Next Article in Journal / Special Issue
Nucleophilic Benzoylation Using a Mandelic Acid Dioxolanone as a Synthetic Equivalent of the Benzoyl Carbanion. Oxidative Decarboxylation of α-Hydroxyacids
Previous Article in Journal / Special Issue
Masked ω-Lithio Ester Enolates: Synthetic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polymer-Supported Cinchona Alkaloid-Derived Ammonium Salts as Recoverable Phase-Transfer Catalysts for the Asymmetric Synthesis of α-Amino Acids

by
Rafael Chinchilla
,
Patricia Mazón
and
Carmen Nájera
*
Departamento de Química Orgánica and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo. 99, 03080 Alicante, Spain
*
Author to whom correspondence should be addressed.
Molecules 2004, 9(5), 349-364; https://doi.org/10.3390/90500349
Submission received: 2 March 2004 / Accepted: 22 March 2004 / Published: 30 April 2004

Abstract

:
Alkaloids such as cinchonidine, quinine and N-methylephedrine have been N-alkylated using polymeric benzyl halides or co-polymerized and then N-alkylated, thus affording a series of polymer-supported chiral ammonium salts which have been employed as phase-transfer catalysts in the asymmetric benzylation of an N-(diphenylmethylene)glycine ester. These new polymeric catalysts can be easily recovered by simple filtration after the reaction and reused. The best ee’s were achieved when Merrifield resin-anchored cinchonidinium ammonium salts were employed.

Introduction

The synthesis of optically active α-amino acids using simple and easily scalable procedures is an important synthetic challenge due to their industrial interest [1]. Amongst all the reported methodologies, the enantioselective synthesis of α-amino acids employing easily available and re-usable chiral catalysts presents clear synthetic advantages for large-scale procedures. Particularly, the phase-transfer catalysis (PTC) [2] methodology applied to the asymmetric alkylation of glycine and alanine Schiff bases is the most simple and easy to scale up. Thus, N-arylmethyl substituted Cinchona alkaloid-derived ammonium salts such as cinchonidine derivatives of the type 1 have been employed as chiral phase-transfer catalysts [3] in the asymmetric alkylation of iminic glycinates, first by O’Donnell [4] (1, R = H, Ar = Ph, X = Cl) and then by Corey [5] (1, R = allyl, Ar = 9-anthryl, X = Br) and Lygo [6] (1, R = H, Ar = 9-anthryl, X = Cl). Interestingly, an opposite sense of the asymmetric induction can be observed changing the alkaloid moiety from cinchonidine [(S)-enantioselectivity] to its so-called pseudoenantiomer cinchonine [(R)-enantioselectivity] [7]. Moreover, dimeric [8], trimeric [9] and even dendrimeric [10] Cinchona alkaloid-derived catalysts, as well as non-Cinchona-derived species such as spiroammonium [11] and phosphonium salts [12], TADDOL [13a,b] and other tartaric acid derivatives [13c,d], guanidinium salts [13e], binaphthyl-derived amines [13b,14] and salen-metal complexes [15] have also been used in these kinds of asymmetric PTC alkylations.
Attaching the chiral catalyst to a solid support can be considered a next step in the development of the PTC methodology due to the resulting ease of separation and possible recycling. As a result, the preparation and uses of all kind of supported reagents is considered nowadays a fast developing topic [16]. Polymeric Cinchona alkaloids have been previously used as catalysts [3] in other processes such as asymmetric Michael addition [17], dihydroxylation [18] and aminohydroxylation [19] reactions. However, the use of polymeric Cinchona alkaloid-derived ammonium salts as PTC catalysts for the asymmetric synthesis of α-amino acid derivatives is very recent and limited. Thus, N-supported Merrifield resin-derived ammonium salts 2 (n = 1, R = H, X = Cl [20]; n = 4, 6 or 8, R = H, OMe, X = I [21]) from cinchonidine and quinine (R = OMe) have been prepared, as well O-supported polymeric derivatives such as 3 [22] (R = H, OMe, Ar = 9-anthryl) and N- and O-alkylated polyethylene glycol (PEG) monomethyl ether ammonium salt derivatives such as 4 [23a] (R = H, OMe, Ar = 9-anthryl) and 5 [R = H, OMe, Ar = 9-anthryl, oblong circle = MeOPEG5000O2C- [23a]; R = H, Ar = 9-anthryl, oblong circle = MeOPEG5000O-C6H4-(CH2)3O- [23b] ], respectively. In addition, a quinine-derived N-methylanthryl ammonium salt attached to a PEG chain at the 6-position of the quinoleine nucleus has been prepared [23b]. All these polymers have being used for the asymmetric alkylation of glycinate imines.
Molecules 09 00349 i001
In this context, and as part of our ongoing studies towards the synthesis of easily recoverable and reusable PTC catalysts for the asymmetric synthesis of α-amino acids [8b,10,20c], we describe in this paper the preparation of a series of ammonium salts derived from alkaloids such as cinchonidine (6), quinine (7) and N-methylephedrine (8), supported mainly at the nitrogen to an array of commercially available or easily prepared polymers, as well as their use as chiral catalysts in the model asymmetric benzylation reaction of a N-(diphenylmethylene)glycine ester under PTC conditions.
Molecules 09 00349 i002

Results and Discussion

Polymer-supported ammonium salt 2a was obtained as previously reported [20c] by N-alkylation of cinchonidine (6) with the Merrifield resin (Fluka, polystyrene crosslinked with 1% divinylbenzene, 1.7 meq Cl/g resin) in refluxing toluene, and is included in this study for comparison. When the N-alkylation reaction was carried out on O-allyl cinchonidine [5a] using Merrifield resin the O-allylated polymer 2b was obtained, whereas resin 2c was prepared similarly to resin 2a, but using quinine (7) instead of cinchonidine. Moreover, we also prepared in the same way the (1R,2S)-N-methylephedrine (8)-supported resin 9, which has been previously used in the PTC ethylation of α-cyanotoluene giving rather low ee’s [24], although its use in the alkylation of glycinimides was never attempted.
After the preparation of these Merrifield resin-derived chiral ammonium salts, we thought to explore the influence of the support in the performance of the polymeric ammonium salts as PTC catalysts. Thus, we prepare the N-tritylated polymer-supported cinchonidine 10a by reaction of 6 with the corresponding polymer bound triphenylchloromethane (Fluka, 1% DVB, 1.1 mmol Cl/g resin), and also its O-allylated counterpart 10b, similarly to 2b. In addition, the cinchonidine 6 was N-anchored to chloromethylated 1-[4-(4-vinylphenoxy)butoxy]-4-vinylbenzene-crosslinked polystyrene (JandaJelTM-Cl, Aldrich, 2% crosslinking, 0.45-0.7 mmol Cl/g resin) and to Wang-Br resin (Novabiochem, 1% DVB, 100-200 mesh, 1.19 mmol Br/g resin) to afford polymeric ammonium salts 11 and 12, respectively. Finally, polymer-supported cinchonidine 13a was obtained by reaction of 6 with chloromethylated polyethyleneglycol-polystyrene copolymer (ArgoGelTM-Cl, Argonaut Technologies, 1% DVB, 0.4 mmol Cl/g resin), whereas its O-allylated analogue 13b was prepared as above.
Molecules 09 00349 i003
After the preparation of all these supported ammonium salts from commercially available polymers, we thought of the synthesis of an anchored cinchonidinium-derived ammonium salt incorporating a 9-anthrylmethyl moiety, which has shown its efficiency as an enantioselectivity-increasing group in Cinchona-derived PTC catalysts [5,6]. Thus, we prepared the mercapto resin 15 by treatment of the Merrifield resin (14) with thiourea and subsequent hydrolysis (Scheme 1) [25]. This resin was deprotonated with sodium hydride and reacted with 9,10-dichloromethylanthracene [26] (2 equiv) and subsequently with cinchonidine (6), affording polymeric salt 16.
Scheme 1.
Scheme 1.
Molecules 09 00349 g001
The incorporation of the alkaloid in all these obtained catalysts was demosntrated by the presence of new bands in the IR spectra attributable to the alkaloid structure, and also by the increase in the initial resin weight and also the elemental analysis, which also allowed the determination of the loading.
Finally, we also obtained co-polymeric cinchonidine-derived ammonium salt 18 by N-alkylation of an acrylonitrile and cinchonidine co-polymer 17 [27]. Thus, a mixture of acrylonitrile and cinchonidine (9:1 molar ratio) and a catalytic amount of azabisisobutyronitrile (AIBN) was heated in degassed DMF at 90ºC affording after precipitation the co-polymer 17 (Scheme 2), which reacted with 9-chloro-methylanthracene to give co-polymeric ammonium salt 18.
Scheme 2.
Scheme 2.
Molecules 09 00349 g002
Polymers 2, 9-13 as well as 16 and 18 were used as insoluble PTC catalysts (0.1 eq) in the model triphase benzylation reaction of glycine-derived N-(diphenylmethylene)glycine isopropyl ester 19 [28] with benzyl bromide in an organic solvent and using an aqueous base (Scheme 3). The isopropyl ester 19 was chosen, instead of the tert-butyl derivative usually employed in asymmetric PTC alkylations, due to preliminary experiments using polymeric ammonium salts such as 2a, which showed higher ee’s and lower reaction times in the alkylation of this glycine derivative [20c]. In addition, toluene was used as solvent and 25% aq NaOH as base when working at r.t. or 0ºC, whereas a mixture of toluene/CHCl3 (7:3 v/v) and 50% aq KOH was used when lower reaction temperatures were employed [8c]. The resulting yields and ee’s are summarized in Table 1. In all cases the (S)-enantiomer 20 was obtained. The ee values were determined by chiral GLC analysis from the corresponding trifluoroacetamide, obtained after 2M HCl hydrolysis of the imine 20 and further reaction with trifluoroacetic anhydride [29]. A racemic reference sample of 20 was prepared using tetrabutylammonium bromide as phase-transfer catalyst.
Scheme 3.
Scheme 3.
Molecules 09 00349 g003
From the results shown in Table 1 it can be observed that the Merrifield-anchored cinchonidine-derived ammonium salt 2a afforded the higher ee at 0ºC (90%, Table 1, entry 2), and lowering the reaction temperature further did not produce an increase in the ee (entries 3 and 4). However, its allylated counterpart 2b gave place to considerably lower ee values, both at r.t. (entry 5) or 0ºC (entry 6). The analogous polymeric quinine derivative 2c was clearly less effective as an asymmetric PTC catalyst than its structurally similar cinchonidine-derivative 2a, giving very low ee’s (Table 1, entries 7 and 8). Moreover, the (1R,2S)-N-methylephedrine-derived resin 9 gave a low yield of an almost racemic 20 (Table 1, entry 9).
The polymeric trityl-anchored cinchonidine salt 10a gave a 44% ee of 20 working at r.t. (Table 1, entry 10), which was raised to 70% ee when the temperature was lowered to 0ºC (Table 1, entry 11), but showed no further increment when the reaction was carried out at -20ºC (Table 1, entry 12). Similarly to 2a, when the benzylation reaction was carried out with the O-allylated trityl-supported resin 10b resin, the ee dropped remarkably (Table 1, entry 13). In addition, the JandaJelTM-anchored cinchonidine ammonium salt 11 gave a 62% ee of 20 at r.t. (Table 1, entry 14) and a slightly lower 56% ee at 0ºC, with an observed increase of the reaction time (Table 1, entry 15), whereas the Wang-supported cinchonidinium salt 12 gave ca. 55% ee, both at r.t. or 0ºC (Table 1, entries 16 and 17). Moreover, ArgoGelTM-supported cinchonidinium salt 13a afforded up to 68% ee from r.t. to -20ºC (Table 1, entries 18-20) and again a tremendous drop in the ee was observed using its O-allylated counterpart 13b (Table 1, entry 21).
Furthermore, when the N-substituted 9-anthrylmethyl derivative 16 was used as PTC catalyst, up to 74% ee of 20 was obtained working at -20ºC (Table 1, entry 23). The use of the co-polymeric cinchonidinium ammonium salt 18 with an anthrylmethyl group at the N gave a 44% ee when working at r.t. (Table 1, entry 25), which increased to 70% ee, almost independently of reductions in the temperature to 0ºC or -20ºC (Table 1, entries 26 and 27).
Table 1. Enantioselective PTC benzylation of glycine derivative 19 using polymeric chiral catalysts
Table 1. Enantioselective PTC benzylation of glycine derivative 19 using polymeric chiral catalysts
EntryCatalystBaseSolventT (ºC)t (h)Yielda (%)eeb (%)
12a25% NaOHPhMe2549066
22a25% NaOHPhMe0179090
32a25% KOHPhMe:CHCl3-20104676
42a50% KOHPhMe:CHCl3-4095685
52b25% NaOHPhMe25106234
62b25% NaOHPhMe01402350
72c25% NaOHPhMe2587618
82c25% NaOHPhMe0968120
9925% NaOHPhMe0160332
1010a25% NaOHPhMe2519644
1110a25% NaOHPhMe0105970
1210a50% KOHPhMe:CHCl3-20187169
1310b25% NaOHPhMe012072
141125% NaOHPhMe2547662
151125% NaOHPhMe0107156
161225% NaOHPhMe25127856
171225% NaOHPhMe01206354
1813a25% NaOHPhMe25107563
1913a25% NaOHPhMe0209164
2013a50% KOHPhMe:CHCl3-20159068
2113b25% NaOHPhMe060288
221625% NaOHPhMe2529259
231625% NaOHPhMe0329470
241650% KOHPhMe:CHCl3-2038674
251825% NaOHPhMe2519644
261825% NaOHPhMe01704970
271850% KOHPhMe:CHCl3-20137471
a Crude yield determined by 1H NMR (300 MHz). b Determined by chiral GLC analysis from the corresponding trifluoroacetamide (see text).
After the benzylation reaction, the polymeric catalysts were filtered off from the reaction mixture and were reused up to three times without any loss of effectivity.

Conclusions

We have prepared chiral polymeric ammonium salts by anchoring a number of alkaloids, mainly from Cinchona, to different commercially available halogenated polymers and to a prepared anthryl-containing polystyrene. In addition, we have also obtained a cinchonidinium salt-acrylonitrile co-polymer. All these polymeric ammonium salts have been employed as solid-supported chiral PTC catalysts for the asymmetric benzylation of N-(diphenylmethylene)glycine isopropyl ester achieving moderate enantioselectivities. The best results were obtained using a Merrifield resin-supported cinchonidinium salt, the use quinine or ephedrine-derived ammonium salts affording poor results. In all cases higher ee’s were obtained when a hydroxyl group was present in the alkaloid moiety, their O-allylated counterparts giving lower enantioselectivities. Lowering the reaction temperature usually resulted in higher ee’s, although temperatures below 0 ºC generally did not affected remarkably the degree of asymmetric induction. All the supported catalysts could be separated from the reaction mixture by simple filtration and recycled.

Acknowledgments

We thank the Ministerio de Educación y Cultura (MEC) of Spain (DGICYT, research project BQU2001-0724-C02-01), the Generalitat Valenciana (GV99-33-1-02) and the Universidad de Alicante for financial support.

Experimental

General

Reagents and solvents from commercial suppliers were of the best grade available and used as provided unless otherwise stated. IR spectra were recorded with a Nicolet 510 P-FT. NMR spectra were measured with a Bruker AC-300 at 300 MHz for 1H- and 75 MHz for 13C- using TMS as internal standard. Elemental analyses were carried out by the Microanalytical Service at the Research Technical Services of the University of Alicante. Chiral GLC analysis were performed using a Chirasil-L-Val column (Chrompack), 1 min 85º, 2º/min to 180º.

General procedure for the preparation of the polymeric ammonium salts 2 and 9-13.

The corresponding halogenated polymer (1 meq) was added to a suspension of the alkaloid 6, 7 or 8 (2 mmol) in toluene (10 mL) and the mixture was stirred under reflux for 24 h. The reaction mixture was cooled to r.t. and the solid was filtered, washed with AcOEt (3 x 15 mL) and dried in vacuo, affording the polymer-supported ammonium salts 2a, 2c, 9, 10a, 11, 12 and 13a. The O-allylated polymeric ammonium salts 2b, 10b and 13b were obtained following the same procedure, but starting from O-allyl cinchonidine [5a].

Polymeric ammonium salt 2a

IR (KBr) cm-1: 3415 (broad), 3060, 2940, 1596, 1430, 1220, 1100, 750.
Microanalysis: % N = 3.25; loading = 1.7 mmol g-1

Polymeric ammonium salt 2b

IR (KBr) cm-1: 3080, 2939, 1590, 1440, 1231, 1103, 760.
Microanalysis: % N = 2.03; loading = 1.2 mmol g-1

Polymeric ammonium salt 2c

IR (KBr) cm-1: 3398 (broad), 3020, 2920, 615, 1497, 1455, 1241, 1012, 757.
Microanalysis: % N = 3.06; loading = 1.6 mmol g-1

Polymeric ammonium salt 9

IR (KBr) cm-1: 3382 (broad), 3033, 2925, 1596, 1499, 1441, 1011, 756.
Microanalysis: % N = 1.72; loading = 1.6 mmol g-1

Polymeric ammonium salt 10a

IR (KBr) cm-1: 3241 (broad), 3015, 2925, 1598, 1460, 1100, 763.
Microanalysis: % N = 1.89; loading = 0.9 mmol g-1

Polymeric ammonium salt 10b

IR (KBr) cm-1: 3054, 2929, 1645, 1448, 1148, 695.
Microanalysis: % N = 0.39; loading = 0.2 mol g-1

Polymeric ammonium salt 11

IR (KBr) cm-1: 3370 (broad), 3019, 2912, 1602, 1488, 1448, 1233, 1025, 751.
Microanalysis: % N = 1.39; loading = 0.6 mmol g-1

Polymeric ammonium salt 12

IR (KBr) cm-1: 3351 (broad), 3063, 2919, 1599, 1457, 1221, 1115, 754.
Microanalysis: % N = 2.51%; loading = 1.2 mmol g-1

Polymeric ammonium salt 13a

IR (KBr) cm-1: 3505 (broad), 3020, 2840, 1602, 1452, 1295, 1240, 1105, 699.
Microanalysis: % N = 0.37; loading = 0.4 mmol g-1

Polymeric ammonium salt 13b

IR (KBr) cm-1: 2924, 1618, 1460, 1356, 1252, 1110, 695.
Microanalysis: % N = 0.30; loading = 0.4 mmol g-1

Preparation of polymeric ammonium salt 16

A suspension of Merrifield resin (Fluka, 1% DVB, 1.7 meq. Cl/g resin) (3 meq, 1.76 g) and thiourea (12.78 mmol, 971 mg) in a mixture of THF (20 mL) and EtOH (6 mL) is refluxed for 2 days and the resin thus obtained is filtered and washed successively with water (3 x 20 mL), THF (3 x 20 mL) and benzene (3 x 20 mL). The solid was then suspended in benzene (27 mL) and a mixture of tetrabutylammonium iodide (0.08 mmol, 28 mg) and NaOH (25.95 mmol, 1.038 g) in water (1.4 mL) was added. The reaction mixture was stirred at 80 ºC under nitrogen for 2 days and the resulting solid was filtered and washed successively with THF (3 x 20 mL), water (3 x 20 mL), THF/6M HCl (3:1 v/v, 3 x 20 mL), water (3 x 20 mL), THF, (3 x 20 mL), acetone (3 x 20 mL), CH2Cl2 (3 x 20 mL) and MeOH (3 x 20 mL). After drying in vacuo (15 Torr), the resin 15 [25] (1.85 g) was obtained. A suspension of 15 (1 mmol, 654 mg) in toluene (12 mL) was treated with NaH (60% mineral oil, 1.2 mmol, 30 mg) and 9,10-dichloromethylanthracene [26] (2 mmol, 550 mg) was added. The mixture was refluxed 24 h and the resulting solid was filtered, washed with AcOEt (4 x 20 mL) and treated with cinchonidine (6) as in the preparation of the former ammonium salts (see above), to give polymeric ammonium salt 16.
IR (KBr) cm-1: 3399 (broad), 3049, 3015, 2910, 1598, 1499, 1451, 755, 698.
Microanalysis: % N = 2.55; loading = 1.4 mmol g-1

Preparation of co-polymeric ammonium salt 18 [27]

A degassed solution of cinchonidine (6) (4 mmol, 1.178 g), acrylonitrile (36.4, 2.4 mL) and AIBN (0.3 mmol, 48 mg) in DMF (12 mL) was heated in a pressure tube at 90 ºC for 48 h. The mixture was cooled to r.t., water (15 mL) and AcOEt (15 mL) were added and the solid was filtered, washed with AcOEt (5 x 15 mL) and dried (15 Torr). To a solution of this solid (606 mg) in DMSO (10 mL) was added 9-chloromethylanthracene (8 mmol, 1.814 g) and the mixture was refluxed for 24 h. The mixture was cooled to r.t. and the solid was filtered, washed with AcOEt (5 x 15 mL) and dried (15 Torr) affording 18 (840 mg).
IR (KBr) cm-1: 3416 (broad), 2939, 2247, 1622, 1448, 1246, 1031, 749.
[α]D25 – 24 (c 0.5, DMSO)

General procedure for the benzylation of 19 using the polymeric ammonium salts as PTC catalysts: Preparation of isopropyl 2-diphenylmethylenamino-3-phenylpropanoate (20).

Benzyl bromide (0.6 mmol, 72 μL) was added to a stirred suspension of 19 [28] (0.5 mmol, 140 mg), the polymeric catalyst (0.1 eq) and the aqueous base (4 mL) in the appropriate solvent (5 mL) and at the selected temperature (see Table 1). When the reaction was considered finished (GLC), the mixture was filtered and the solid was washed with AcOEt (25 mL), thus giving the recovered polymeric catalyst. The filtrate was washed with water (3 x 15 mL) and the organic phase was dried (MgSO4), filtered off and evaporated (15 Torr) to give the title compound: 1H-NMR (CDCl3) δ: 1.19, 1.21 (6H, 2d, J = 6.5), 3.17 (1H, dd, J = 13.1, 9.2), 3.27 (1H, dd, J = 13.1, 4.3), 4.19 (1H, dd, J = 9.2, 4.3), 5.04 (1H, heptet, J = 6.1), 6.62 (1H, m), 7.03-7.60 (13H, m), 7.79 (1H, d, J = 7.9); 13C-NMR (CDCl3) δ: 21.7, 39.5, 67.3, 68.2, 126.1, 127.5, 127.8, 128.2, 128.6, 129.7, 129.9, 130.1, 132.3, 136.1, 138.0, 139.4, 170.5, 171.1; IR (thin film) cm-1: 3063, 3032, 1742, 1629; HRMS (EI) for C22H25NO2 (M+): Calcd 371.1885; Found 371.1847.

References and Notes

  1. For recent reviews: Seebach, D.; Sting, A.R.; Hoffmann, M. Self-regeneration of stereocenters (SRS) - applications, limitations, and abandonment of a synthetic principle. Angew. Chem. Int. Ed. Engl. 1996, 35, 2708–2748. [Google Scholar] Wirth, T. New strategies to α-alkylated α-amino acids. Angew. Chem. Int. Ed. Engl. 1997, 36, 225–227. [Google Scholar] Cativiela, C.; Díaz-de-Villegas, M.D. Stereoselective synthesis of quaternary α-amino acids. Part 1: acyclic compounds. Tetrahedron:Asymmetry 1998, 9, 3517–3599. [Google Scholar] Cativiela, C.; Díaz-de-Villegas, M.D. Stereoselective synthesis of quaternary α-amino acids. Part 2: cyclic compounds. Tetrahedron: Asymmetry 2000, 11, 645–732. [Google Scholar] Abellán, T.; Chinchilla, R.; Galindo, N.; Guillena, G.; Nájera, C.; Sansano, J.M. Glycine and alanine imines as templates for asymmetric synthesis of α-amino acids. Eur. J. Org. Chem. 2000, 2689–2697. [Google Scholar] Nájera, C. From α-amino acids to peptides: all you need for the journey. Synlett 2002, 1388–1403. [Google Scholar]
  2. Phase-Transfer Catalysis; Halpern, M.E. American Chemical Society: Washington D.C., 1997. [Google Scholar]
  3. Kacprzak, K.; Gawronski, J. Cinchona alkaloids and their derivatives: versatile catalysts and ligands in asymmetric synthesis. Synthesis 2001, 961–998. [Google Scholar]
  4. O’Donnell, M.J.; Bennett, W.D.; Wu, S. The stereoselective synthesis of α-amino acids by phase-transfer catalysis. J. Am. Chem. Soc. 1989, 111, 2353–2355. [Google Scholar] Lipkowitz, K.B.; Cavanaugh, M.W.; Baker, B.; O’Donnell, M.J. Theoretical studies in molecular recognition: asymmetric induction of benzophenone imine ester enolates by the benzylcinchoninium ion. J. Org. Chem. 1991, 56, 5181–5192. [Google Scholar] O’Donnell, M.J.; Wu, S. A catalytic enantioselective synthesis of α-methyl amino acid derivatives by phase-transfer catalysis. Tetrahedron: Asymmetry 1992, 3, 591–594. [Google Scholar] O’Donnell, M.J.; Wu, S.; Huffman, J.C. A new active catalyst species for enantioselective alkylation by phase-transfer catalysis. Tetrahedron 1994, 50, 4507–4518. [Google Scholar] O’Donnell, M.J. The preparation of optically active α-amino acids from the benzophenone imines of glycine derivatives. Aldrichimica Acta 2001, 34, 3–15. [Google Scholar]
  5. Corey, E.J.; Xu, F.; Noe, M.C. A rational approach to catalytic enantioselective enolate alkylation using a structurally rigidified and defined chiral quaternary ammonium salt under phase transfer conditions. J. Am. Chem. Soc. 1997, 119, 12414–12415. [Google Scholar] Corey, E.J.; Noe, M.C.; Xu, F. Highly enantioselective synthesis of cyclic and functionalized α-amino acids by means of a chiral phase transfer catalyst. Tetrahedron Lett. 1998, 39, 5347–5350. [Google Scholar] Horikawa, M.; Busch-Petersen, J.; Corey, E.J. Enantioselective synthesis of α-hydroxy-α-amino acid esters by aldol coupling using a chiral quaternary ammonium salt as catalyst. Tetrahedron Lett. 1999, 40, 3843–3846. [Google Scholar]
  6. Lygo, B.; Wainwright, P.G. A new class of asymmetric phase-transfer catalysts derived from Cinchona alkaloids - application in the enantioselective synthesis of α-amino acids. Tetrahedron Lett. 1997, 38, 8595–8598. [Google Scholar] Lygo, B.; Crosby, J.; Peterson, J.A. Enantioselective synthesis of bis-α-amino acid esters via asymmetric phase-transfer catalysis. Tetrahedron Lett. 1999, 40, 1385–1388. [Google Scholar] Lygo, B. Enantioselective synthesis of dityrosine and isodityrosine via asymmetric phase-transfer catalysis. Tetrahedron Lett. 1999, 40, 1389–1392. [Google Scholar] Lygo, B.; Crosby, J.; Peterson, J.A. Enantioselective alkylation of alanine-derived imines using quaternary ammonium catalysts. Tetrahedron Lett. 1999, 40, 8671–8674. [Google Scholar] Lygo, B.; Crosby, J.; Lowdon, T.R.; Peterson, J.A.; Wainwright, P.G. Studies on the enantioselective synthesis of α-amino acids via asymmetric phase-transfer catalysis. Tetrahedron 2001, 57, 2403–2409. [Google Scholar] Lygo, B.; Andrews, B. I.; Crosby, J.; Peterson, J.A. Asymmetric alkylation of glycine imines using in situ generated phase-transfer catalysts. Tetrahedron Lett. 2002, 43, 8015–8018. [Google Scholar]
  7. O’Donnell, M.J.; Delgado, F.; Pottorf, R.S. Enantioselective solid-phase synthesis of α-amino acid derivatives. Tetrahedron 1999, 55, 6347–6362. [Google Scholar]
  8. Jew, S.; Jeong, B.; Yoo, M.; Huh, H.; Park, H. Synthesis and application of dimeric Cinchona alkaloid phase-transfer catalysts: α,α'-bis[O(9)-allylcinchonidinium]-o, m, or p-xylene dibromide. Chem. Commun. 2001, 1244–1245. [Google Scholar] Park, H.; Jeong, B.; Yoo, M.; Lee, J.; Park, M.; Lee, Y.; Kim, M.; Jew, J. Highly enantioselective and practical Cinchona-derived phase-transfer catalysts for the synthesis of α-amino acids. Angew. Chem. Int. Ed. 2002, 41, 3036–3038. [Google Scholar] Chinchilla, R.; Mazón, P.; Nájera, C. New dimeric anthracenyl-derived Cinchona quaternary ammonium salts as phase-transfer catalysts for the asymmetric synthesis of α-amino acids. Tetrahedron:Asymmetry 2002, 13, 927–931. [Google Scholar]
  9. Park, H.; Jeong, B.; Yoo, M.; Park, M.; Huh, H.; Jew, S. Trimeric Cinchona alkaloid phase-transfer catalyst: α,α',α''-tris[O(9)-allylcinchonidinium]mesitylene tribromide. Tetrahedron Lett. 2001, 42, 4645–4648. [Google Scholar]
  10. Mazón, P.; Chinchilla, R.; Nájera, C.; Guillena, G.; Kreiter, R.; Klein Gebbink, R.J.M.; van Koten, G. Unexpected metal base-dependent inversion of the enantioselectivity in the asymmetric synthesis of α-amino acids using phase-transfer catalysts derived from cinchonidine. Tetrahedron:Asymmetry 2002, 13, 2181–2185. [Google Scholar] Guillena, G.; Kreiter, R.; van de Coevering, R.; Klein Gebbink, R.J.M.; van Koten, G.; Mazón, P.; Chinchilla, R.; Nájera, C. Chiroptical properties and applications in PTC of new dendritic cinchonidine-derived ammonium salts. Tetrahedron:Asymmetry 2003, 14, 3705–3712. [Google Scholar]
  11. Ooi, T.; Kameda, M.; Maruoka, K. Molecular design of a C2-symmetric chiral phase-transfer catalyst for practical asymmetric synthesis of α-amino acids. J. Am. Chem. Soc. 1999, 121, 6519–6520. [Google Scholar] Ooi, T.; Tayama, E.; Doda, K.; Takeuchi, M.; Maruoka, K. Dramatic rate enhancement by ultrasonic irradiation in liquid-liquid phase-transfer catalytic reactions. Synlett 2000, 1500–1502. [Google Scholar] Ooi, T.; Takeuchi, M.; Kameda, M.; Maruoka, K. Practical catalytic enantioselective synthesis of α,α-dialkyl-α-amino acids by chiral phase-transfer catalysis. J. Am. Chem. Soc. 2000, 122, 5228–5229. [Google Scholar] Ooi, T.; Uematsu, Y.; Maruoka, K. Evaluation of the efficiency of the chiral quaternary ammonium salt β-Np-NAS-Br in the organic-aqueous phase-transfer alkylation of a protected glycine derivative. Adv. Synth. Catal. 2002, 344, 288–291. [Google Scholar] Ooi, T.; Uematsu, Y.; Maruoka, K. Highly enantioselective alkylation of glycine methyl and ethyl ester derivatives under phase-transfer conditions: its synthetic advantage. Tetrahedron Lett. 2004, 45, 1675–1678. [Google Scholar]
  12. Manabe, K. Asymmetric phase-transfer alkylation catalyzed by a chiral quaternary phosphonium salt with a multiple hydrogen-bonding site. Tetrahedron Lett. 1998, 39, 5807–5810. [Google Scholar] Manabe, K. Synthesis of novel chiral quaternary phosphonium salts with multiple hydrogen-bonding sites and their application to asymmetric phase-transfer alkylation. Tetrahedron 1998, 54, 14465–14476. [Google Scholar]
  13. Belokon’, Y.N.; Kotchetkov, K.A.; Churkina, T.D.; Ikonnikov, N.S.; Chesnokov, A.A.; Larionov, A.V.; Parmár, V.S.; Kumar, R.; Kagan, H.B. Asymmetric PTC C-alkylation mediated by TADDOL-novel route to enantiomerically enriched α-alkyl-α-amino acids. Tetrahedron: Asymmetry 1998, 9, 851–857. [Google Scholar] Belokon’, Y.N.; Kotchetkov, K.A.; Churkina, T.D.; Ikonnikov, N.S.; Chesnokov, A.A.; Larionov, A.V.; Singh, I.; Parmár, V.S.; Vyskocil, S.; Kagan, H.B. Asymmetric PTC C-alkylation catalyzed by chiral derivatives of tartaric acid and aminophenols. Synthesis of (R)- and (S)-α-methyl amino acids. J. Org. Chem. 2000, 65, 7041–7048. [Google Scholar] Shibuguchi, T.; Fukuta, Y.; Akachi, Y.; Sekine, A.; Oshima, T.; Shibasaki, M. Development of new asymmetric two-center catalysts in phase-transfer reactions. Tetrahedron Lett. 2002, 43, 9539–9543. [Google Scholar] Arai, S.; Tsuji, R.; Nishida, A. Phase-transfer-catalyzed asymmetric Michael reaction using newly-prepared chiral quaternary ammonium salts derived from l-tartrate. Tetrahedron Lett. 2002, 43, 9535–9537. [Google Scholar] Kita, T.; Georgieva, A.; Hashimoto, Y.; Nakata, T.; Nagasawa, K. C2-symmetric chiral pentacyclic guanidine: A phase-transfer catalyst for the asymmetric alkylation of tert-butyl glycinate Schiff base. Angew. Chem. Int. Ed. 2002, 41, 2832–2834. [Google Scholar]
  14. Belokon’, Y.N.; Kotchetkov, K.A.; Churkina, T.D.; Ikonnikov, N.S.; Vyskocil, S.; Kagan, H. B. Enantiomerically enriched (R)- and (S)-α -methylphenylalanine via asymmetric PTC C-alkylation catalyzed by NOBIN. Tetrahedron:Asymmetry 1999, 10, 1723–1728. [Google Scholar] Belokon’, Y.N.; Kotchetkov, K.A.; Churkina, T.D.; Ikonnikov, N.S.; Larionov, O.V.; Harutyunyan, S.R.; Vyskocil, S.; North, M.; Kagan, H.B. Highly efficient catalytic synthesis of α-amino acids under phase-transfer conditions with a novel catalyst/substrate pair. Angew Chem. Int. Ed. 2001, 40, 1948–1951. [Google Scholar] Casas, J.; Nájera, C.; Sansano, J.M.; González, J.; Saá, J.M.; Vega, M. Enantioselective synthesis of (S)-α-methylphenylalanine using (S)-BINOLAMs as new phase-transfer catalysts. Tetrahedron:Asymmetry 2001, 12, 699–702. [Google Scholar]
  15. Belokon’, Y.N.; North, M.; Kublitski, V.S.; Ikonnikov, N.S.; Krasik, P.E.; Maleev, V.L. Chiral salen-metal complexes as novel catalysts for asymmetric phase transfer alkylations. Tetrahedron Lett. 1999, 40, 6105–6108. [Google Scholar] Belokon’, Y.N.; Davies, R.G.; North, M. A practical asymmetric synthesis of α-methyl α-amino acids using a chiral Cu-salen complex as a phase transfer catalyst. Tetrahedron Lett. 2000, 41, 7245–7248. [Google Scholar]
  16. For recent reviews on polymer-supported chemistry see: Lorsbach, B. A.; Kurth, M. J. Carbon-carbon bond forming solid-phase reactions. Chem. Rev. 1999, 99, 1549–1581. [Google Scholar] James, I.W. Linkers for solid phase organic synthesis. Tetrahedron 1999, 55, 4855–4946. [Google Scholar] Shuttleworth, S.J.; Allin, S.M.; Wilson, R.D.; Nasturica, D. Functionalized polymers in organic chemistry; Part 2. Synthesis 2000, 1035–1074. [Google Scholar] Ley, S.V.; Baxendale, I.R.; Bream, R.N.; Jackson, P.S.; Leach, A.G.; Longbottom, D.A.; Nesi, M.; Scott, J.S.; Storer, R.I.; Taylor, S.J. Multi-step organic synthesis using solid-supported reagents and scavengers: a new paradigm in chemical library generation. J. Chem. Soc., Perkin Trans. 1 2000, 3815–4195. [Google Scholar] Orain, D.; Ellard, J.; Bradley, M. Protecting Groups in Solid-Phase Organic Synthesis. J. Comb. Chem. 2002, 4, 1–16. [Google Scholar] Benaglia, M.; Puglisi, A.; Cozzi, F. Polymer-Supported Organic Catalysts. Chem. Rev. 2003, 103, 3401–3429. [Google Scholar]
  17. Hermann, K.; Wynberg, H. Polymer-bound Cinchona alkaloids as catalysts in the Michael reaction. Helv. Chim. Acta 1977, 60, 2208–2212. [Google Scholar] Kobayashi, N.; Iwai, K. Functional polymers. 1. Poly(cinchona alkaloid-co-acrylonitrile)s. New polymer catalysts for asymmetric synthesis. J. Am. Chem. Soc. 1978, 100, 7071–7072. [Google Scholar] Hodge, P.; Khoshdel, E.; Waterhouse, J. Michael reactions catalyzed by polymer-supported quaternary ammonium salts derived from cinchona and ephedra alkaloids. J. Chem. Soc., Perkin Trans. 1 1983, 2205–2209. [Google Scholar] Hodge, P.; Khoshdel, E.; Waterhouse, J.; Fréchet, J.M.J. Michael additions catalyzed by Cinchona alkaloids bound via their vinyl groups to preformed crosslinked polymers. J. Chem. Soc., Perkin Trans. 1 1985, 2327–2331. [Google Scholar] Inagaki, M.; Hiratane, J.; Yamamoto, Y.; Oda, J. Asymmetric induction in the base-catalyzed reactions using polymer-supported quinines with spacer groups. Bull. Chem. Soc. Jpn. 1987, 60, 4121–4126. [Google Scholar] Sera, A.; Takagi, K.; Katayama, H.; Yamada, H.; Matsumoto, K. High-pressure asymmetric Michael additions of thiols, nitromethane, and methyl oxoindancarboxylate to enones. J. Org. Chem. 1988, 53, 1157–1161. [Google Scholar] Alvarez, R.; Hourdin, M.-A.; Cavé, C.; d’Angelo, J.; Chaminade, P. New polymer-supported catalysts derived from cinchona alkaloids: their use in the asymmetric Michael reaction. Tetrahedron Lett. 1999, 40, 7091–7094. [Google Scholar]
  18. Han, H.; Janda, K.D. Soluble Polymer-Bound Ligand-Accelerated Catalysis: Asymmetric Dihydroxylation. J. Am. Chem. Soc. 1996, 118, 7632–7633. [Google Scholar] Han, H.; Janda, K.D. A soluble polymer-bound approach to the Sharpless catalytic asymmetric dihydroxylation (AD) reaction: preparation and application of a [(DHQD)2PHAL-PEG-OMe) ligand. Tetrahedron Lett. 1997, 38, 1527–1530. [Google Scholar] Nandanan, E.; Sudalai, A.; Ravindranathan, T. New polymer supported Cinchona alkaloids for heterogeneous catalytic asymmetric dihydroxylation of olefins. Tetrahedron Lett. 1997, 38, 2577–2580. [Google Scholar] Canali, L.; Song, E.S.; Sherrington, D.C. Polymer-supported bis-Cinchona alkaloid ligands for asymmetric dihydroxylation of alkenes-a cautionary tale. Tetrahedron: Asymmetry 1998, 9, 1029–1034. [Google Scholar]
  19. Song, C.E.; Oh, C.R.; Lee, S.W.; Lee, S.; Canali, L.; Sherrington, D.C. Heterogeneous asymmetric aminohydroxylation of alkenes using a silica gel-supported bis-Cinchona alkaloid. Chem. Commun. 1998, 2435–2436. [Google Scholar]
  20. Wang, Y.; Zhengpu, Z.; Zhen, W.; Jiben, M.; Hodge, P. The asymmetric synthesis of amino acids under polymer-supported phase transfer catalytic condition. Chin. J. Polym. Sci. 1998, 16, 356–361. [Google Scholar] Zhengpu, Z.; Yongmei, W.; Zhen, W.; Hodge, P. Asymmetric synthesis of a-amino acids using polymer-supported chiral phase transfer catalysts. React. Funct. Polym. 1999, 41, 37–43. [Google Scholar] Chinchilla, R.; Mazón, P.; Nájera, C. Asymmetric synthesis of a-amino acids using polymer-supported Cinchona alkaloid-derived ammonium salts as chiral phase-transfer catalysts. Tetrahedron: Asymmetry 2000, 11, 3277–3281. [Google Scholar]
  21. Thierry, B.; Plaquevent, J.-C.; Cahard, D. New polymer-supported chiral phase-transfer catalysts in the asymmetric synthesis of α-amino acids: the role of a spacer. Tetrahedron: Asymmetry 2001, 12, 983–986. [Google Scholar]
  22. Thierry, B.; Perrard, T.; Audouard, C.; Plaquevent, J.-C.; Cahard, D. Solution- and solid-phase approaches in asymmetric phase-transfer catalysis by Cinchona alkaloid derivatives. Synthesis 2001, 1742–1746. [Google Scholar]
  23. Thierry, B.; Plaquevent, J.-C.; Cahard, D. Poly(ethylene glycol) supported Cinchona alkaloids as phase transfer catalysts: application to the enantioselective synthesis of α-amino acids. Tetrahedron: Asymmetry 2003, 14, 1671–1677. [Google Scholar] Danelli, T.; Annunziata, R.; Benaglia, M.; Cinquini, M.; Cozzi, F.; Tocco, G. Immobilization of catalysts derived from Cinchona alkaloids on modified poly(ethylene glycol). Tetrahedron: Asymmetry 2003, 14, 461–467. [Google Scholar]
  24. Chiellini, E.; Solaro, R. Stereo-ordered macromolecular matrixes bearing ammonium groups as catalysts in alkylation and carbenation reactions. J. Chem. Soc., Chem. Commun. 1977, 231–232. [Google Scholar]
  25. Fréchet, J.M.J.; de Smet, M.D.; Marc, D.; Farrall, M.J. Functionalization of crosslinked polystyrene resins. 2. Preparation of nucleophilic resins containing hydroxyl or thiol functionalities. Polymer 1979, 20, 675–680. [Google Scholar]
  26. Prepared by bubbling HCl (g) through a solution of anthracene and paraformaldehyde in dioxane: Miller, M.W.; Amidon, R.W.; Tawney, P.O. Some meso substituted anthracenes. I. 9,10-Bis-(chloromethyl)anthracene as a synthetic intermediate. J.Am. Chem. Soc. 1955, 77, 2845–2848. [Google Scholar]
  27. Kobayashi, N.; Iwai, K. Functional polymers. 1. Poly(cinchona alkaloid-co-acrylonitrile)s. New polymer catalysts for asymmetric synthesis. J. Am. Chem. Soc. 1978, 100, 7071–7072. [Google Scholar]
  28. Prepared in 80% overall yield by reaction of glycine with thionyl chloride in the presence of isopropanol (Patel, R.P.; Price, S. Synthesis of benzyl esters of α-amino acids. J. Org. Chem. 1965, 30, 3575–3576. [Google Scholar] ), followed by treatment of the crude product with benzophenone imine (O’Donnell, M.J.; Polt, R.L. A mild and efficient route to Schiff base derivatives of amino acids. J. Org. Chem. 1982, 47, 2663–2666. [Google Scholar] ).
  29. Oppolzer, W.; Moretti, R.; Zhou, C. Asymmetric alkylations of a sultam-derived glycine equivalent: practical preparation of enantiomerically pure α-amino acids. Helv. Chim. Acta 1994, 77, 2363–2380. [Google Scholar]
  • Sample Availability: Samples of the catalysts are available from the authors.

Share and Cite

MDPI and ACS Style

Chinchilla, R.; Mazón, P.; Nájera, C. Polymer-Supported Cinchona Alkaloid-Derived Ammonium Salts as Recoverable Phase-Transfer Catalysts for the Asymmetric Synthesis of α-Amino Acids. Molecules 2004, 9, 349-364. https://doi.org/10.3390/90500349

AMA Style

Chinchilla R, Mazón P, Nájera C. Polymer-Supported Cinchona Alkaloid-Derived Ammonium Salts as Recoverable Phase-Transfer Catalysts for the Asymmetric Synthesis of α-Amino Acids. Molecules. 2004; 9(5):349-364. https://doi.org/10.3390/90500349

Chicago/Turabian Style

Chinchilla, Rafael, Patricia Mazón, and Carmen Nájera. 2004. "Polymer-Supported Cinchona Alkaloid-Derived Ammonium Salts as Recoverable Phase-Transfer Catalysts for the Asymmetric Synthesis of α-Amino Acids" Molecules 9, no. 5: 349-364. https://doi.org/10.3390/90500349

Article Metrics

Back to TopTop