Next Article in Journal
Association of Moderate Beer Consumption with the Gut Microbiota and SCFA of Healthy Adults
Previous Article in Journal
Antibacterial Activity of Honey/Chitosan Nanofibers Loaded with Capsaicin and Gold Nanoparticles for Wound Dressing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Step One-Pot Reductive Amination of Furanic Aldehydes Using CuAlOx Catalyst in a Flow Reactor

by
Alexey L. Nuzhdin
*,
Marina V. Bukhtiyarova
and
Valerii I. Bukhtiyarov
Boreskov Institute of Catalysis SB RAS, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(20), 4771; https://doi.org/10.3390/molecules25204771
Submission received: 26 September 2020 / Revised: 12 October 2020 / Accepted: 15 October 2020 / Published: 17 October 2020

Abstract

:
Aminomethylhydroxymethylfuran derivatives are well known compounds which are used in the pharmaceutical industry. Reductive amination of 5-hydroxymethylfurfural (HMF) derived from available non-edible lignocellulosic biomass is an attractive method for the synthesis of this class of compounds. In the present study, the synthesis of N-substituted 5-(hydroxymethyl)-2-furfuryl amines and 5-(acetoxymethyl)-2-furfuryl amines was performed by two-step process, which includes the condensation of furanic aldehydes (HMF and 5-acetoxymethylfurfural) with primary amines in methanol on the first step and the reduction of obtained imines with hydrogen in a flow reactor over CuAlOx catalyst derived from layered double hydroxide on the second step. This process does not require isolation and purification of intermediate imines and can be used to synthesize a number of aminomethylhydroxymethylfurans in good to excellent yield.

1. Introduction

N-Substituted 5-(hydroxymethyl)-2-furfuryl amines are an important class of compounds due to their usage in pharmaceutical industry for production of calcium antagonists, muscarinic agonists, cholinergic agents, carcinogenesis inhibitors, etc. (for instance, anti-hepatitis B virus drug and antihypertensives, Scheme 1) [1,2,3]. Generally, these structures are prepared by a Mannich-type reaction starting from furfural alcohol (or furfural), formaldehyde, and primary amines. However, harsh reaction conditions are usually required, that leads to a low yield of the target product [4,5,6]. Meanwhile, reductive amination of 5-hydroxymethylfurfural (HMF) derived from available non-edible lignocellulosic biomass is an attractive method for the synthesis of this type of amines [3,7,8,9,10,11].
In general, the reductive amination of HMF is realized in the presence of noble metal-based catalysts such as dichlorobis(2,9-dimethyl-1,10-phenanthroline) ruthenium(II) complex [7], Pd/C [8], Au/TiO2 [3], and Pd nanoparticles immobilized in a MOF/polymer composite [9] using batch reactors. However, the high price and limited availability of precious metals are stimulated interest in catalysts based on earth-abundant elements. Chieffi et al. studied the reductive amination of HMF over carbon-supported Fe-Ni catalyst under flow conditions. However, the amines used were limited to 3-aminopropanol and alanine sodium and the yield of the target aminomethylhydroxymethylfuran (AMHMF) derivatives did not exceed 78% [11].
In our previous study [12], an environmentally friendly procedure has been developed for the synthesis of secondary amines through a two-step one-pot reductive amination of aromatic aldehydes with primary amines. The condensation of benzaldehyde derivatives (or furfural) with amines in methanol at room temperature gives imines, which are then hydrogenated in a flow reactor using Cu-Al mixed oxide (denoted hereafter as CuAlOx) derived from layered double hydroxide (LDH). In present work, we investigated the efficiency of this procedure for reductive amination of furanic aldehydes (HMF and 5-acetoxymethylfurfural) with primary amines via hydrogenation of the intermediate imines over CuAlOx catalyst in a flow reactor.

2. Results and Discussion

Methanol is a suitable solvent for the synthesis of N-substituted 5-(hydroxymethyl)-2-furfuryl amines, since it is a good solvent for both the condensation of aromatic aldehydes with primary amines and hydrogenation of imines [10,12,13]. We found that the reaction of HMF with aniline in low water methanol (0.02 wt% H2O) occurs at room temperature with imine yield of 98% (Table 1, entry 1) and time dependence of imine yield let us to see that equilibrium concentrations are reached within 2 h (Supplementary Materials, Figure S1). The use of ethanol instead of methanol leads to a decrease in the yield of imine 1a to 70% (Table 1, entry 2). Simultaneously, the yield of 1a in isopropanol is only 31% under the same reaction conditions (Table 1, entry 3).
The reaction mixtures obtained by condensation of HMF with primary amines were further hydrogenated over CuAlOx catalyst in a flow reactor at 100 °C and H2 pressure of 10 bar. Before each catalytic run, the CuO phase of CuAlOx was reduced in situ by hydrogen to metallic copper particles [12,14]. The average size of copper crystallites in the reduced sample was about 7 nm [14] that is much smaller than in case of the supported Cu/Al2O3 catalysts with high copper content. Thus, CuAlOx material prepared by calcination of LDH precursor is a promising non-noble metal catalyst for hydrogenation reactions [12,14,15].
The hydrogenation of imine 1a, which is obtained by condensation of HMF and aniline in methanol, leads to the formation of AMHMF derivative 2a with the yield of 97% (Table 2, entry 1). In addition to 2a, undesired bis-(hydroxymethyl)furan 3 was observed among the reaction products. The use of ethanol or isopropanol instead of methanol leads to a strong drop in the yield of the target product (Table 2, entries 2 and 3) due to the much lower yield of imine in the first stage (Table 1). As a result, a greater amount of HMF is hydrogenated to 3. Raising the concentration of reactants in methanol to 0.2 M reduces the yield to 94% (Table 2, entry 4). However, an increase of reaction temperature to 120 °C at high reactant concentration leads to rise of the imine hydrogenation rate and allows synthesizing 2a with the yield of 98% (Table 2, entry 5).
The effect of primary amine structure on the formation of AMHMF derivatives was investigated by reductive amination of HMF with different aromatic and aliphatic amines over CuAlOx catalyst (Table 2, entries 1 and 6–16). Formation of 2b, 2c, and 2e with the excellent yield (95–97%) is observed for aromatic amines containing electron-donating substituent (CH3 and OCH3) in meta- and para-position (Table 2, entries 6, 7, and 9). It means that the rates of imine formation in the condensation of HMF with these compounds are similar to that in the reaction of HMF with aniline while using o-toluidine gives lower rate of imine formation due to steric hindrance created by methyl group in ortho-position [12]. A much longer period of time (16 h) is needed for the condensation of HMF with o-toluidine at the first stage to obtain 2d with the yield of 88% by imine hydrogenation over the CuAlOx catalyst at the second stage (Table 2, entry 8).
Introduction of F and Cl substituent in the para-position of aniline slightly decreases the yield of compounds 2 to 93–94% (Table 2, entries 10 and 11). At the same time, in the case of p-bromoaniline the significant drop in the yield of the target product to 60% was observed while imine amount in the final mixture was increased (Table 2, entry 12). It is worth to mention that reaction was performed at 110 °C that is higher than in case of other aromatic amines. This result can be explained by the steric hindrance created by bulky Br substituent during the diffusion of imine to active sites of CuAlOx [12]. It should be noted that hydrodehalogenated product practically did not form during the reaction.
The position of electron-withdrawing substituent in aromatic amine has a strong effect on the reaction. The imine formation rate in the condensation of HMF with m-chloroaniline is much lower than with p-chloroaniline. As a result, the reaction of HMF with m-chloroaniline for 3 h on the first stage leads to formation of 2i in a low yield (Table 2, entry 13). However, carrying out the condensation of HMF and m-chloroaniline for 16 h allows obtaining 2i in 95% yield (Table 2, entry 14). At the same time, o-chloroaniline gave a moderate yield of 2j (52%) after the reaction with HMF on the first stage for 16 h (Table 2, entry 15). These observations are explained by a decrease in the nucleophilic properties of chloroanilines in the following order: p-chloroaniline > m-chloroaniline > o-chloroaniline.
Thus, the yield of 2 is determined by the imine formation at the first stage. Higher yield of the intermediate imine provides higher yield of AMHMF derivative. In the case of primary amines with weak nucleophilic properties (o-toluidine, m-chloroaniline, o-chloroaniline), a longer condensation time (16 h) is required to obtain compound 2 with a higher yield.
Using aliphatic hexylamine gave a slightly lower yield of AMHMF derivative than aniline (Table 2, entries 1 and 16) that can be explained by instability of intermediate imines containing alkyl substituent [16].
The time-on-stream measurements for reductive amination of HMF with aniline over the CuAlOx catalyst were performed for~3 h (Supplementary Materials, Figure S3). The yield of 2a (97%) remained unchanged meaning that the catalyst exhibits good stability during the reaction [12,14].
5-Acetoxymethylfurfural (AMF), which can be prepared from carbohydrates without isolating HMF, is less reactive, more stable and easily extracted from the aqueous reaction mixture. Therefore, AMF represents a good substitute for HMF [17,18,19]. Similar to HMF, AMF reacts with aniline in methanol to form imine with the yield of 98% (Supplementary Materials, Figure S2). The hydrogenation of imines obtained by condensation of AMF with different amines was carried out on the CuAlOx catalyst at lower temperature of 80 °C than in case of HMF in order to avoid the reduction of the acetoxymethyl group. It was found that reductive amination of AMF with aniline and its derivatives containing CH3 group in para- and meta- positions leads to the formation of N-substituted 5-(acetoxymethyl)-2-furfuryl amines 6a–c with the yields about 99% (Table 3, entries 1–3). Introduction of methyl substituent in ortho-position of aniline decreases the yield of compound 6 to 96% (Table 3, entry 4). It is worth to mention that condensation of AMF with o-toluidine was performed for 16 h while temperature of following hydrogenation of obtained imine was raised to 90 °C. In the case of anilines with CH3O, F, and Cl substituents in para-position, AMHMF derivatives 6e, 6f, and 6g were obtained with the yield of 99, >98 and 97%, respectively (Table 3, entries 5–7). Simultaneously, the reaction of AMF with m-chloroaniline for 16 h at the first stage, followed by imine hydrogenation at 80 °C, allows synthesizing 6h with the yield of 96% (Table 3, entry 8). In contrast to aromatic amines, hexylamine gives only 82% yield of the AMHMF derivative due to hydrogenation of acetyl group in 6i (Table 3, entry 9).

3. Materials and Methods

3.1. Chemicals

Aniline (99.8%), o-toluidine (99%), m-toluidine (99%), p-toluidine (99%), p-anisidine (99%), o-chloroaniline (>98%), m-chloroaniline (99%), p-chloroaniline (98%), p-fluoroaniline (98%), p-bromoaniline (>99%), n-hexylamine (99%), and 5-(acetoxymethyl)-2-furaldehyde (97%) from Acros Organics (Geel, Belgium), as well as 5-hydroxymethylfurfural (99%) from Sigma-Aldrich (St Louis, MO, USA), were used without additional purification. Methanol (99.8%) from J.T. Baker (Phillipsburg, NJ, USA), ethanol (99.9%) from J.T. Baker and isopropanol (99.5%) from Acros Organics were employed as a solvent.

3.2. Catalyst Preparation

Cu-Al layered double hydroxide was prepared by co-precipitation method at pH 9.0 and temperature of 70 °C using the mixture of copper (II) nitrate and aluminum nitrate as starting material and solution containing NaOH and Na2CO3 ([CO32]/[Al3+] = 0.86, [OH] = 1.6([Al3+] + [Cu2+]) as precipitant agent. The aging of the obtained suspension was performed at 70 °C for 4 h. Afterwards, the precipitate was filtered, washed with hot water and dried at 110 °C for 14 h. The as-prepared LDH was calcined at 650 °C for 4 h to obtain desired CuAlOx catalyst [11,13] containing 47.2 wt% of Cu and 19.9 wt% of Al. The texture properties of CuAlOx were found to be as follows: BET specific surface area—68 m2 g−1 and total pore volume—0.15 cm3 g−1.

3.3. Characterization Techniques

The Cu and Al content was measured by atomic absorption spectroscopy on an Optima 4300 DV instrument (Perkin Elmer, Waltham, MA, USA). Textural characteristics were determined from nitrogen adsorption–desorption isotherms (77 K) obtained on a Micromeritics ASAP 2400 analyzer (Micromeritics, Norcross, GA, USA).

3.4. General Procedure for Reductive Amination of Furanic Aldehydes with Primary Amines

Depending on the nucleophilic properties of the amine, the solution of furanic aldehyde (0.05 M) and primary amine (0.05 M) in methanol was kept at 25 °C for 3 or 16 h. Then, the reaction mixture was mixed with H2 and pumped through the flow reactor packed with the CuAlOx material. Catalytic experiments were performed in H-Cube Pro setup (Thalesnano, Budapest, Hungary) equipped with a 30 mm CatCart cartridge (CatCart®30, 0.30 mL empty volume) [11,13]. Before the catalytic run, the catalyst (0.165 g of 250–500 μm particles) was reduced by a mixture of hydrogen with methanol at 120 °C for 1 h (pressure of 10 bar, flow rates of methanol and H2 were 0.5 and 30 mL min−1, respectively). Afterwards, the reaction mixture was pumped through the reactor instead of pure solvent, and this point in time was chosen as the starting point of the experiment.
The reaction was carried out at temperature of 80–120 °C, H2 pressure of 10 bar, liquid feed rate of 0.5 mL min−1 and hydrogen flow rate of 30 mL min−1 (inlet H2/substrate molar ratio was 54). The use of a large excess of H2 in the reaction medium is necessary to avoid the influence of external mass transfer on the reaction progress [13]. The performance of the catalyst was evaluated by analysis of the samples taken in the interval of 30–33 min from the beginning of the experiment. The composition of the reaction products was determined using 1H NMR spectroscopy in CDCl3. The error in determining the yield is ±1%.

4. Conclusions

The two-step one-pot reductive amination of furanic aldehydes, including non-catalytic condensation of HMF (or AMF) with primary amines and the subsequent flow hydrogenation of the obtained imine over Cu-based catalyst derived from layered double hydroxide, was successfully performed using methanol as a solvent. This inexpensive and environmentally friendly process can be utilized to obtain a wide range of N-substituted 5-(hydroxymethyl)-2-furfuryl amines and 5-(acetoxymethyl)-2-furfuryl amines in good to excellent yields (up to 98 and 99%, respectively). The proposed approach is a new method for the synthesis of aminomethylhydroxymethylfuran derivatives which can be used as pharmaceuticals or their precursors.

Supplementary Materials

The following are available online, Figure S1: The time dependence of imine yield in the condensation of HMF (0.05 M) and aniline (0.05 M) at room temperature; Figure S2: The time dependence of imine yield in the reaction between AMF (0.05 M) and aniline (0.05 M) in methanol at room temperature; Figure S3: Time dependence of 2a yield in the two-step one-pot reductive amination of HMF (0.05 M) with aniline (0.05 M) at 100 °C; 1H-NMR spectra of the final reaction mixtures.

Author Contributions

A.L.N. planned and designed the experiments, performed the catalytic activity tests, and wrote the manuscript. M.V.B. performed the catalyst synthesis and revised the manuscript. V.I.B. supervised the project. All authors discussed the results and approved the final version of the manuscript.

Funding

The work was supported by the Russian Science Foundation (grant no. 20-43-05002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Feriani, A.; Gaviraghi, G.; Toson, G.; Mor, M.; Barbieri, A.; Grana, E.; Boselli, C.; Guarneri, M.; Simoni, D.; Manfredini, S. Cholinergic Agents Structurally Related to Furtrethonium. 2. Synthesis and Antimuscarinic Activity of a Series of N-[5-[(1′-Substituted-acetoxy)methyl]-2-furfuryl]dialkylamines. J. Med. Chem. 1994, 37, 4278–4287. [Google Scholar] [CrossRef] [PubMed]
  2. Caetano, J.A.T.; Fernandes, A.C.; Catano, J.A.T. One-pot synthesis of amines from biomass resources catalyzed by HReO4. Green Chem. 2018, 20, 2494–2498. [Google Scholar] [CrossRef]
  3. Zhu, M.-M.; Tao, L.; Zhang, Q.; Dong, J.; Liu, Y.-M.; He, H.-Y.; Cao, Y. Versatile CO-assisted direct reductive amination of 5-hydroxymethylfurfural catalyzed by a supported gold catalyst. Green Chem. 2017, 19, 3880–3887. [Google Scholar] [CrossRef]
  4. Li, X.; Xu, R.; Yang, J.; Nie, S.; Liu, D.; Liu, Y.; Si, C. Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulosic biomass and catalytic upgradation. Ind. Crop. Prod. 2019, 130, 184–197. [Google Scholar] [CrossRef]
  5. Hu, L.; Lin, L.; Wu, Z.; Zhou, S.; Liu, S. Recent advances in catalytic transformation of biomass-derived 5-hydroxymethylfurfural into the innovative fuels and chemicals. Renew. Sustain. Energy Rev. 2017, 74, 230–257. [Google Scholar] [CrossRef]
  6. Froidevaux, V.; Negrell, C.; Caillol, S.; Pascault, J.-P.; Boutevin, B. Biobased Amines: From Synthesis to Polymers; Present and Future. Chem. Rev. 2016, 116, 14181–14224. [Google Scholar] [CrossRef] [PubMed]
  7. Xu, Z.; Yan, P.; Xu, W.; Jia, S.; Xia, Z.; Chung, B.; Hou, G. Direct reductive amination of 5-hydroxymethylfurfural with primary/secondary amines via Ru-complex catalyzed hydrogenation. RSC Adv. 2014, 4, 59083–59087. [Google Scholar] [CrossRef]
  8. García-Ortiz, A.; Vidal, J.D.; Climent, M.J.; Concepción, P.; Corma, A.; Iborra, S. Chemicals from Biomass: Selective Synthesis of N-Substituted Furfuryl Amines by the One-Pot Direct Reductive Amination of Furanic Aldehydes. ACS Sustain. Chem. Eng. 2019, 7, 6243–6250. [Google Scholar] [CrossRef]
  9. Karve, V.V.; Sun, D.T.; Trukhina, O.; Yang, S.; Oveisi, E.; Luterbacher, J.S.; Queen, W.L. Efficient reductive amination of HMF with well dispersed Pd nanoparticles immobilized in a porous MOF/polymer composite. Green Chem. 2020, 22, 368–378. [Google Scholar] [CrossRef]
  10. Cukalovic, A.; Stevens, C.V. Production of biobased HMF derivatives by reductive amination. Green Chem. 2010, 12, 1201–1206. [Google Scholar] [CrossRef]
  11. Chieffi, G.; Braun, M.; Esposito, D. Continuous Reductive Amination of Biomass-Derived Molecules over Carbonized Filter Paper-Supported FeNi Alloy. ChemSusChem 2015, 8, 3590–3594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nuzhdin, A.L.; Bukhtiyarova, M.V.; Bukhtiyarova, G.A. Cu-Al mixed oxide derived from layered double hydroxide as an efficient catalyst for continuous-flow reductive amination of aromatic aldehydes. J. Chem. Technol. Biotechnol. 2020. [Google Scholar] [CrossRef]
  13. Hizartzidis, L.; Cossar, P.J.; Robertson, M.J.; Simone, M.I.; Young, K.A.; McCluskey, A.; Gordon, C.P. Expanding the utility of flow hydrogenation—A robust protocol restricting hydrodehalogenation. RSC Adv. 2014, 4, 56743–56748. [Google Scholar] [CrossRef]
  14. Nuzhdin, A.L.; Bukhtiyarova, M.V.; Bulavchenko, O.A.; Bukhtiyarova, G.A. Flow hydrogenation of 5-acetoxymethylfurfural over Cu-based catalysts. Mol. Catal. 2020, 494, 111132. [Google Scholar] [CrossRef]
  15. Kumalaputri, A.J.; Bottari, G.; Erne, P.M.; Heeres, H.; Barta, K. Tunable and Selective Conversion of 5-HMF to 2,5-Furandimethanol and 2,5-Dimethylfuran over Copper-Doped Porous Metal Oxides. ChemSusChem 2014, 7, 2266–2275. [Google Scholar] [CrossRef] [PubMed]
  16. Layer, R.W. The Chemistry of Imines. Chem. Rev. 1963, 63, 489–510. [Google Scholar] [CrossRef]
  17. Kang, E.-S.; Hong, Y.-W.; Chae, D.W.; Kim, B.; Kim, B.; Kim, Y.J.; Cho, J.K.; Kim, Y.G. From Lignocellulosic Biomass to Furans via 5-Acetoxymethylfurfural as an Alternative to 5-Hydroxymethylfurfural. ChemSusChem 2015, 8, 1179–1188. [Google Scholar] [CrossRef] [PubMed]
  18. Gavilà, L.; Esposito, D. Cellulose acetate as a convenient intermediate for the preparation of 5-acetoxymethylfurfural from biomass. Green Chem. 2017, 19, 2496–2500. [Google Scholar] [CrossRef] [Green Version]
  19. Shinde, S.; Deval, K.; Chikate, R.; Rode, C.V. Cascade Synthesis of 5-(Acetoxymethyl)furfural from Carbohydrates over Sn-Mont Catalyst. ChemistrySelect 2018, 3, 8770–8778. [Google Scholar] [CrossRef]
Sample Availability: Not available.
Scheme 1. Examples of N-substituted 5-(hydroxymethyl)-2-furfuryl amines with pharmaceutical activity.
Scheme 1. Examples of N-substituted 5-(hydroxymethyl)-2-furfuryl amines with pharmaceutical activity.
Molecules 25 04771 sch001
Table 1. Condensation of 5-hydroxymethylfurfural (HMF) with aniline 1.
Table 1. Condensation of 5-hydroxymethylfurfural (HMF) with aniline 1.
Molecules 25 04771 i001
EntrySolventYield, % 2
1Methanol98
2Ethanol70
3Isopropanol31
1 Reaction conditions: HMF (0.05 M), aniline (0.05 M), batch reactor, T = 25 °C, 2 h. 2 GC yields.
Table 2. Two-step one-pot reductive amination of HMF with primary amines 1.
Table 2. Two-step one-pot reductive amination of HMF with primary amines 1.
Molecules 25 04771 i002
EntryRStep 1, hT, °C2Yield, % 2
1234
1Ph31002an.d.973n.d.
2 3Ph31002an.d.7426n.d.
3 4Ph31002an.d.6733n.d.
4 5Ph31002a3943<0.5
5 5Ph31202a198<10.5
6p-CH3C6H431002bn.d.973n.d.
7m-CH3C6H431002cn.d.964n.d.
8o-CH3C6H4161002dn.d.8812<0.5
9p-CH3OC6H431002en.d.955n.d.
10p-FC6H431002fn.d.946n.d.
11p-ClC6H431002gn.d.937n.d.
12p-BrC6H431102h2260161
13m-ClC6H431002in.d.5545n.d.
14m-ClC6H4161002in.d.9541
15o-ClC6H4161002jn.d.52426
16n-hexyl31002kn.d.937n.d.
1 Reaction conditions: HMF (0.05 M), primary amine (0.05 M), CuAlOx catalyst (165 mg), 10 bar H2, liquid flow rate of 0.5 mL min−1, hydrogen flow rate of 30 mL min−1. 2 The product yields were determined by 1H-NMR. 3 Ethanol was used as a solvent. 4 Isopropanol was used as a solvent. 5 Reactant concentrations were 0.2 M.
Table 3. Two-step one-pot reductive amination of 5-Acetoxymethylfurfural (AMF) with primary amines 1.
Table 3. Two-step one-pot reductive amination of 5-Acetoxymethylfurfural (AMF) with primary amines 1.
Molecules 25 04771 i003
EntryRStep 1, h6Yield, % 2
56789
1Ph36an.d.99<0.5<1n.d.
2 p-CH3C6H436bn.d.>99n.d.<1n.d.
3m-CH3C6H436c<0.5>98n.d.1n.d.
4 3o-CH3C6H4166dn.d.9613n.d.
5p-CH3OC6H436en.d.990.5<0.5<0.5
6p-FC6H436f<0.5>98<0.51n.d.
7p-ClC6H436g<0.5970.52<0.5
8m-ClC6H4166hn.d.96<0.522
9n-hexyl36in.d.82 4n.d.n.d.n.d.
1 Reaction conditions: AMF (0.05 M), primary amine (0.05 M), CuAlOx catalyst (165 mg), 80 °C, 10 bar H2, liquid flow rate of 0.5 mL min−1, hydrogen flow rate of 30 mL min−1. 2 The product yields were determined by 1H-NMR. 3 T = 90 °C. 4 Compound 2k was also formed.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nuzhdin, A.L.; Bukhtiyarova, M.V.; Bukhtiyarov, V.I. Two-Step One-Pot Reductive Amination of Furanic Aldehydes Using CuAlOx Catalyst in a Flow Reactor. Molecules 2020, 25, 4771. https://doi.org/10.3390/molecules25204771

AMA Style

Nuzhdin AL, Bukhtiyarova MV, Bukhtiyarov VI. Two-Step One-Pot Reductive Amination of Furanic Aldehydes Using CuAlOx Catalyst in a Flow Reactor. Molecules. 2020; 25(20):4771. https://doi.org/10.3390/molecules25204771

Chicago/Turabian Style

Nuzhdin, Alexey L., Marina V. Bukhtiyarova, and Valerii I. Bukhtiyarov. 2020. "Two-Step One-Pot Reductive Amination of Furanic Aldehydes Using CuAlOx Catalyst in a Flow Reactor" Molecules 25, no. 20: 4771. https://doi.org/10.3390/molecules25204771

Article Metrics

Back to TopTop