direct leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the...

4
Direct Leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the preparation of secondary and tertiary amines Danielle O’Connor, Ashley Lauria, Steven P. Bondi, Shahrokh Saba Department of Chemistry, Fordham University, Bronx, NY 10458, USA article info Article history: Received 12 July 2010 Revised 29 October 2010 Accepted 1 November 2010 Available online 5 November 2010 Keywords: Reductive amination Secondary amines Tertiary amines Aldehydes Ketones abstract A high-yielding and facile one-pot Leuckart-type reaction for rapid access to a number of 2° and 3° amines is described. Ó 2010 Elsevier Ltd. All rights reserved. In modern organic synthesis, the direct reductive amination (DRA) 1 of aldehydes and ketones is a powerful and highly attrac- tive protocol for the synthesis of primary, secondary, and tertiary amines. 2 Its one-step procedure offers operational convenience and avoids preformation or isolation of the imine or iminium inter- mediates prior to their reduction. The selective reduction of the in situ formed C@N bond and the stability of the reducing agent under the reaction conditions, which are often acidic, are critical to the overall success of the process. Catalytic hydrogenation methods, 3 both economical and eco-friendly, are effective for DRAs but are not compatible with substrates with reducible moieties, such as cyano 4 and nitro 5 groups, and may yield alcohols as by-products. Among metal hydride reducing agents, sodium cya- noborohydride (NaBH 3 CN) 6 and sodium triacetoxyborohydride (NaBH(OAc) 3 ), 7 both commercially available, have been widely used. Sodium cyanoborohydride is highly toxic and formation of toxic by-products, such as HCN and NaCN create problematic dis- posal issues and may contaminate the product. 8 Other drawbacks include the need to use a fivefold excess of the amine and slow reaction rates with aromatic ketones and amines of lower basicity. Sodium triacetoxyborohydride, though less toxic, is flammable and water-reactive, and similar limitations are encountered with aro- matic, a,b-unsaturated, and sterically hindered ketones. 7a To over- come such limitations, a variety of other reductant systems, though not as widely used, have been developed. The majority of these reagents employ a catalyst combined with a reducing agent. 9 An alternative DRA approach for the preparation of amines is the Leuckart reaction, 10 whereby an aldehyde or a ketone is heated in the presence of ammonium formate 11 or formamide, 12 or mix- tures of formamide and formic acid. 13,14 The product of the Leuck- art reaction is often the formyl derivative of the desired amine, 11–13 which must be hydrolyzed with acid. This requirement makes the Leuckart reaction unsuitable for substrates with acid-sensitive moieties. Ingersoll et al. 15 improved the original Leuckart proce- dure and expanded the scope of the reaction by developing an ammonium formate-formamide reagent that, when combined with a ketone and heated in the range of 160–185 °C for several hours followed by acid hydrolysis, afforded 1° amines in 60–80% yield. Over the years, different mechanisms have been suggested for the Leuckart reaction; 16 imines, iminium ions, N-formyl imines, or N-formyl iminium cations have been implicated as intermedi- ates in the process. As part of our continued efforts on the use of simple ammonium salts in synthetic transformations, 17 we recently reported a high- yielding one-pot procedure for the preparation of a wide range of iminium salts (1) by direct combination of an aldehyde or a ketone with a 2° amine free base in the presence of ammonium tetrafluo- roborate, ammonium perchlorate, or ammonium hexafluorophos- phate (Scheme 1). 18 Since iminium ions are possible intermediates in the Leuckart reaction, we envisioned that reduction of the in situ formed C@N bond of 1 should occur readily by substituting ammonium for- mate 19 for the ammonium salts previously used for the preparation of iminium salts 1. Therefore, we decided to explore the preparation 0040-4039/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.tetlet.2010.11.006 Corresponding author. Tel.: +1 718 817 4440; fax: +1 718 817 4432. E-mail address: [email protected] (S. Saba). Tetrahedron Letters 52 (2011) 129–132 Contents lists available at ScienceDirect Tetrahedron Letters journal homepage: www.elsevier.com/locate/tetlet

Upload: danielle-oconnor

Post on 02-Jul-2016

253 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Direct Leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the preparation of secondary and tertiary amines

Tetrahedron Letters 52 (2011) 129–132

Contents lists available at ScienceDirect

Tetrahedron Letters

journal homepage: www.elsevier .com/ locate/ tet le t

Direct Leuckart-type reductive amination of aldehydes and ketones: a facileone-pot protocol for the preparation of secondary and tertiary amines

Danielle O’Connor, Ashley Lauria, Steven P. Bondi, Shahrokh Saba ⇑Department of Chemistry, Fordham University, Bronx, NY 10458, USA

a r t i c l e i n f o

Article history:Received 12 July 2010Revised 29 October 2010Accepted 1 November 2010Available online 5 November 2010

Keywords:Reductive aminationSecondary aminesTertiary aminesAldehydesKetones

0040-4039/$ - see front matter � 2010 Elsevier Ltd. Adoi:10.1016/j.tetlet.2010.11.006

⇑ Corresponding author. Tel.: +1 718 817 4440; faxE-mail address: [email protected] (S. Saba).

a b s t r a c t

A high-yielding and facile one-pot Leuckart-type reaction for rapid access to a number of 2� and 3�amines is described.

� 2010 Elsevier Ltd. All rights reserved.

In modern organic synthesis, the direct reductive amination(DRA)1 of aldehydes and ketones is a powerful and highly attrac-tive protocol for the synthesis of primary, secondary, and tertiaryamines.2 Its one-step procedure offers operational convenienceand avoids preformation or isolation of the imine or iminium inter-mediates prior to their reduction. The selective reduction of thein situ formed C@N bond and the stability of the reducing agentunder the reaction conditions, which are often acidic, are criticalto the overall success of the process. Catalytic hydrogenationmethods,3 both economical and eco-friendly, are effective for DRAsbut are not compatible with substrates with reducible moieties,such as cyano4 and nitro5 groups, and may yield alcohols asby-products. Among metal hydride reducing agents, sodium cya-noborohydride (NaBH3CN)6 and sodium triacetoxyborohydride(NaBH(OAc)3),7 both commercially available, have been widelyused. Sodium cyanoborohydride is highly toxic and formation oftoxic by-products, such as HCN and NaCN create problematic dis-posal issues and may contaminate the product.8 Other drawbacksinclude the need to use a fivefold excess of the amine and slowreaction rates with aromatic ketones and amines of lower basicity.Sodium triacetoxyborohydride, though less toxic, is flammable andwater-reactive, and similar limitations are encountered with aro-matic, a,b-unsaturated, and sterically hindered ketones.7a To over-come such limitations, a variety of other reductant systems,though not as widely used, have been developed. The majority ofthese reagents employ a catalyst combined with a reducing agent.9

ll rights reserved.

: +1 718 817 4432.

An alternative DRA approach for the preparation of amines isthe Leuckart reaction,10 whereby an aldehyde or a ketone is heatedin the presence of ammonium formate11 or formamide,12 or mix-tures of formamide and formic acid.13,14 The product of the Leuck-art reaction is often the formyl derivative of the desired amine,11–13

which must be hydrolyzed with acid. This requirement makes theLeuckart reaction unsuitable for substrates with acid-sensitivemoieties. Ingersoll et al.15 improved the original Leuckart proce-dure and expanded the scope of the reaction by developing anammonium formate-formamide reagent that, when combinedwith a ketone and heated in the range of 160–185 �C for severalhours followed by acid hydrolysis, afforded 1� amines in 60–80%yield.

Over the years, different mechanisms have been suggested forthe Leuckart reaction;16 imines, iminium ions, N-formyl imines,or N-formyl iminium cations have been implicated as intermedi-ates in the process.

As part of our continued efforts on the use of simple ammoniumsalts in synthetic transformations,17 we recently reported a high-yielding one-pot procedure for the preparation of a wide range ofiminium salts (1) by direct combination of an aldehyde or a ketonewith a 2� amine free base in the presence of ammonium tetrafluo-roborate, ammonium perchlorate, or ammonium hexafluorophos-phate (Scheme 1).18

Since iminium ions are possible intermediates in the Leuckartreaction, we envisioned that reduction of the in situ formed C@Nbond of 1 should occur readily by substituting ammonium for-mate19 for the ammonium salts previously used for the preparationof iminium salts 1. Therefore, we decided to explore the preparation

Page 2: Direct Leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the preparation of secondary and tertiary amines

R1

C

R2

O

R3

HN

R4

NH4+ NCA

R1

R2

R3

R4

A+ + toluene/benzene NH3 + H2O+ +

A = BF4, ClO4, PF6

heat

1

Scheme 1. One-pot preparation of iminium salts 1.

R1

C

R2

O

R3

HN

R4

NH4+ N

R1

R2

R3

R4

+ + toluene/benzene NH3 + H2O + CO2+heat

432

HCO2 H

Scheme 2. One-pot Leuckart-type preparation of amines 4.

Table 1One-pot preparation of 2� and 3� amines (4)

Entry Aldehyde/ketone Amine Product Yield (%) Time (h)

Benzene Toluene

1O N H

3a

N84 54 1.5

2 O 3a N 89 72 1.5

3O

3aN

80 73 1.5

4O

OO 3a

O

ON 87 75 1.5

5 O 3a N 82 71 1.5

6O

3a N 79 71 1.5

7

O

3aN

66 75 1.5

8

CHO

Me2N2a

3a

N

Me2N 89 70 1.5

9 2a

HN3b

N

Me2N85 75 1.5

10

O

O

CHO

2b

3a

O

O

N

76 73 1.5

11 2b 3a

O

O

N73 72 1.5

130 D. O’Connor et al. / Tetrahedron Letters 52 (2011) 129–132

Page 3: Direct Leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the preparation of secondary and tertiary amines

Table 1 (continued)

Entry Aldehyde/ketone Amine Product Yield (%) Time (h)

Benzene Toluene

12

CHO

2c

3a

N

77 68 1.5

13 2c 3b

N

80 69 1.5

14 O 3b N 67 68 1.5

15O

NH2

3c

HN

85 74 1.5

16

CHO

3c NH

93 71 1.5

17O NH2 H

N 86 60 1.5

D. O’Connor et al. / Tetrahedron Letters 52 (2011) 129–132 131

of amines 4 by a direct, one-step procedure: heating an aldehyde ora ketone (2) with a 1� or 2� amine (3) in the presence of solid ammo-nium formate (Scheme 2). In this Letter, we report the preparationand characterization of various 2� and 3� amines by this Leuckart-DRA procedure.

Typically, amines 420 are prepared by heating a mixture of analdehyde or a ketone, a 1� or 2� amine and solid ammonium for-mate in toluene or benzene in a Dean-Stark apparatus with contin-uous removal of water formed. After the theoretical quantity ofwater was collected, the solution was cooled and the solvent wasremoved by rotary evaporation. The crude products were thenpurified by distillation under reduced pressure. A number of 2�and 3� amines were prepared in high yields by this method in veryshort time (Table 1). While the isolated yield for most of theamines prepared were slightly higher in benzene, good yields ofthe products were obtained in the less toxic and more eco-friendlysolvent toluene. This reaction also worked well with substratescontaining acid-labile acetal groups (Table 1, entries 4 and 10). Fur-thermore, NMR spectra of the crude products did not reveal forma-tion of N-formylamines as by-products, avoiding the need forsubsequent acid hydrolysis. We are currently adapting this DRAprotocol to employ arylamines but we describe herein a samplingof amines derived from iminium ions that we previouslyreported.18

In conclusion, a one-pot, high-yielding, direct reductive amina-tion protocol was developed using commercially available andinexpensive bulk chemicals for rapid access to a variety of 2� and3� amines. A further advantage of this new variant of the Leuckartreaction is the avoidance of subsequent acid hydrolysis of oftenformed N-formyl derivatives of 2� amines and formation of 3�amines as their formates under previously reported Leuckart reac-tion conditions.10

Acknowledgments

The authors gratefully acknowledge financial support fromFordham University and the National Science Foundation’s Divi-sion of Undergraduate Education through Grant DUE #9650684for the NMR spectrometer at Fordham. The authors also thankthe Fordham College Dean’s Office for a Summer Science Internshipaward to S.P.B. Thanks also go to James A. Ciaccio for helpful com-ments on the manuscript.

References and notes

1. In a leading article by Abdel-Magid, et al. (Ref. 7a) reductive amination is‘described as a direct reaction when the carbonyl compound and the amine aremixed with the proper reducing agent without prior formation of theintermediate imine or iminium salt.’

2. Hutchins, R. O.; Hutchins, M. K. In Comprehensive Organic Synthesis; Trost, B. M.,Fleming, I., Eds.; Pergamon Press: New York, 1991; Vol. 8, pp 25–78.

3. (a) Emerson, W. S. In Organic Reactions; Adams, R., Ed.; John Wiley & Sons: NewYork, 1948; Vol. V, pp 174–255; (b) Robichaud, A.; Ajjou, A. N. Tetrahedron Lett.2006, 47, 3633–3636. and references cited therein.

4. Rylander, P. N. In Catalytic Hydrogenation Over Platinum Metals; AcademicPress: New York, 1967; p 128.

5. Roe, A.; Montgomery, J. A. J. Am. Chem. Soc. 1953, 75, 910–912.6. (a) Borch, R. F.; Bernstein, M. D.; Durst, H. D. J. Am. Chem. Soc. 1971, 93, 2897–

2904; (b) Hutchins, R. O.; Natale, N. R. Org. Prep. Proced. Int. 1979, 11, 201–246;(c) Mattson, R. J.; Pham, K. M.; Leuck, D. J.; Cowen, K. A. J. Org. Chem. 1990, 55,2552–2554; For a review on sodium cyanoborohydride, see: (d) Lane, C. F.Synthesis 1975, 135–146.

7. (a) Abdel-Magid, A. F.; Carson, K. G.; Harris, B. D.; Maryanoff, C. A.; Shah, R. D.J.Org. Chem. 1996, 61, 3849–3862; (b) Xiao, S.; Li, Y.; Li, Y.; Liu, H.; Li, H.;Zhuang, J.; Liu, Y.; Lu, F.; Zhang, D.; Zhu, D. Tetrahedron Lett. 2004, 45, 3975–3978; (c) Bailey, H. V.; Heaton, W.; Vicker, N.; Potter, B. V. L. Synlett 2006,2444–2448; (d) Fray, M. J.; Allen, P.; Bradley, P. R.; Challenger, C. E.; Closier, M.;Evans, T. J.; Lewis, M. L.; Mathias, J. P.; Nichols, C. L.; Po-Ba, Y. M.; Snow, H.;Stefaniak, M. H.; Vuong, H. V. Tetrahedron 2006, 62, 6869–6875; (e) Reddy, T. J.;Leclair, M.; Proulx, M. Synlett 2005, 583–586.

Page 4: Direct Leuckart-type reductive amination of aldehydes and ketones: a facile one-pot protocol for the preparation of secondary and tertiary amines

132 D. O’Connor et al. / Tetrahedron Letters 52 (2011) 129–132

8. Moormann, A. E. Synth. Commun. 1993, 23, 789–795.9. For InCl3-Et3SiH, see: Lee, O.-Y.; Law, K.-L.; Yang, D. Org. Lett. 2009, 11, 3302–

3305; For Bu2SnCl2-PhSiH3, see: Apodaca, R.; Xiao, W. Org. Lett. 2001, 3, 1745–1748; For PhMe2SiH-B(Ph)3, see: Blackwell, J. M.; Sonmor, E. R.; Scoccitti, T.;Piers, W. Org. Lett. 2000, 2, 3921–3923; For Et3SiH-[IrCl(cod)]2 or Et3SiH-IrCl3,see: Mizuta, T.; Sakaguchi, S.; Ishii, Y. J. Org. Chem. 2005, 70, 2195–2199; ForEt3SiH-CF3CO2H, see: Chen, B.; Sundeen, J. E.; Guo, P.; Bednarz, M. S.; Zhao, R.Tetrahedron Lett. 2001, 42, 1245–1246; For Cl3SiH-DMF, see: Kobayashi, S.;Yasuda, M.; Hachiya, I. Chem. Lett. 1996, 407–408; Forpolymethylhydrosiloxane-Ti(OiPr)4, see: Chandrasekhar, S.; Reddy, C. R.;Ahmed, M. Synlett 2000, 1655–1657; For Bu3SnH-SiO2, see: Hiroi, R.;Miyoshi, N.; Wada, M. Chem. Lett. 2002, 274–275; For Bu2SnClH-Bu2SnIH,see: (a) Shibata, I.; Suwa, T.; Sugiyama, E.; Baba, A. Synlett 1998, 1081–1082; (b)Suwa, T.; Sugiyama, E.; Shibata, I.; Baba, A. Synthesis 2000, 789–800; (c)Shibata, I.; Moriuchi-Kawakami, T.; Tanizawa, D.; Suwa, T.; Sugiyama, E.;Matsusda, H.; Baba, A. J. Org. Chem. 1998, 63, 383–385; For Cu(PPh3)2BH4-NH2SO3H, see: Bhanushali, M. J.; Nandurkar, N. S.; Bhor, M. D.; Bhalchandra, M.B. Tetrahedron Lett. 2007, 48, 1273–1276; For poly-2,4-ionene borohydride, see:Tajbakhsh, M.; Lakouraj, M. M.; Mahalli, M. S. Synth. Commun. 2008, 38, 1976–1983; For NaBH4-H3PW12O40, see: Heydari, A.; Khaksar, S.; Akbari, J.;Esfandyari, M.; Pourayoubi, M.; Tajbakhsh, M. Tetrahedron Lett. 2007, 68,1135–1138; For Zn(BH4)2-silica gel, see: Ranu, B. C.; Majee, A.; Sarkar, A. J. Org.Chem. 1998, 63, 370–373; For N-methylpiperdine-Zn(BH4)2, see: (a) Alinezhad,H.; Tajbakhsh, M.; Zamani, R. Synlett 2006, 431–434; (b) Alinezhad, H.;Tajbakhsh, M.; Zamani, R. Synth. Commun. 2006, 36, 3609–3615; For pyridineborane, see: (a) Bomann, M. D.; Guch, I. C.; DiMare, M. J. Org. Chem. 1995, 60,5995–5996; (b) Pelter, A.; Rosser, R. M.; Mills, S. J. Chem. Soc., Perkin Trans. 11984, 717–720; For 2-picoline borane, see: Sato, S.; Sakamoto, T.; Miyazawa, E.;Kikugawa, Y. Tetrahedron 2004, 60, 7899–7906; For 5-ethyl-2-methylpyridineborane, see: Burkhardt, E. R.; Coleridge, B. M. Tetrahedron Lett. 2008, 49, 5152–5155; For Sc(OTf)3-Hantzsch dihydropyridine, see: Itoh, T.; Nagata, K.;Kurihara, A.; Miyazaki, M.; Ohsawa, A. Tetrahedron Lett. 2002, 43, 3105–3108.

10. For a review of the Leuckart reaction, see: Moore, M. L. In Organic Reactions;Adams, R., Ed.; John Wiely & Sons, Inc.: New York, 1948; Vol. V, pp 301–330.

11. (a) Leuckart, R. Ber. Dtsch. Chem. Ges. 1886, 18, 2341–2344; (b) Leuckart, R.;Bach, E. Ber. Dtsch. Chem. Ges. 1886, 19, 2128–2133.

12. For examples, see: (a) Johns, I. B.; Burch, J. M. J. Am. Chem. Soc. 1938, 60, 919–920; (b) Carlson, R.; Lejon, T.; Lundstedt, T.; Le Clouerec, E. Acta Chem. Scand.1993, 47, 1046–1049.

13. For examples, see: (a) Bach, R. D. J. Org. Chem. 1968, 33, 1647–1649; (b) Loupy,A.; Monteux, D.; Petit, A.; Aizpurua, J. M.; Domínguez, E.; Palomo, C.Tetrahedron Lett. 1996, 37, 8177–8180.

14. deBenneville, P. L.; Macartney, J. H. J. Am. Chem. Soc. 1950, 72, 3073–3075.15. Ingersoll, A. W.; Brown, J. H.; Kim, C. K.; Beauchamp, W. D.; Jennings, G. J. Am.

Chem. Soc. 1936, 58, 1808–1811.16. (a) Awachie, P. I.; Agwada, V. C. Tetrahedron 1990, 46, 1899–1910; (b) Ito, K.;

Oba, H.; Sekiya, M. Bull. Chem. Soc. Jpn. 1976, 49, 2485–2490; (c) Pollard, C. B.;Young, D. C., Jr. J. Org. Chem. 1951, 16, 661–672; (d) Webers, V. J.; Bruce, W. F. J.Am. Chem. Soc. 1948, 70, 1422–1424; (e) Alexander, E. R.; Wildman, R. B. J. Am.Chem. Soc. 1948, 70, 1187–1189; (f) Ref. 12b.

17. Saba, S.; Brescia, A.; Kaloustian, M. K. Tetrahedron Lett. 1991, 32, 5031–5034.

18. Saba, S.; Vrkic, D.; Cascella, C.; DaSilva, I.; Carta, K.; Kojtari, A. J. Chem. Res. 2008,6, 301–304.

19. For a review on the use of ammonium formate in hydrogen transfer reactions,see: Ram, S.; Ehrenkaufer, R. E. Synthesis 1988, 91–95.

20. Typical procedure for one-pot preparation of secondary and tertiary amines 4(Table 1, entry 10): To a 50-mL round-bottomed flask containing ammoniumformate (0.69 g, 11 mmol) was added piperonal (1.65 g, 11 mmol) followed bypyrrolidine (0.78 g, 11 mmol) and toluene (20 mL). The mixture wasmagnetically stirred and heated at reflux with continuous removal of thewater formed (Dean–Stark trap). The mixture was then cooled to roomtemperature and the solvent was removed by rotary evaporation. The crudeproduct was purified by distillation under reduced pressure and afforded theamine as a colorless oil (1.65 g, 73%; bp 163–165 �C/11 torr); 1H NMR(300 MHz, CDCl3): d 6.85 (s, 1H), 6.75–6.72 (m, 2H), 5.92 (s, 2H), 3.51 (s, 2H),2.50–2.46 (m, 4H), 1.79–1.73 (m, 4H); 13C NMR (75 MHz, CDCl3): d 147.48,146.34, 133.38, 121.81, 109.34, 107.82, 100.77, 60.40, 53.99, 23.37.