Next Article in Journal
Enhancement of Sono-Fenton by P25-Mediated Visible Light Photocatalysis: Analysis of Synergistic Effect and Influence of Emerging Contaminant Properties
Next Article in Special Issue
Flow-Through Macroporous Polymer Monoliths Containing Artificial Catalytic Centers Mimicking Chymotrypsin Active Site
Previous Article in Journal
Impact of Coordination Features of Co(II)-Glycine Complex on the Surface Sites of Co/SiO2 for Fischer–Tropsch Synthesis
Previous Article in Special Issue
Baeyer-Villiger-Including Domino Two-Step Oxidations of β-O-Substituted Primary Alcohols: Reflection of the Migratory Aptitudes of O-Substituted Alkyl Group in the Outcome of the Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Organocatalyzed Michael Addition to Nitroalkenes via Masked Acetaldehyde

1
Department of Physical and Chemical Sciences, Università degli Studi dell’Aquila, via Vetoio, 67100 L’Aquila, Italy
2
Istituto per la Sintesi Organica e la Fotoreattività, Consiglio Nazionale delle Ricerche, Via P. Gobetti 101, 40129 Bologna, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(11), 1296; https://doi.org/10.3390/catal10111296
Submission received: 9 October 2020 / Revised: 3 November 2020 / Accepted: 4 November 2020 / Published: 9 November 2020
(This article belongs to the Special Issue Organocatalysis: Advances, Opportunity, and Challenges)

Abstract

:
A novel and safe reaction protocol for the enantioselective enamine-catalysed addition of acetaldehyde to nitroalkenes is presented; this protocol makes use of a safe acetaldehyde precursor to access important intermediates to Active Pharmaceutical Ingredients (APIs), and allows the use of fewer equivalents of acetaldehyde and lower catalyst loadings. The reaction developed proved to be suitable to be performed on gram-scale and to produce key intermediates for the synthesis of pharmacologically active compounds such as pregabalin.

Graphical Abstract

1. Introduction

Nowadays, organocatalysis is a key technology platform and is routinely assessed in industry when taking a process to manufacture [1,2,3,4]. In fact, organocatalysis can bring many benefits to an industrial process; the catalysts are in general non-toxic, the reactions are robust, metals are avoided, and strictly controlled conditions are non-necessary. Organocatalysis can be employed to access very valuable γ-amino acids, such as pregabalin, and a great number of organocatalytic synthetic routes to this simple API has been disclosed [5].
Among these, the enamine-catalysed [6,7,8,9] addition of acetaldehyde to a nitroalkene holds high potential for a cost-efficient process; the raw materials are widely available at low prices and the catalyst needed can be accessed at a reasonable cost. Whereas different types of ketones and aldehydes have been activated as nucleophiles, few attempts of using acetaldehyde as nucleophile have been reported with aminocatalysis [10,11]. Controlling acetaldehyde’s reactivity is extremely challenging; in fact, it can easily undergo self-condensation reactions. Furthermore, the product of the enamine-mediated reaction of acetaldehyde is an aldehyde carrying a methylene group, which is still considerably reactive and could potentially react with either a nucleophile or an electrophile to give collateral products. Hayashi [12,13] and List [14] showed independently that, by carefully choosing the reaction conditions, an efficient aminocatalytic enantioselective addition of acetaldehyde to nitroalkenes can be accomplished (Scheme 1a,b).
Despite the great advancement of the reports by Hayashi and List, the approach still suffers from the use of relatively high catalyst loading (10–20 mol%) and the use of a large excess of acetaldehyde; in fact, because of its high tendency to form oligomers, it was used in large excess and added slowly. Furthermore, by using two supported catalysts, Pericàs ingeniously employed paraldehyde to enable the same reaction and avoid the use of acetaldehyde [15] (Scheme 1c). Nevertheless, they used 10 equivalents of masked acetaldehyde (i.e., 3.3 eq. of paraldehyde) and a relatively high catalyst loading of supported organocatalysts that, despite the potential recyclability, bring a considerable cost contribution to the manufacture process.
Herein, we report the use of acetaldehyde dimethyl acetal in the aminocatalytic enantioselective addition to nitroalkenes. By employing a simple masked acetaldehyde, we could tackle the challenges that acetaldehyde brings to an industrial process, lower the catalyst loading, use fewer equivalents of acetaldehyde, and use affordable raw materials and catalyst. The desired γ-nitroaldehydes derivatives were obtained in high yields and enantioselectivities using a very simple, safe, and cost-efficient protocol.

2. Results and Discussion

A range of organic, inorganic, and immobilized acids were tested on the in-situ deprotection of acetaldehyde dimethyl acetal 5 (Table 1). While organic acids afforded negligible amounts of deprotected products (Table 1, entries 1–4), encouraging results were obtained with inorganic acids and acidic resins (Table 1, entries 5–6 and 7–8 and 10) and trifluoroacetic acid (Table 1, entry 9). The most promising acids proved to be TFA and Amberlyst-15, providing the deprotected acetaldehyde 2 in 16% and 18% conversion, respectively (Table 1, entry 9–10). Nafion NRE 212, albeit showing a similar conversion, was discarded for further screening given its higher cost and the challenges that a polymeric sheet brings in a process.
Encouraged by these results, we further optimized the reaction conditions between nitrostyrene 1a and acetaldehyde dimethyl acetal 5 in the presence of a catalytic amount of Hayashi/Jørgensen catalyst 6, by employing Amberlyst-15 as a catalyst to effect the deprotection (Table 2, see also Supplementary Materials); the use of TFA was discarded as no 7 is formed in the presence of TFA, 1a, 6 and 5 in CHCl3. Adventitious water is not enough to afford a high conversion and the addition of water is needed (Table 2, entries 1–2). A solvent screening was carried out with 5 eq. of 5, 10 eq. of water (i.e., 2 eq. with respect to 5) over 72 h (Table 2, entries 2–9). Chloroform afforded the desired product 3a in 94% conversion and 93% ee (Table 2, entry 2). Solvents such as ethyl acetate, acetonitrile, and acetone showed lower conversion and enantioselectivity (Table 2, entries 3–5), while toluene, diethyl ether, and dichloromethane provided lower conversion with enantioselectivity comparable to CHCl3 (Table 2, entries 6–8). The use of dioxane provided similar conversion and ee to CHCl3 over 72 h (Table 2, entry 9). However, when the reaction time was shortened to 24 h, CHCl3 proved to be a better solvent than dioxane both with 10 eq. (Table 2, entries 10–11) and 15 eq. of H2O (Table 2, entries 12–13).
Lowering the equivalents of 5 afforded a more sluggish reaction (Table 2, entries 14–16); the reactivity at 3 eq. could be restored by increasing the concentration to 0.8 M (Table 2, entry 17), while the use of 2 eq. called for 1.6 M conditions (Table 2, entries 18–19).
Using different amounts of Amberlyst-15 proved detrimental to the reaction (Table 2, entries 20–21), slowing it down dramatically in case of a larger amount, probably because the acetaldehyde was released too quickly, giving rise to oligomers (Table 2, entry 21).
Variation of the catalyst (Table 2, entries 22–23) and testing dioxane, one of the promising solvents, in the conditions of entry 19 (Table 2, entry 24), did not bring any improvement.
Based on the results, we chose the conditions in Table 2, entry 19, to evaluate the generality of the reaction (Table 3, see also Supplementary Materials). As expected, and in agreement with the previous reports [13,14,15], other nitroalkenes are less reactive than nitrostyrene. Nevertheless, nitrostyrene derivatives, having either electron-rich or electro-deficient substituents, successfully afforded the desired Michael adducts in high yields and enantioselectivity (Table 3, entries 1–4). The protocol developed proved to be successful also with alkyl-substituted nitroalkenes to give the desired products in high yields and optical purity (Table 3, entries 5–8).
As anticipated at the outset, the catalytic Michael reaction affords nitroaldehydes that are versatile and key intermediates to access important APIs, such as pregabalin [16], a widely used anti-epileptic drug. The developed protocol was tested on >1 g scale of the starting nitroalkene 1h to provide a more reliable expectation of how it should behave on bigger scales (Scheme 2, see also Supplementary Materials). The reaction was stopped after 72 h; the yield proved to be in line with the smaller scale, while the enantioselectivity was higher than on a smaller scale and nearly complete (>99%), pointing to the fact that the protocol is very promising for further development.

3. Conclusions

In conclusion, we have developed an industrially interesting protocol for the Michael addition of acetaldehyde to nitroalkenes, affording the corresponding products in high yields and ee. A current limitation of the presented reaction is the use of a class 2 solvent; however, we believe that further R&D can tackle this issue. The presented reaction makes use of a masked acetaldehyde to avoid the use of a highly toxic and reactive intermediate. Furthermore, the use of an acidic resin and low amounts of an affordable organocatalyst make the overall protocol appealing for more in-depth studies to assess its application in manufacture.

4. Materials and Methods

Typical procedure: acetaldehyde dimethyl acetal 5 (148 µL, 1.4 mmol) was added to a mixture of (S)-diphenyltrimethylsiloxymethyl pyrrolidine 6 (11.4 mg, 0.035 mmol), trans-β-nitrostyrene 1a (105 mg, 0.7 mmol), Amberlyst-15 (10 mol%, H+ exchange capacity (4.7 meq/g)), water (76 µL, 4.2 mmol) and CHCl3 (0.44 mL). The reaction mixture was stirred at room temperature for 24 h and then quenched with 1 mL 1M HCl. Then, the aqueous mixture was extracted with ethyl acetate (3 × 3 mL). The combined organic layers were dried over anhydrous MgSO4, filtered and concentrated in vacuo. Purification by flash chromatography on silica gel (petroleum ether/ethyl acetate = 90:10) gave (S)-4-nitro-3-phenylbutanal 3a (122 mg, 0.63 mmol) in 90% yield and 95% ee. The enantiomeric excess was determined by chiral HPLC analysis after conversion of the aldehyde into the corresponding alcohol by reduction with NaBH4.

Supplementary Materials

General procedures, preparation of starting materials, characterization of compounds, 1H and 13C NMRs, and HPLC traces are available online at https://www.mdpi.com/2073-4344/10/11/1296/s1.

Author Contributions

Methodology, investigation, analysis, data curation, writing—original draft preparation: G.G. and V.N. Writing—review and editing, supervision and funding acquisition: L.P., A.B. and A.C. Conceptualisation and project administration: A.C. All authors have read and agreed to the published version of the manuscript.

Funding

G.G. is grateful to PON—DOT13OV2OC for an industrial PhD fellowship. A.B. acknowledges PON—AIM1842894, CUP E18D19000560001 for support. A.C. and A.B. acknowledge Royal Society of ChemistryResearchFund (RF19-5583 and R19-3106) for support.

Acknowledgments

A.C. is extremely grateful to Fred Cederbaum and Nikola Kokosar, Syngenta Crop Protection Research Center—Stein (CH), Angelini Fine Chemicals Process Development—Aprilia (I), Reddy’s—Cambridge (UK), and Discovery Department, Dompè farmaceutici SpA—L’Aquila (I), for their generous support. Francesco Fini (Università di Modena e Reggio Emilia) is acknowledged for HRMS analysis. Giorgio Bencivenni (Alma Mater Studiorum—Università di Bologna) is acknowledged for help with chiral HPLC analysis of 3e.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Carlone, A.; Bernardi, L. Enantioselective organocatalytic approaches to active pharmaceutical—Selected industrial examples. Phys. Sci. Rev. 2019, 4, 20180097. [Google Scholar] [CrossRef]
  2. Blaser, H.U.; Federsel, H.J. Asymmetric Catalysis on Industrial Scale: Challenges, Approaches and Solutions, 2nd ed.; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010; ISBN 9783527324897. [Google Scholar]
  3. Wallace, D.; Challenger, S.; Ding, Z.D.; Osminski, W.E.G.; Ren, H.; Wulff, W.D.; Desai, A.A.; Munmun, M.; Scott, J.P.; Alam, M.; et al. Special feature section: Asymmetric synthesis on large scale. Org. Process. Res. Dev. 2011, 15, 1088–1211. Available online: https://pubs.acs.org/toc/oprdfk/15/5 (accessed on 9 November 2020). [CrossRef]
  4. Dunn, P.J.; Hii, K.K.; Krische, M.J.; Williams, M.T. Sustainable Catalysis: Challenges and Practices for the Pharmaceutical and Fine Chemical Industries; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013; ISBN 9781118155424. [Google Scholar]
  5. Adamo, M.F.A.; Kelly, B.G.; Moccia, M. Recent advances in the preparation of active pharmaceutical ingredient (S)-Pregabalin. Chim. Oggi Chem. Today 2016, 34, 54–57. [Google Scholar]
  6. Dalko, P.I. Comprehensive Enantioselective Organocatalysis: Catalysts, Reactions, and Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; Volume 1–3, ISBN 9783527658862. [Google Scholar]
  7. List, B. Asymmetric Organocatalysis in Topics in Current Chemistry; Springer: Berlin, Germany; New York, NY, USA, 2009; ISBN 978-3-642-02815-1. [Google Scholar]
  8. Torres, R.R. Stereoselective Organocatalysis: Bond Formation Methodologies and Activation Modes; Wiley: Hoboken, NJ, USA, 2013; ISBN 978-1-118-60470-0. [Google Scholar]
  9. Melchiorre, P.; Marigo, M.; Carlone, A.; Bartoli, G. Asymmetric aminocatalysis—Gold rush in organic chemistry. Angew. Chem. Int. Ed. 2008, 47, 6138–6171. [Google Scholar] [CrossRef] [PubMed]
  10. Córdova, A.; Notz, W.; Barbas, C.F. Proline-catalyzed one-step asymmetric synthesis of 5-hydroxy-(2E)-hexenal from acetaldehyde. J. Org. Chem. 2002, 67, 301–303. [Google Scholar] [CrossRef] [PubMed]
  11. Bøgevig, A.; Kumaragurubaran, N.; Jørgensen, K.A. Direct catalytic asymmetric aldol reactions of aldehydes. Chem. Commun. 2002, 2, 620–621. [Google Scholar] [CrossRef] [PubMed]
  12. Hayashi, Y.; Itoh, T.; Aratake, S.; Ishikawa, H. A diarylprolinol in an asymmetric, catalytic, and direct crossed-aldol reaction of acetaldehyde. Angew. Chem. Int. Ed. 2008, 47, 2082–2084. [Google Scholar] [CrossRef]
  13. Hayashi, Y.; Itoh, T.; Ohkubo, M.; Ishikawa, H. Asymmetric Michael reaction of acetaldehyde catalyzed by diphenylprolinol silyl ether. Angew. Chem. Int. Ed. 2008, 47, 4722–4724. [Google Scholar] [CrossRef] [PubMed]
  14. García-García, P.; Ladépêche, A.; Halder, R.; List, B. Catalytic asymmetric michael reactions of acetaldehyde. Angew. Chem. Int. Ed. 2008, 47, 4719–4721. [Google Scholar] [CrossRef] [PubMed]
  15. Fan, X.; Rodríguez-Escrich, C.; Sayalero, S.; Pericàs, M.A. Paraldehyde as an acetaldehyde precursor in asymmetric michael reactions promoted by site-isolated incompatible catalysts. Chem. Eur. J. 2013, 19, 10814–10817. [Google Scholar] [CrossRef] [PubMed]
  16. Gotoh, H.; Ishikawa, H.; Hayashi, Y. Diphenylprolinol silyl ether as catalyst of an asymmetric, catalytic, and direct michael reaction of nitroalkanes with α,β-unsaturated aldehydes. Org. Lett. 2007, 9, 5307–5309. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Literature-reported examples of enantioselective Michael addition of acetaldehyde to nitroalkenes compared to our synthetic strategy.
Scheme 1. Literature-reported examples of enantioselective Michael addition of acetaldehyde to nitroalkenes compared to our synthetic strategy.
Catalysts 10 01296 sch001
Scheme 2. Application of the developed protocol on >1 g scale to afford synthetically useful intermediate 3h.
Scheme 2. Application of the developed protocol on >1 g scale to afford synthetically useful intermediate 3h.
Catalysts 10 01296 sch002
Table 1. Results of the acid-catalysed deprotection of acetaldehyde dimethyl acetal 1.
Table 1. Results of the acid-catalysed deprotection of acetaldehyde dimethyl acetal 1.
Catalysts 10 01296 i001
EntryAcidConv. (%) 2
1Benzoic Acid0
2AcOH1
3p-NO2-C6H4CO2H0
4pTSA0
5HCl14
6H2SO414
7Amberlyst-3612
8Nafion NRE 21215
9TFA16
10Amberlyst-1518
1 Acetaldehyde dimethyl acetal (0.8 mmol) and acid (10 mol%) were mixed in CDCl3 for 2 h. 2 Conversion of acetaldehyde dimethyl acetal 5 into acetaldehyde 2.
Table 2. Optimization of conditions for the Michael addition 1.
Table 2. Optimization of conditions for the Michael addition 1.
Catalysts 10 01296 i002
Entry5 (eq)H2O (eq)Solventt (h)Conv. [%] 2ee (%) 3
150CHCl3724696
2510CHCl3729493
3510AcOEt723288
4510MeCN72544
5510Acetone727587
6510Toluene728594
7510Et2O727194
8510CH2Cl2726192
9510Dioxane729394
10510Dioxane247193
11510CHCl3249393
12515Dioxane248495
13515CHCl3249793
141.23.6CHCl3245791
1526CHCl3247490
1639CHCl3248192
17 439CHCl324>9994
18 426CHCl3248192
19 526CHCl3249490
20 5,626CHCl3249190
21 5,726CHCl3241093
22 5,826CHCl3240n.d
23 926CHCl3244891
24 526Dioxane246592
1 Reactions performed with catalyst 6 (7.5 mg, 0.02 mmol, 0.05 eq), trans-β-nitrostyrene 1a (60 mg, 0.4 mmol, 1 eq.), acetaldehyde dimethyl acetal 5, water, Amberlyst-15 (10 mol%) and solvent (1 mL, 0.4 M) at room temperature. 2 Measured by 1H NMR spectroscopy. 3 Determined by chiral HPLC analysis after conversion of the aldehyde into the corresponding alcohol by reduction with NaBH4. See the Supporting Information for details. 4 0.5 mL of solvents were used, M = 0.8 mol/L. 5 0.25 mL of solvents were used, M = 1.6 mol/L. 6 7 mg of Amberlyst-15 were used. 7 28 mg of Amberlyst-15 were used. 8 L-proline was used as a catalyst. 9 (S)-α,α-Bis [3,5-bis(trifluoromethyl)phenyl]-2-pyrrolidinemethanol trimethylsilyl ether (Jørgensen catalyst) was used as a catalyst.
Table 3. Catalytic Michael addition of acetaldehyde dimethyl acetal 5 with various nitroalkenes 1.
Table 3. Catalytic Michael addition of acetaldehyde dimethyl acetal 5 with various nitroalkenes 1.
Catalysts 10 01296 i003
EntryR3t (h)Yield (Conv.) (%) 2ee (%) 3
1 Catalysts 10 01296 i0043a2490 (94)95
2 Catalysts 10 01296 i0053b7262 (65)94
3 Catalysts 10 01296 i0063c7276 (80)92
4 Catalysts 10 01296 i0073d4883 (85)93
5 Catalysts 10 01296 i0083e4875 (78)96
6 4 Catalysts 10 01296 i0093f4889 (92)−95 5
7 4 Catalysts 10 01296 i0103g7293 (94)93
8 Catalysts 10 01296 i0113h7283 (86)94 6
1 The reaction was performed with nitroalkenes (0.7 mmol, 1 eq), acetaldehyde dimethyl acetal 5 (1.4 mmol, 2 eq), catalyst 6 (0.035 mmol, 0.05 eq), Amberlyst-15 (10 mol%), water (4.2 mmol, 6 eq) in CHCl3 (0.44 mL, 1.6 M) at room temperature. 2 Yield of isolated product. 3 Optical purity was determined by chiral HPLC/GC analysis after conversion of the aldehyde into the corresponding alcohol by reduction with NaBH4. 4 Amberlyst-15 (5 mol%) was used. 5 (R)-6 was used as a catalyst. 6 Optical purity was determined after conversion of the corresponding alcohol into the tosylate derivative.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Giorgianni, G.; Nori, V.; Baschieri, A.; Palombi, L.; Carlone, A. Organocatalyzed Michael Addition to Nitroalkenes via Masked Acetaldehyde. Catalysts 2020, 10, 1296. https://doi.org/10.3390/catal10111296

AMA Style

Giorgianni G, Nori V, Baschieri A, Palombi L, Carlone A. Organocatalyzed Michael Addition to Nitroalkenes via Masked Acetaldehyde. Catalysts. 2020; 10(11):1296. https://doi.org/10.3390/catal10111296

Chicago/Turabian Style

Giorgianni, Giuliana, Valeria Nori, Andrea Baschieri, Laura Palombi, and Armando Carlone. 2020. "Organocatalyzed Michael Addition to Nitroalkenes via Masked Acetaldehyde" Catalysts 10, no. 11: 1296. https://doi.org/10.3390/catal10111296

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop