Synlett 2020; 31(06): 581-586
DOI: 10.1055/s-0037-1610751
cluster
© Georg Thieme Verlag Stuttgart · New York

FAST Hydrogenations as a Continuous Platform for Green Aromatic Nitroreductions

MD Taifur Rahman
a   Almac Sciences Ltd., 20 Seagoe Industrial Estate, Craigavon, BT63 5QD, UK   Email: megan.smyth@almacgroup.com
b   Theoretical and Applied Catalysis Research Cluster, School of Chemistry and Chemical Engineering, Queen’s University Belfast, David Keir Building, Stranmillis Road, Belfast, BT9 5AG, UK   Email: h.manyar@qub.ac.uk
,
Scott Wharry
a   Almac Sciences Ltd., 20 Seagoe Industrial Estate, Craigavon, BT63 5QD, UK   Email: megan.smyth@almacgroup.com
,
Megan Smyth
a   Almac Sciences Ltd., 20 Seagoe Industrial Estate, Craigavon, BT63 5QD, UK   Email: megan.smyth@almacgroup.com
,
Haresh Manyar
b   Theoretical and Applied Catalysis Research Cluster, School of Chemistry and Chemical Engineering, Queen’s University Belfast, David Keir Building, Stranmillis Road, Belfast, BT9 5AG, UK   Email: h.manyar@qub.ac.uk
,
Thomas S. Moody
a   Almac Sciences Ltd., 20 Seagoe Industrial Estate, Craigavon, BT63 5QD, UK   Email: megan.smyth@almacgroup.com
c   Arran Chemical Company, Unit 1 Monksland Industrial Estate, Athlone, Co. Roscommon, Ireland
› Author Affiliations
The authors gratefully acknowledge the financial support from Innovate UK under the Knowledge Transfer Partnership (KTP) programme for KTP Associate funding for MDTR for the project ‘Flow-mediated Amine Synthesis Technology Platform’.
Further Information

Publication History

Received: 03 December 2019

Accepted after revision: 27 January 2020

Publication Date:
13 February 2020 (online)


Published as part of the ISySyCat2019 Special Issue

Abstract

A continuous-flow packed-bed catalytic reactor has been developed for hydrogenation of aromatic nitrobenzoic acids in water to produce the corresponding anilines. These hydrogenations are green, more efficient, less consumptive, and safer than the conventional ­reduction process. Various industrially important aromatic amines have been produced in excellent yields and with high throughput. The optimized continuous-flow reduction process produces no detectable genotoxic intermediates, unlike the corresponding batch reduction. The reactor is modular in design and can be scaled up to produce several kilograms of product per day without extensive redesign.

Supporting Information

 
  • References and Notes

    • 1a Orlandi M, Brenna D, Harms R, Jost S, Benaglia M. Org. Process Res. Dev. 2018; 22: 430
    • 1b Song J, Huang Z.-F, Pan K, Li K, Zhang X, Wag L, Zou J.-J. Appl. Catal., B 2018; 227: 386
    • 1c Russell CC, Baker JR, Cossar PJ, McCluskey A. In New Advances in Hydrogenation Processes: Fundamentals and Applications . IntechOpen; London: 2017: 269 ; DOI: 10.5772/65518
    • 2a Cossar PJ, Hizartzidis L, Simone MI, McCluskey A, Gordon CP. Org. Biomol. Chem. 2015; 13: 7119
    • 2b Irfan M, Glasnov TM, Kappe CO. ChemSusChem 2011; 4: 300
    • 2c Yoswathananont N, Nitta K, Nishiuchi Y, Sato M. Chem. Commun. 2005; 40
    • 2d Hutchings M, Wirth T. Synlett 2016; 27: 1832
    • 3a Teasdale A, Elder D, Chang S.-J, Wang S, Thompson R, Benz N, Sanchez Flores IH. Org. Process Res. Dev. 2013; 17: 221
    • 3b Robinson DI. Org. Process Res. Dev. 2010; 14: 946
    • 4a Edgar MJ. US 332417 5, 1967
    • 4b Levy J. US 3882171, 1975
    • 5a Kudaibergenov S, Dauletbekova M, Toleutay G, Kabdrakhmanova S, Seilkhanov T, Abdullin K. J. Inorg. Organomet. Polym. 2018; 28: 2427
    • 5b Winterbottom M, Fishwick R, Stitt H. Can. J. Chem. Eng. 2003; 81: 588
    • 5c Andersson B. AIChE J. 1982; 28: 333
    • 6a Pošta M, Soós V, Beier P. Tetrahedron 2016; 72: 3809
    • 6b Liu H, Wang P, Yang H, Niu J, Ma J. New J. Chem. 2015; 39: 4343
    • 6c Pandarus V, Ciriminna R, Beland F, Pagliaro M. Catal. Sci. Technol. 2011; 1: 1616
    • 7a Khan M, Joshi S, Ranade V. Chem. Eng. J. (Amsterdam, Neth.) 2019; 377: 120512 ; and references cited therein
    • 7b Purwanto Purwanto, Deshpande RM, Chaudhari RV, Delmas H. J. Chem. Eng. Data 1996; 41: 1414
    • 8a Vardanyan RS, Hruby VJ. Synthesis of Essential Drugs . Elsevier; Amsterdam: 2006: 9
    • 8b Osgood PI, Moss SH, Davies DJ. G. J. Invest. Dermatol. 1982; 79: 354
  • 9 Batch Hydrogenation of p-Nitrobenzoic Acid The batch reactions were all performed in a 100 mL stainless-steel stirred-tank reactor (Autoclave Engineers, Erie, PA). In a typical reaction, the reactor was charged with 5 wt% Pd/C catalyst (0.02 g) and an aqueous mixture (30 mL) of p-nitrobenzoic acid (PNBA; 2.5 g, 15 mmol) and NaOH (0.695 g, 17.4 mmol). The concentrations of PNBA and NaOH in the aq solution were 0.50 M and 0.58 M, respectively. The reactor was sealed and purged three times with N2, and then the mixture was heated to 80 °C with stirring at 1250 rpm. The reactor was pressurized to 10 or 5 bar with H2, at which point the reaction started. 200 μL aliquots of the reaction mixtures were withdrawn from the reactor at 2, 4, 6, 8 10, 15, 20, 25, 30, 35, 40, 45, 60, and 90 min. Each sample was diluted with 1.8 mL of D2O containing a known amount of DMSO as an external standard, and filtered through a syringe filter to remove any suspended catalyst. The percentage conversion of PNBA and the percentage yields of p-aminobenzoic acid (PABA) and its various intermediates were quantified by means of 1H NMR analysis. Reference samples for PNBA and PABA were prepared from commercially available HPLC-grade reagent samples. Intermediates 4, 5, and 6, observed in the reaction mixtures, were subsequently isolated and synthesized by separate batch reactions, as described in the Supporting Information. Quantitative analysis of these intermediates in the batch reaction mixtures was performed by comparison with these authentic samples.
    • 10a McManus IJ, Daly H, Manyar HG, Taylor SF. R, Thompson JM, Hardacre C. Faraday Discuss. 2016; 188: 451
    • 10b Figueras F, Coq B. J. Mol. Catal. A: Chem. 2001; 173: 223
    • 10c Haber F. Z. Elektrochem. Angew. Phys. Chem. 1898; 22: 506
  • 11 Hessel V, Angeli P, Gavriilidis A, Löwe H. Ind. Eng. Chem Res. 2005; 44: 9750
    • 12a Ranade VV, Chaudhari RV, Gunjal PR. Trickle Bed Reactors: Reactor Engineering and Applications . Elsevier; Oxford: 2011
    • 12b Duduković MP, Larachi F, Mills PL. Catal. Rev.: Sci. Eng. 2002; 44: 123
  • 13 Flow Hydrogenation of p-Nitrobenzoic Acid The construction of the packed-bed reactor and its associated gas- and liquid-feed lines, gas–liquid separators, etc., are described in the Supporting Information. For a flow experiment, the reactor was primed with a continuous flow of 0.58 M NaOH solution under 10 bar of hydrogen pressure. The gas flow rate was measured by using bubble flowmeter; the average flow rate under atmospheric conditions was 83 mL/min. The gas and liquid flow were kept under 10 bar pressure for 30 min to equilibrate the gas–liquid flow inside the packed bed and to stabilize the gas–liquid flow through the packed bed of 5% Pd/C. The liquid flow was then stopped and the packed bed was sealed by closing the upstream and downstream one-way valves to maintain a 10 bar hydrogen pressure at 80 °C to prereduce the catalyst. After 2 h of prereduction of the catalyst, the valves were opened and hydrogen flow was maintained along with the desired flow rate of the PNBA solution (0.5 M in 0.58 M NaOH aq solution). The initial reaction mixture (three times the internal volume of the packed-bed reactor column) collected at the gas–liquid separator was discarded. The subsequent mixtures, after flow equilibration and attainment of a steady state, were collected for 1H NMR analysis as described for the batch reaction.9
  • 14 Yang C, Teixeira AR, Shi Y, Born SC, Lin H, Song YL, Martin B, Schenkel B, Lachegurabia MP, Jensen KF. Green Chem. 2018; 20: 886