generation water-soluble cyclopentadienone iron complex Fe3 did not catalyze the reduction of 4-methoxybenzaldehyde in these conditions, while Fe5 appeared as active as Fe4 (entries 10–15, Table 1). Finally, the best conditions for the reduction of 4-methoxybenzaldehyde into the corresponding alcohol 4a were: 1 mmol of aldehyde, in the presence of five equivalents of sodium formate in 2 mL of water, 2 mol % of Fe4 or Fe5 and 2.5 mol % of Me3NO at 80 °C for 24 h.
Having established the optimized conditions, we delineated the scope of the carbonyl derivatives (Table 2). Both electron-donating (methoxy, methyl, acetal substituents) and electron-withdrawing (nitro, nitrile, and ester substituents) groups were tolerated in this reduction. The corresponding alcohols 4am were isolated in excellent yields in all examples (91–99%, Table 2). No reduction of halogen-carbon bonds in the substituted phenyl group (compounds 4ij) was observed. Other reducible functions, such as ester or nitrile, were preserved in these conditions. Heteroaromatic derivatives, such as pyridine or thiophene carboxaldehyde, furfural, did not impede the catalytic activity and provided the alcohols 4nr in 75–98%yield (Table 2). Finally, to extend the scope, aliphatic aldehydes were also engaged in this reduction and the corresponding alcohols 4sv were isolated in 92–99% yield (Table 2). It is worth to mention that (i) ethanol was used as a co-solvent with some substrates to facilitate the solubility and consequently enhanced the reactivity; and (ii) no reaction occurred in a mixture of water and ethanol without sodium formate.

2.3. Iron-Catalyzed Reductive Amination

Having established a simple protocol for the reduction of aldehydes in water, we thought to extend this work to the synthesis of amines. Amines are usually prepared via the reduction of C=N bonds either in catalytic conditions under hydrogen pressure or in stoichiometric conditions in the presence of aluminum/boron hydride [45]. However, imines are not always easily prepared and cannot be stable. Reductive amination of aldehydes constitutes a direct route to amines, without requiring any purification of the imine intermediate. Many efforts have been devoted to the development of reductive amination [46,47,48]. For example, in iron chemistry, Bhanage described that a combination of iron sulfate and ethylenediaminetetraacetic acid (EDTA) catalyzed a reductive amination under hydrogen pressure (400 psi) in water at elevated temperatures (150 °C) [20]. Beller reported a reductive amination with anilines catalyzed by Fe2(CO)9 under high hydrogen pressure and elevate temperature [49]. We have reported that cyclopentadienone iron tricarbonyl complexes [21,34,35] or cyclopentadienyl iron(II) tricarbonyl complex [50] were able to catalyzed the reductive alkylation of various amines and carbonyl derivatives under 5 bar of hydrogen, at 40–70 °C and even room temperature. To avoid the use of a large amount of hydride (and the concomitant formation of wastes) or the handling of gas, hydrogen transfer with formate derivatives appears as a simple and versatile procedure. The reductive alkylation of N-methylbenzylamine with citronellal was chosen as a model reaction for the optimization of the reaction conditions. Three formate salts were tested, and the cation appeared to be crucial for the catalytic activity (entries 1–4, Table 3). Indeed, the ammonium favored both the condensation and the reduction (via the formation of an iminium intermediate). Fe4 and Fe5 provided the alkylated amines with the same conversion and selectivities (entries 1–2, Table 3). Both water-soluble complexes Fe4 and Fe5 could then be used in this reaction (as it was also mentioned in the reduction of aldehydes), but for the rest of the study, we will use Fe5 as pre-catalyst as the overall conversion in reductive amination is somewhat higher. Without a hydride donor, no reduction occurred (entries 5, Table 3), and only the imine was obtained. Decreasing the temperature was detrimental to the catalytic activity as a drop of the conversion in amine was noticed (entries 6–8, Table 3). Finally, an increase of the amount of ammonium formate to 6.5 equivalent furnished the alkylated amine in 70% isolated yield (entry 9, Table 3).
With the optimized conditions in hands, we evaluated some aliphatic and benzylic amines with aromatic and aliphatic aldehydes (Table 4). Whatever the benzylic amine used with citronellal, the isolated yield was good (5ab, 61–70%), while the alkylated amine 5c was obtained in a low yield (21%) from the 2-phenylethylamine (Table 4). Amines 5dk were prepared in 11–64% yield from various benzaldehydes. First, as observed previously, no reduction of halogen-carbon bonds in the substituted phenyl group was observed, and the corresponding alkylated amines 5fh were obtained in around 50% yield. The reductive alkylation of N-methylbenzylamine with thiophene carboxaldehyde furnished the corresponding amine in a 64% yield. The reductive amination with electron-rich benzaldehyde and N-methylbenzylamine led to the alkylated amine 5j in very modest yields (13%, Table 4).

2.4. Recycling of the Water-Soluble Iron Complex

One of the main goals, when reactions are carried out in the water, is the study of the recyclability of the complex used. Due to the high solubility of the iron complexes, Fe45 in water, separation and recycling should be possible to perform. Fe5-catalyzed, the reductive alkylation of N-methylbenzylamine with citronellal, was chosen as a model reaction for this study. At the end of the first run, ethyl acetate was added under an argon atmosphere to extract the organic compounds, and the aqueous phase was re-engaged directly in another run after the addition of ammonium formate, amine, and aldehyde (Table 5). As showed in Table 5, catalytic activity was maintained after five runs without any decrease in the conversion.

3. Materials and Methods

All air- and moisture-sensitive manipulations were carried out using standard vacuum line Schlenk tubes techniques. All solvent and substrates were degassed prior to use by bubbling argon gas directly in the reaction medium. Other solvents and chemicals were purchased from different suppliers and used as received. Deuterated solvents for NMR spectroscopy were purchased from Sigma Aldrich (Saint-Quentin Fallavier, France) and used as received. NMR spectra were recorded on a 500 MHz Bruker spectrometer. Proton (1H) NMR information is given in the following format: multiplicity (s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet), coupling constant(s) (J) in Hertz (Hz), number of protons, type. The prefix app is occasionally applied when the true signal multiplicity was unresolved, and br indicates the signal in question broadened. Carbon (13C) NMR spectra are reported in ppm (δ) relative to CDCl3 unless noted otherwise. Infrared spectra were recorded over a PerkinElmer Spectrum 100 FT-IR Spectrometer using neat conditions. HRMS analyses were performed by using the Laboratoire de Chimie Moléculaire et Thioorganique analytical Facilities.

3.1. General Procedure for the Reduction of Aldehydes

In a dried flamed Schlenk tube under argon, the corresponding aldehyde (1 equiv.) and sodium formate (5 equiv.) were mixed in water (0.5 M solution). The iron complex Fe4 (2 mol %) and Me3NO (2.5 mol %) were then added. The mixture was stirred and heated at 80 °C for 24 h. After cooling down to room temperature, the resulting solution was quenched with a saturated aqueous solution of sodium bicarbonate and extracted three times with ethyl acetate. The organic phase was dried over MgSO4, filtrated, and concentrated under vacuum to afford the crude product. Purification by flash chromatography on silica gel furnished the alcohol.

3.2. General Procedure for the Reductive Amination of Aldehydes

In a dried flamed Schlenk tube under argon, the aldehyde (1 equiv.), the amine (2 equiv.) and ammonium formate (6.5 equiv.) were mixed in water (0.5 M solution). The iron complex Fe5 (2 mol %) and Me3NO (2.5 mol %) were then added. The mixture was stirred and heated at 90 °C for 24–48 h. After cooling down to room temperature, the resulting solution was quenched with a saturated aqueous solution of sodium bicarbonate and extracted three times with ethyl acetate. The organic phase was dried over MgSO4, filtrated, and concentrated under vacuum to afford the crude product. Purification by flash chromatography on silica gel furnished the amine.

4. Conclusions

In conclusion, we have described the application of water-soluble cyclopentadienone iron tricarbonyl complexes in the reduction of aldehydes and in reductive amination under hydride transfer conditions in pure water. Recyclability of iron complex Fe5 was also demonstrated in a model reductive amination. This system tolerated a variety of functional groups such as halides, ethers, heteroaromatic derivatives without impeding the chemical yields. These water-soluble iron complexes allow an efficient, green, and practical procedure for the synthesis of amines and alcohols.

Supplementary Materials

The following are available online. Table S1: Optimization of the reaction conditions for aldehyde reduction by hydride transfer. Table S2: Optimization of the reaction conditions for reductive amination by hydride transfer. Figure S1–S64: The 1H, 13C, and 19F-NMR spectra of compounds.

Author Contributions

Conceptualization—J.-L.R. and S.C.; methodology—J.-L.R.; validation—J.-L.R., S.G. and M.D.M.; formal analysis—S.C. and D.N.; investigation—S.C. and D.N.; resources—J.-L.R.; data curation—S.C. and D.N.; writing—original draft preparation—J.-L.R.; writing—review and editing—J.-L.R., S.G. and M.D.M.; visualization—J.-L.R., S.G. and M.D.M.; supervision—J.-L.R.; project administratio—J.-L.R.; funding acquisition—J.-L.R. and M.D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We gratefully acknowledge financial support from the “Ministère de la Recherche et des Nouvelles Technologies”, Normandie Université, CNRS, “Région Normandie”, and the LABEX SynOrg (ANR-11-LABX-0029). Ademe Agency is acknowledged for a grant to S. C. and Coopération Française-Sénégal for a grant to D. N.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Klingler, F.D. Asymmetric Hydrogenation of Prochiral Amino Ketones to Amino Alcohols for Pharmaceutical Use. Accounts Chem. Res. 2007, 40, 1367–1376. [Google Scholar] [CrossRef]
  2. Hems, W.P.; Groarke, M.; Zanotti-Gerosa, A.; Grasa, G.A. [(Bisphosphine) Ru(II) Diamine] Complexes in Asymmetric Hydrogenation: Expanding the Scope of the Diamine Ligand. Accounts Chem. Res. 2007, 40, 1340–1347. [Google Scholar] [CrossRef] [PubMed]
  3. Johnson, N.B.; Lennon, I.C.; Moran, P.H.; Ramsden, J.A. Industrial-Scale Synthesis and Applications of Asymmetric Hydrogenation Catalysts. Accounts Chem. Res. 2007, 40, 1291–1299. [Google Scholar] [CrossRef] [PubMed]
  4. Saudan, L.A. Hydrogenation Processes in the Synthesis of Perfumery Ingredients. Accounts Chem. Res. 2007, 40, 1309–1319. [Google Scholar] [CrossRef] [PubMed]
  5. Shimizu, H.; Nagasaki, I.; Matsumura, K.; Sayo, N.; Saito, T. Developments in Asymmetric Hydrogenation from an Industrial Perspective. Accounts Chem. Res. 2007, 40, 1385–1393. [Google Scholar] [CrossRef] [PubMed]
  6. Blaser, H.-U.; Studer, M. Cinchona-Modified Platinum Catalysts: From Ligand Acceleration to Technical Processes. Accounts Chem. Res. 2007, 40, 1348–1356. [Google Scholar] [CrossRef]
  7. Blaser, H.-U.; Pugin, B.; Spindler, F.; Thommen, M. From a Chiral Switch to a Ligand Portfolio for Asymmetric Catalysis. Accounts Chem. Res. 2007, 40, 1240–1250. [Google Scholar] [CrossRef]
  8. Shultz, C.S.; Krska, S.W. Unlocking the Potential of Asymmetric Hydrogenation at Merck. Accounts Chem. Res. 2007, 40, 1320–1326. [Google Scholar] [CrossRef]
  9. Wang, N.; Astruc, D. The Golden Age of Transfer Hydrogenation. Chem. Rev. 2015, 115, 6621–6686. [Google Scholar] [CrossRef]
  10. Wang, C.; Wu, X.; Xiao, J. Broader, Greener, and More Efficient: Recent Advances in Asymmetric Transfer Hydrogenation. Chem. Asian J. 2008, 3, 1750–1770. [Google Scholar] [CrossRef]
  11. Wei, D.; Darcel, C. Iron Catalysis in Reduction and Hydrometalation Reactions. Chem. Rev. 2018, 119, 2550–2610. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, W.; Sahoo, B.; Junge, K.; Beller, M. Cobalt Complexes as an Emerging Class of Catalysts for Homogeneous Hydrogenations. Accounts Chem. Res. 2018, 51, 1858–1869. [Google Scholar] [CrossRef] [PubMed]
  13. Ai, W.; Zhong, R.; Liu, X.; Liu, Q. Hydride Transfer Reactions Catalyzed by Cobalt Complexes. Chem. Rev. 2018, 119, 2876–2953. [Google Scholar] [CrossRef] [PubMed]
  14. Filonenko, G.A.; Van Putten, R.; Hensen, E.J.M.; Pidko, E.A. Catalytic (de)hydrogenation promoted by non-precious metals—Co, Fe and Mn: Recent advances in an emerging field. Chem. Soc. Rev. 2018, 47, 1459–1483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lindström, U.M. Stereoselective Organic Reactions in Water. Chem. Rev. 2002, 102, 2751–2772. [Google Scholar] [CrossRef]
  16. Cornils, B.; Hermann, W.A. Aqueous-Phase Organometallic Catalysis; Wiley-VCH: Weinheim, Germany, 2002. [Google Scholar]
  17. Robertson, A.; Matsumoto, T.; Ogo, S. The development of aqueous transfer hydrogenation catalysts. Dalton Trans. 2011, 40, 10304–10310. [Google Scholar] [CrossRef]
  18. Wu, X.; Xiao, J. Aqueous Phase Asymmetric transfer hydrogenation of ketones-a greener approach to chiral alcohols. Chem. Commun. 2007, 24, 2449–2466. [Google Scholar] [CrossRef]
  19. Mérel, D.S.; Elie, M.; Lohier, J.-F.; Gaillard, S.; Renaud, J.-L. Bifunctional Iron Complexes: Efficient Catalysts for C=O and C=N Reduction in Water. ChemCatChem 2013, 5, 2939–2945. [Google Scholar] [CrossRef]
  20. Bohr, M.D.; Bhanushali, M.J.; Nandurkar, N.S.; Bhanage, B.M. Direct Reductive Amination of Carbonyl Compounds with Primary/Secondary Amines Using Recyclable Water-Soluble FeII/EDTA Complex as Catalyst. Tetrahedron Lett. 2008, 49, 965–969. [Google Scholar] [CrossRef]
  21. Thai, T.-T.; Mérel, D.S.; Poater, A.; Gaillard, S.; Renaud, J.-L. Highly active phosphine-free bifunctional iron complex for hydrogenation of bicarbonate and reductive amination. Chem. Eur. J. 2015, 21, 7066–7070. [Google Scholar] [CrossRef]
  22. Lator, A.; Gaillard, S.; Poater, A.; Renaud, J.-L. Iron-Catalyzed Chemoselective Reduction of α,β-Unsaturated Ketones. Chem. Eur. J. 2018, 24, 5770–5774. [Google Scholar] [CrossRef] [PubMed]
  23. Coufourier, S.; Gaillard, S.; Clet, G.; Serre, C.; Daturi, M.; Renaud, J.-L. A MOF-assisted phosphine free bifunctional iron complex for the hydrogenation of carbon dioxide, sodium bicarbonate and carbonate to formate. Chem. Commun. 2019, 55, 4977–4980. [Google Scholar] [CrossRef] [PubMed]
  24. Seck, C.; Mbaye, M.D.; Coufourier, S.; Lator, A.; Lohier, J.-F.; Poater, A.; Ward, T.R.; Gaillard, S.; Renaud, J. Alkylation of Ketones Catalyzed by Bifunctional Iron Complexes: From Mechanistic Understanding to Application. ChemCatChem 2017, 9, 4410–4416. [Google Scholar] [CrossRef]
  25. Polidano, K.; Allen, B.D.W.; Williams, J.M.J.; Morrill, L.C. Iron-Catalyzed Methylation Using the Borrowing Hydrogen Approach. ACS Catal. 2018, 8, 6440–6445. [Google Scholar] [CrossRef]
  26. Bettoni, L.; Seck, C.; Mbaye, M.D.; Gaillard, S.; Renaud, J.-L. Iron-Catalyzed Tandem Three-Component Alkylation: Access to α-Methylated Substituted Ketones. Org. Lett. 2019, 21, 3057–3061. [Google Scholar] [CrossRef]
  27. Latham, D.E.; Polidano, K.; Williams, J.M.J.; Morrill, L.C. One-Pot Conversion of Allylic Alcohols to α-Methyl Ketones via Iron-Catalyzed Isomerization-Methylation. Org. Lett. 2019, 21, 7914–7918. [Google Scholar] [CrossRef]
  28. Lator, A.; Gaillard, S.; Poater, A.; Renaud, J.-L. Well-Defined Phosphine-Free Iron-Catalyzed N-Ethylation and N-Methylation of Amines with Ethanol and Methanol. Org. Lett. 2018, 20, 5985–5990. [Google Scholar] [CrossRef]
  29. Dambatta, M.B.; Polidano, K.; Northey, A.D.; Williams, J.M.J.; Morrill, L.C. Iron-Catalyzed Borrowing Hydrogen C-Alkylation of Oxindoles with Alcohols. ChemSusChem 2019, 12, 2345–2349. [Google Scholar] [CrossRef] [Green Version]
  30. Seck, C.; Mbaye, M.D.; Gaillard, S.; Renaud, J.-L.; Seck-Diouf, C. Bifunctional Iron Complexes Catalyzed Alkylation of Indoles. Adv. Synth. Catal. 2018, 360, 4640–4645. [Google Scholar] [CrossRef]
  31. Bettoni, L.; Gaillard, S.; Renaud, J.-L. Iron-Catalyzed β-Alkylation of Alcohols. Org. Lett. 2019, 21, 8404–8408. [Google Scholar] [CrossRef]
  32. Polidano, K.; Williams, J.M.J.; Morrill, L.C. Iron-Catalyzed Borrowing Hydrogen β-C(sp3)-Methylation of Alcohols. ACS Catal. 2019, 9, 8575–8580. [Google Scholar] [CrossRef] [Green Version]
  33. Coufourier, S.; Gaillard, Q.G.; Lohier, J.-F.; Poater, A.; Gaillard, S.; Renaud, J.-L. Hydrogenation of CO2, Hydrogenocarbonate and Carbonate to Formate in Water using Phosphine Free Bifunctional Iron Complexes. ACS Catal. 2020. [Google Scholar] [CrossRef]
  34. Pagnoux-Ozherelyeva, A.; Pannetier, N.; Mbaye, D.M.; Gaillard, S.; Renaud, J.-L. Knölker’s Iron Complex: An Efficient In Situ Generated Catalyst for Reductive Amination of Alkyl Aldehydes and Amines. Angew. Chem. Int. Ed. 2012, 51, 4976–4980. [Google Scholar] [CrossRef] [PubMed]
  35. Moulin, S.; Dentel, H.; Gaillard, S.; Poater, A.; Cavallo, L.; Lohier, J.-F.; Renaud, J.-L.; Pagnoux-Ozherelyeva, A. Bifunctional (Cyclopentadienone)Iron-Tricarbonyl Complexes: Synthesis, Computational Studies and Application in Reductive Amination. Chem. A Eur. J. 2013, 19, 17881–17890. [Google Scholar] [CrossRef] [PubMed]
  36. Luh, T.-Y. Trimethylamine N-oxide—A versatile reagent for organometallic chemistry. Coord. Chem. Rev. 1984, 60, 255–276. [Google Scholar] [CrossRef]
  37. Moyer, S.A.; Funk, T.W. Air-Stable iron catalyse for the oppenauer type oxidation of alcohols. Tetrahedron Lett. 2010, 51, 5430–5433. [Google Scholar] [CrossRef]
  38. Johnson, T.C.; Clarkson, G.J.; Wills, M. (Cyclopentadienone)iron Shvo Complexes: Synthesis and applications to hydrogen transfer reactions. Organometallics 2011, 30, 1859–1868. [Google Scholar] [CrossRef]
  39. Plank, T.N.; Drake, J.L.; Kim, D.K.; Funk, T.W. Air-Stable, Nitrile-Ligated (Cyclopentadienone)iron Dicarbonyl Compounds as Transfer Reduction and Oxidation Catalysts. Adv. Synth. Catal. 2012, 354, 597–601. [Google Scholar] [CrossRef]
  40. Knölker, H.-J.; Heber, J. Transition Metal-Diene Complexes in Organic Synthesis, Part 18.1 Iron-Mediated [2 + 2 + 1] Cycloadditions of Diynes and Carbon Monoxide: Selective Demetalation Reactions. Synlett 1993, 12, 924–926. [Google Scholar] [CrossRef]
  41. Knölker, H.-J.; Baum, E.; Heber, J. Transition Metal-Diene Complexes in Organic Synthesis, Part 25.1 Cycloadditions of Annulated 2,5-Bis(trimethylsilyl)cyclopentadienones. Tetrahedron Lett. 1995, 36, 7647–7650. [Google Scholar] [CrossRef]
  42. Knölker, H. Trimethylamine N-Oxide—A useful oxidizing reagent. J. Für Prakt. Chem. 1996, 338, 190–192. [Google Scholar] [CrossRef]
  43. Knölker, H.-J.; Goesmann, H.; Klauss, R. A Novel Method for the Demetalation of Tricarbonyliron-Diene Complexes by a Photolytically Induced Ligand Exchange Reaction with Acetonitrile. Angew. Chem. Int. Ed. 1999, 38, 702–705. [Google Scholar] [CrossRef]
  44. Knölker, H.-J.; Baum, E.; Goesmann, H.; Klauss, R. Demetalation of Tricarbonyl(cyclopentadienone)iron Complexes Initiated by a Ligand Exchange Reaction with NaOH—X-Ray Analysis of a Complex with Nearly Square-Planar Coordinated Sodium. Angew. Chem. Int. Ed. 1999, 38, 2064–2066. [Google Scholar] [CrossRef]
  45. Abdel-Magid, A.F.; Mehrman, S.J. A Review on the Use of Sodium Triacetoxyborohydride in the Reductive Amination of Ketones and Aldehydes. Org. Process. Res. Dev. 2006, 10, 971–1031. [Google Scholar] [CrossRef]
  46. Nugent, T.C.; El-Shazly, M. Chiral amine synthesis—Recent developments and trends for enamide reduction, reductive amination, and imine reduction. Adv. Synth. Catal. 2010, 352, 753–819. [Google Scholar] [CrossRef]
  47. Alinezhad, H.; Yavari, H.; Salehian, F. Recent Advances in Reductive Amination Catalysis and Its Applications. Curr. Org. Chem. 2015, 19, 1021–1049. [Google Scholar] [CrossRef]
  48. Gusak, K.N.; Ignatovich, Z.V.; Koroleva, E.V. New potential of the reductive alkylation of amines. Russ. Chem. Rev. 2015, 84, 288–309. [Google Scholar] [CrossRef]
  49. Fleischer, S.; Zhou, S.; Junge, K.; Beller, M. An Easy and General Iron-catalyzed Reductive Amination of Aldehydes and Ketones with Anilines. Chem. Asian J. 2011, 6, 2240–2245. [Google Scholar] [CrossRef]
  50. Lator, A.; Gaillard, Q.G.; Mérel, D.S.; Lohier, J.-F.; Gaillard, S.; Poater, A.; Renaud, J.-L. Room-Temperature Chemoselective Reductive Alkylation of Amines Catalyzed by a Well-Defined Iron(II) Complex Using Hydrogen. J. Org. Chem. 2019, 84, 6813–6829. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Previous water-soluble cyclopentadienone iron complex and new complexes for application reduction.
Figure 1. Previous water-soluble cyclopentadienone iron complex and new complexes for application reduction.
Molecules 25 00421 g001
Scheme 1. Synthesis of the iron complexes Fe4, Fe5, Fe6, and Fe7.
Scheme 1. Synthesis of the iron complexes Fe4, Fe5, Fe6, and Fe7.
Molecules 25 00421 sch001
Table 1. Optimization of the reaction conditions for the aldehyde reduction a.
Table 1. Optimization of the reaction conditions for the aldehyde reduction a.
Molecules 25 00421 i001
EntryHCO2XFeTemperature (°C)Time (h)Conv. (%) b
1HCO2HFe4100240
2HCO2H/Et3N (1/1)Fe410024100
3HCO2NaFe410024100
4HCO2KFe410024100
5HCO2CsFe410024100
6-Fe4100240
7HCO2Na-100240
8HCO2NaFe41001683
9HCO2NaFe48024100 (99%) c
10HCO2NaFe380240
11HCO2NaFe4801681
12HCO2NaFe4602475
13 dHCO2NaFe4802453
14 eHCO2NaFe4802486
15HCO2NaFe58024100 (98%) c
a General conditions: HCO2X (5 mmol, 5 equiv.), 4-methoxybenzaldehyde (1 mmol), pre-catalyst (2 mol %), Me3NO (2.5 mol %), water (2 mL). b Conversion was determined by 1H-NMR spectroscopy analysis. c Isolated yield in the bracket. d Fe 0(1 mol %), Me3NO (1.25 mol %) were used. e HCO2Na (3 mmol, three equiv.) were used.
Table 2. Iron-catalyzed reduction of aldehydes with sodium formate a.
Table 2. Iron-catalyzed reduction of aldehydes with sodium formate a.
Molecules 25 00421 i002
Molecules 25 00421 i003
a General conditions: aldehyde (1 mmol), HCO2Na (5 mmol, 5 equiv.), pre-catalyst Fe4 (2 mol %), Me3NO (2.5 mol %), water (2 mL). b H2O/EtOH 1/1.
Table 3. Optimization of the reaction conditions for the reductive amination a.
Table 3. Optimization of the reaction conditions for the reductive amination a.
Molecules 25 00421 i004
EntryHCO2X (equiv.)[Fe]T (°C)Conv. (%) bSelectivity (5a)/(5a’) b
1HCO2NH4 (5)Fe4909377/23
2HCO2NH4 (5)Fe5909577/23
3HCO2K (5)Fe5909460/40
4HCO2Cs (5)Fe5909340/60
5-Fe5901000/100
6HCO2NH4 (5)Fe5859167/33
7HCO2NH4 (5)Fe5808369/31
8HCO2NH4 (5)Fe5408052.5/47.5
9HCO2NH4 (6.5)Fe5909691/9 (70) c
a General conditions: HCO2X (5 mmol, 5 equiv.), citronellal (1 mmol), N-methylbenzylamine (2 equiv.), pre-catalyst Fe (2 mol %), Me3NO (2.5 mol %), water (2 mL). b Conversion and selectivity were determined by 1H-NMR spectroscopy analysis. c Isolated yield in the bracket.
Table 4. Iron-catalyzed reductive amination with ammonium formate a.
Table 4. Iron-catalyzed reductive amination with ammonium formate a.
Molecules 25 00421 i005
Molecules 25 00421 i006
a General conditions: aldehyde (1 mmol), amine (2 equiv.), HCO2NH4 (6.5 mmol, 6.5 equiv.), pre-catalyst Fe4 or Fe5 (2 mol %), Me3NO (2.5 mol %), water (2 mL), 90 °C for 24 h. b for 48 h.
Table 5. Recycling of the pre-catalyst Fe5 a,b.
Table 5. Recycling of the pre-catalyst Fe5 a,b.
Molecules 25 00421 i007
EntryRunConv. (%) cSelectivity (5a)/(5a’) c
119691/9
229596/4
339596/4
449697/3
559896/4
a General conditions for the initial run: citronellal (1 mmol), N-methylbenzylamine (2 equiv.), HCO2NH4 (6.5 equiv.), pre-catalyst Fe5 (2 mol %), Me3NO (2.5 mol %), water (2 mL), 90 °C for 24 h. b General conditions for run 2–5: citronellal (1 mmol), N-methylbenzylamine (2 equiv.), HCO2NH4 (6.5 equiv.) were added to the former solution, and the mixture was heated to 90 °C. c Conversion and selectivity were determined by 1H-NMR spectroscopy analysis.

Share and Cite

MDPI and ACS Style

Ndiaye, D.; Coufourier, S.; Mbaye, M.D.; Gaillard, S.; Renaud, J.-L. Cyclopentadienone Iron Tricarbonyl Complexes-Catalyzed Hydrogen Transfer in Water. Molecules 2020, 25, 421. https://doi.org/10.3390/molecules25020421

AMA Style

Ndiaye D, Coufourier S, Mbaye MD, Gaillard S, Renaud J-L. Cyclopentadienone Iron Tricarbonyl Complexes-Catalyzed Hydrogen Transfer in Water. Molecules. 2020; 25(2):421. https://doi.org/10.3390/molecules25020421

Chicago/Turabian Style

Ndiaye, Daouda, Sébastien Coufourier, Mbaye Diagne Mbaye, Sylvain Gaillard, and Jean-Luc Renaud. 2020. "Cyclopentadienone Iron Tricarbonyl Complexes-Catalyzed Hydrogen Transfer in Water" Molecules 25, no. 2: 421. https://doi.org/10.3390/molecules25020421

Article Metrics

Back to TopTop