Next Article in Journal
Evaluation of Sulfated Polysaccharides from the Brown Seaweed Dictyopteris Justii as Antioxidant Agents and as Inhibitors of the Formation of Calcium Oxalate Crystals
Next Article in Special Issue
Examination of the Potential for Adaptive Chirality of the Nitrogen Chiral Center in Aza-Aspartame
Previous Article in Journal
Ultrasound-Promoted One-Pot, Four-Component Synthesis of Pyridin-2(1H)-One Derivatives
Previous Article in Special Issue
Deracemization of Axially Chiral Nicotinamides by Dynamic Salt Formation with Enantiopure Dibenzoyltartaric Acid (DBTA)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Perfluoroalkanesulfonamide Organocatalysts for Asymmetric Conjugate Additions of Branched Aldehydes to Vinyl Sulfones

1
Tokyo University of Pharmacy and Life Sciences, 1432-1 Horinouchi, Hachioji, Tokyo 192-0392, Japan
2
Gifu Pharmaceutical University 1-25-4, Daigaku-nishi, Gifu 501-1196, Japan
*
Author to whom correspondence should be addressed.
Molecules 2013, 18(12), 14529-14542; https://doi.org/10.3390/molecules181214529
Submission received: 30 October 2013 / Revised: 20 November 2013 / Accepted: 21 November 2013 / Published: 25 November 2013
(This article belongs to the Special Issue Dynamic Stereochemistry)

Abstract

:
Asymmetric conjugate additions of branched aldehydes to vinyl sulfones promoted by sulfonamide organocatalyst 6 or 7 have been developed, allowing facile synthesis of the corresponding adducts with all-carbon quaternary stereocenters in excellent yields with up to 95% ee.

1. Introduction

All-carbon quaternary stereocenters are one of the most important motifs in many natural products and bioactive compounds; however, relatively harsh reaction conditions are required to construct these stereocenters due to their steric hindrance. In addition, combinations of electrophile and nucleophile are limited, and the stereoselective construction of all-carbon quaternary stereocenters is not generally straightforward. Therefore, development of efficient synthetic methods to stereoselectively construct all-carbon quaternary stereocenters under mild reaction conditions is highly desirable in organic synthesis [1,2,3,4,5,6,7,8,9,10,11,12,13]. Among various methodologies to construct these centers, organocatalysis is one of the most effective processes that can be performed under mild conditions. The synthetic methods for compounds with quaternary stereogenic centers using organocatalysts have received considerable attention, particularly in the field of green chemistry [14,15]. Michael additions of various carbonyl compounds to 1,1-bis(benzenesulfonyl)ethylene (11) using organocatalysts are efficient synthetic methods, and several research groups have reported findings in this area [16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33]; however, successful conjugate additions of α-branched aldehydes and 11 for the construction of such all-carbon quaternary stereocenters have been rarely reported [34,35,36,37,38]. Alexakis and coworkers reported that l-proline derivatives catalyze the reaction of 11 with α-branched aldehyde 12a to give the corresponding adduct 13a with up to 73% ee [34,35]. Lu and coworkers reported that the sulfonamide organocatalyst derived from l-threonine promotes the conjugate addition of 12a to 11 in the unusual reaction solvent p-fluorotoluene to afford the corresponding adduct 13a in high yield with high enantioselectivity (up to 83% ee) [36]. Furthermore, Maruoka and coworkers reported efficient conjugate additions of α-heterosubstituted aldehydes with 11 using a sulfonamide organocatalyst with a dihydroanthracene framework (up to 95% ee) [37]. Recently, we also reported that a diaminomethylenemalononitrile organocatalyst catalyzes similar conjugate additions to afford 13a with high enantioselectivity (up to 89% ee) [38].
On the other hand, fluorous compounds with a perfluoroalkyl group can be easily separated from nonfluorous compounds by fluorous organic solvent extraction or fluorous solid phase extraction (FSPE) using fluorous silica gel [39]. Several research groups have reported asymmetric reactions in which fluorous organocatalysts are recyclable [40]. We have also reported a direct aldol reaction in water using fluorous sulfonamide organocatalyst 3 and related catalysts [41,42,43], Michael addition reactions using a fluorous thiourea organocatalyst [44], and an oxidation reaction using fluorous IBX [45]. In addition, we have reported a method for the synthesis of both enantiomeric aldol products in water using sulfonamide organocatalysts 1 [46] and 2 [47,48], prepared from l-phenylalanine. Very recently, we reported in a preliminary communication that perfluoroalkanesulfonamides 5 and 6 catalyze the conjugate additions of branched aldehydes to vinyl sulfone 11 to give the corresponding adducts with excellent stereoselectivities [49]; however, development of a protocol for recovery and reuse of 5 and 6 is yet to be reported. Herein, we describe the full details of the conjugate additions of branched aldehydes to vinyl sulfone using 6 and novel fluorous sulfonamide 7 (Figure 1).
Figure 1. Structure of organocatalysts.
Figure 1. Structure of organocatalysts.
Molecules 18 14529 g001

2. Results and Discussion

We initially examined the sulfonamide organocatalysts 17 for the conjugate addition of 12a to 11 as a test reactant (Table 1). Sulfonamide organocatalysts 14 derived from l-phenylalanine were superior to catalyst 5 derived from l-valine for the direct aldol reactions in water [37,46,48]; however, 5 bearing the valine skeleton resulted more suitable for the conjugate addition with vinyl sulfone 11 (entries 1–5). Furthermore, to develop a more powerful organocatalyst, we synthesized 6, which enhanced the acidity of the sulfonamide group by the introduction of the perfluorobutyl group. Treatment of compound 8 [50] with perfluorobutanesulfonyl fluoride in presence of triethylamine in dichloromethane provided the intermediate 9 in 79% yield. The Boc protective group was removed by treatment of hydrogen chloride in ethyl acetate to give the desired perfluorobutanesulfonamide 6 in 90% yield (Scheme 1). Organocatalyst 6 was more effective for conjugate additions with vinyl sulfone 11, resulting in the highest enantioselectivity (91% ee) and excellent yield (entry 6). Furthermore, to develop an organocatalyst that can be recovered and reused, 7 was synthesized by a similar procedure (Scheme 2). The stereoselectivity was slightly reduced in the reaction using 7 (entry 7).
Table 1. Selection of organocatalyst. Molecules 18 14529 i001
Table 1. Selection of organocatalyst. Molecules 18 14529 i001
EntryCatalystTime (h)Yield a (%)ee b (%)
11195−2
221.59980
3328886
4429779
5529588
66210091
7729989
a isolated yield; b determined by HPLC analysis.
Scheme 1. Preparation of organocatalyst 6.
Scheme 1. Preparation of organocatalyst 6.
Molecules 18 14529 g002
Scheme 2. Preparation of organocatalyst 7.
Scheme 2. Preparation of organocatalyst 7.
Molecules 18 14529 g003
We investigated the optimal reaction conditions for the enantioselective conjugate additions using 6, various solvents, and additives (Table 2). Conjugate additions were performed with vinyl sulfone 11 and 2-methylphenylethanal (12a) as test reactants in the presence of a catalytic amount of 6 and trifluoroacetic acid (TFA) at room temperature. A slight reduction in enantioselectivity and much longer reaction time were observed without TFA (entries 1 and 2). Aprotic solvents such as dichloromethane, diethyl ether, ethyl acetate, acetonitrile, chloroform, 1,2-dichloroethane, and p, m, and o-xylene were accepted well in this conjugate addition with good enantioselectivity (entries 3 and 5–12). A protic polar solvent such as methanol is a poor solvent for this reaction and provided low yield and enantioselectivity (entry 4). Among the solvents probed, the best results (95% yield and 93% ee) were achieved when the reaction was performed in m-xylene (entry 11). We also examined the effects associated with the presence of other protic acids, including benzoic acid, p-nitrobenzoic acid, and trifluoromethanesulfonic acid; however, TFA was found to be the most suitable additive (entries 13–15). Additions of 0.2 or 0.05 equiv of TFA resulted in a slight reduction in stereoselectivity (entries 16 and 17). The highest enantioselectivity (95% ee) was obtained when the reaction was performed at 0 °C or −10 °C, although longer reaction time (21 h or 72 h) was required (entries 18 and 19). Enantioselectivity was slightly reduced when the catalyst loading was lowered to 0.05 equiv (entry 20). Considering the reaction time, the optimal conditions were determined to be 0.1 equiv of 6 and 0.1 equiv of TFA in m-xylene at room temperature (entry 11).
Table 2. Optimization of reaction conditions using organocatalyst 6. Molecules 18 14529 i002
Table 2. Optimization of reaction conditions using organocatalyst 6. Molecules 18 14529 i002
EntrySolventTempAdditive (equiv)Time (h)Yield a (%)ee b (%)
1toluenertnone247586
2toluenertTFA (0.1)210091
3CH2Cl2rtTFA (0.1)2.59791
4MeOHrtTFA (0.1)243332
5Et2OrtTFA (0.1)58782
6EtOAcrtTFA (0.1)2.59886
7MeCNrtTFA (0.1)48777
8CHCl3rtTFA (0.1)29991
9ClCH2CH2ClrtTFA (0.1)29892
10p-xylenertTFA (0.1)29492
11m-xylenertTFA (0.1)29593
12o-xylenertTFA (0.1)29789
13m-xylenertPhCO2H (0.1)5.54680
14m-xylenert4-NO2C6H4CO2H (0.1)248582
15m-xylenertTfOH (0.1)242483
16m-xylenertTFA (0.2)29791
17m-xylenertTFA (0.05)29891
18m-xylene0 °CTFA (0.1)219995
19m-xylene−10 °CTFA (0.1)729495
20 cm-xylenertTFA (0.05)39990
21 dm-xylenertTFA (0.01)209689
a Isolated yield; b Determined by HPLC analysis; c Catalyst (0.05 equiv) was used; d Catalyst (0.01 equiv) was used.
In order to identify the scope and limitations of aldehyde substrates, we investigated substituent effects of the branched aromatic aldehydes on the conjugate additions (Table 3). A range of electron-withdrawing substituents such as bromo and fluoro moieties, and electron-donating substituents such as methyl and methoxy groups on the aromatic ring of branched aldehydes 12bg provided the corresponding adducts in excellent yields with good enantioselectivities (83%–92% ee) (entries 2–7). The additions of branched aldehydes possessing a naphthalene motif, 12h and 12i, to vinyl sulfone 11 proceeded smoothly in the presence of a catalytic amount of 6 to afford the corresponding adducts 13h and 13i in excellent yields with 92% ee, respectively (entries 8 and 9). Interestingly, 2-methoxy-2-phenylacetaldehyde (12j) was also applicable and gave the corresponding adduct 13j in high yield, albeit with reduced enantioselectivity (entry 10). In addition, 6 promoted the reaction of N-Boc α-aminophenylacetaldehyde (12k) with 11 to yield the corresponding adduct 13k in 68% yield with 60% ee (entry 11).
Table 3. Conjugate additions using organocatalyst 6. Molecules 18 14529 i003
Table 3. Conjugate additions using organocatalyst 6. Molecules 18 14529 i003
EntryAldehydeProductTime (h)Yield a (%)ee b (%)
1 Molecules 18 14529 i004 Molecules 18 14529 i00529593
2 Molecules 18 14529 i006 Molecules 18 14529 i00739789
3 Molecules 18 14529 i008 Molecules 18 14529 i00929591
4 Molecules 18 14529 i010 Molecules 18 14529 i01129992
5 Molecules 18 14529 i012 Molecules 18 14529 i01349992
6 Molecules 18 14529 i014 Molecules 18 14529 i015109883
7 Molecules 18 14529 i016 Molecules 18 14529 i01749991
8 Molecules 18 14529 i018 Molecules 18 14529 i01939992
9 Molecules 18 14529 i020 Molecules 18 14529 i02149792
10 Molecules 18 14529 i022 Molecules 18 14529 i023778868
11 Molecules 18 14529 i024 Molecules 18 14529 i02566860
a Isolated yield; b Determined by HPLC analysis.
Based on the optimal conditions for conjugate additions using 6, the reaction conditions were optimized for the enantioselective conjugate additions using 7 (Table 4). 1,2-Dichloroethane was the most suitable solvent among those examined in the presence of 0.1 equiv of TFA at room temperature. The reaction in 1,2-dichloroethane provided high yield and enantioselectivity (entry 8). It should be noted that 7 can promote the conjugate additions in brine because the perfluoroalkyl chain of 7 functions as the hydrophobic reaction field in water as described in our previous report [42,43].
Table 4. Optimization of reaction conditions using organocatalyst 7. Molecules 18 14529 i026
Table 4. Optimization of reaction conditions using organocatalyst 7. Molecules 18 14529 i026
EntrySolventAdditive (Equiv)Time (h)Yield a (%)ee b (%)
1toluenenone248583
2tolueneTFA (0.1)29989
3CH2Cl2TFA (0.1)29091
4hexaneTFA (0.1)29186
5Et2OTFA (0.1)28589
6brineTFA (0.1)246878
7CHCl3TFA (0.1)210091
8ClCH2CH2ClTFA (0.1)28792
9m-xyleneTFA (0.1)28491
a Isolated yield; b Determined by HPLC analysis.
The generality and substrate scope were probed for the optimal conditions (Table 5). The tendency of reactivities using 7 was quite similar to that using 6; however, aldehydes 12e, 12i, and 12j were poor substrates and gave low to moderate yields (entries 5, 9, and 10). Interestingly, the stereoselectivity with 12g was improved up to 94% ee (entry 7). In addition, the yield in the reaction with 12k was improved up to 100% yield (entry 11).
The recyclability of 7 was evaluated. After use of 7 in the conjugate addition of 12a to 11 under the optimal conditions, it was readily recovered by the FSPE technique using fluorous silica gel. Furthermore, the recovered catalyst 7 can be reused without further purification, and its catalytic activity was retained for the first reuse. Unfortunately, the catalytic activity of the recovered catalyst 7 decreased significantly for the second reuse.
We infer that the conjugate additions of aldehydes 12 to vinyl sulfone 11 using 6 or 7 proceed via a plausible transition state (Scheme 3) based on the stereochemistry of addition products 13ai. The primary amino group of 6 or 7 condenses with aldehydes 12 to generate the corresponding imine intermediate. The imine intermediate is subsequently isomerized to the E-enamine intermediate because of the resonance stabilizing effect of the aromatic ring. Then, the acidic proton of the sulfonamide group, which coordinates intramolecularly to nitrogen in the enamine transition state, successfully interacts with the oxygen of vinyl sulfone to control the approach direction of vinyl sulfone to the Re face of the enamine intermediate. This ultimately affords the corresponding addition products with high stereoselectivities. We believe that the acidity of 6 and 7 is enhanced by the powerful electron-withdrawing effect of the perfluoroalkyl chains, enabling strong coordination to vinyl sulfone and stabilizing the rigid transition states during conjugate additions. Moreover, the addition of TFA to the conjugate additions might accelerate the formation of the imine and enamine intermediates as well as reinforce the rigid transition state of the conjugate additions.
Table 5. Conjugate additions using organocatalyst 7. Molecules 18 14529 i027
Table 5. Conjugate additions using organocatalyst 7. Molecules 18 14529 i027
EntryAldehydeProductTime (h)Yield a (%)ee b (%)
1 Molecules 18 14529 i028 Molecules 18 14529 i02928792
2 Molecules 18 14529 i030 Molecules 18 14529 i03149090
3 Molecules 18 14529 i032 Molecules 18 14529 i033610092
4 Molecules 18 14529 i034 Molecules 18 14529 i03529282
5 Molecules 18 14529 i036 Molecules 18 14529 i03731380
6 Molecules 18 14529 i038 Molecules 18 14529 i039510083
7 Molecules 18 14529 i040 Molecules 18 14529 i04148194
8 Molecules 18 14529 i042 Molecules 18 14529 i04337692
9 Molecules 18 14529 i044 Molecules 18 14529 i045244589
10 Molecules 18 14529 i046 Molecules 18 14529 i047246468
11 Molecules 18 14529 i048 Molecules 18 14529 i049610064
a Isolated yield; b Determined by HPLC analysis.
Scheme 3. Plausible mechanism and transition state model of reaction.
Scheme 3. Plausible mechanism and transition state model of reaction.
Molecules 18 14529 g004

3. Experimental

3.1. General

1H-NMR and 13C-NMR spectra were measured with a JEOL AL 400 spectrometer (400 MHz for 1H-NMR and 100 MHz for 13C-NMR), or JEOL ECA-500 spectrometer (500 MHz for 1H-NMR and 125 MHz for 13C-NMR). The chemical shifts are expressed in ppm downfield from tetramethylsilane (δ = 0.00) as an internal standard. For thin layer chromatographic (TLC) analyses, Merck precoated TLC plates (silica gel 60 F254, Art 5715) were used. The products were isolated by flash column chromatography on silica gel (Kanto Chemical, Tokyo, Japan, silica gel 60N, spherical, neutral, 40–50 µm).

3.2. Preparation of Organocatalyst 6

(S)-tert-Butyl 3-methyl-1-(perfluorobutanesulfonamido)butan-2-ylcarbamate (9). To a solution of (S)-tert-butyl 1-amino-3-methylbutan-2-ylcarbamate (8, 300 mg, 1.48 mmol) [50] in dry CH2Cl2 (5 mL) was added triethylamine (0.46 mL, 3.06 mmol) at room temperature under an argon atmosphere. After stirring for 5 min, perfluorobutanesulfonyl fluoride (0.87 mL, 4.45 mmol) was added to the reaction mixture at 0 °C. After stirring for 1 h at 0 °C, the reaction mixture was additionally stirred for 45 h at room temperature. The reaction mixture was added to water and extracted three times with EtOAc. The EtOAc layers were combined, washed with brine, dried over anhydrous MgSO4, and evaporated. The residue was purified by flash column chromatography on silica gel with a 7:1 mixture of hexane and EtOAc to give the pure 9 (566 mg, 79%) as a colorless powder. Mp = 74–75 °C; Molecules 18 14529 i050 = −5.4° (c = 0.62 in MeOH); 1H-NMR (400 MHz, CD3OD): δ = 0.81 (d, J = 6.8 Hz, 3H), 0.85 (d, J = 6.8 Hz, 3H), 1.35 (s, 9H), 1.65–1.71 (m, 1H), 3.08 (dd, J = 8.1, 13.5 Hz, 1H), 3.28 (dd, J = 4.5, 13.5 Hz, 1H), 3.31–3.37 (m, 1H); 13C-NMR (125 MHz, CD3OD): δ = 18.2, 19.9, 28.8, 31.0, 47.1, 57.6, 80.2, 110.2–121.0 (complex signals of –CF2 and –CF3), 158.5; HRMS (ESI-TOF): calcd for C14H21F9N2O4SNa (M+Na)+: 507.0976, Found: 507.0991.
(S)-N-(2-Amino-3-methylbutyl)-perfluorobutanesulfonamide (6). To a solution of 9 (300 mg, 0.619 mmol) in EtOAc (2.5 mL) was added a 4 M solution of hydrochloric acid in EtOAc (2.5 mL) at 0 °C. After stirring for 2.5 h at room temperature, the reaction mixture was evaporated. The residue was added to saturated aqueous NaHCO3 and extracted three times with EtOAc. The EtOAc layers were combined, washed with brine, dried over anhydrous MgSO4, and evaporated. The residue was purified by flash column chromatography on silica gel with a 20:1 mixture of CHCl3 and MeOH to give the pure 6 (214 mg, 90%) as a colorless powder. Mp = 134–136 °C; Molecules 18 14529 i051 = +7.9° (c = 1.01 in MeOH); 1H-NMR (500 MHz, CD3OD): δ = 1.01 (d, J = 6.9 Hz, 3H), 1.02 (d, J = 6.9 Hz, 3H), 1.90–1.97 (m, 1H), 2.82–2.86 (m, 1H), 3.14 (dd, J = 8.6, 13.1 Hz, 1H), 3.41 (dd, J = 3.5, 13.1 Hz, 1H); 13C-NMR (125 MHz, CD3OD): δ = 18.9, 19.0, 30.1, 47.4, 60.7, 110.2–120.4 (complex signals of –CF2 and –CF3); Anal. Calcd for C9H13F9N2O2S: C, 28.13; H, 3.41; N, 7.29. Found: C, 28.07; H, 3.39; N, 7.26.

3.3. Preparation of Organocatalyst 7

(S)-tert-Butyl 3-methyl-1-(perfluorooctanesulfonamido)butan-2-ylcarbamate (10). To a solution of (S)-tert-butyl 1-amino-3-methylbutan-2-ylcarbamate (8) [50] (385 mg, 1.90 mmol) in dry CH2Cl2 (20 mL) was added triethylamine (0.80 mL, 5.71 mmol) at room temperature under an argon atmosphere. After stirring for 5 min, perfluorooctanesulfonyl fluoride (1.57 mL, 5.71 mmol) was added to the reaction mixture at 0 °C. After stirring for 2 h at 0 °C, the reaction mixture was additionally stirred for 90 h at room temperature. The reaction mixture was added to water and extracted three times with EtOAc. The EtOAc layers were combined, washed with brine, dried over anhydrous MgSO4, and evaporated. The residue was purified by flash column chromatography on silica gel with a 6:1 mixture of hexane and EtOAc to give the pure 10 (602 mg, 46%) as a pale yellow oil. Molecules 18 14529 i050 = −4.2° (c = 1.28 in MeOH); 1H-NMR (500 MHz, CDCl3): δ = 0.95 (d, J = 7.4 Hz, 3H), 0.97 (d, J = 6.8 Hz, 3H), 1.44 (s, 9H), 1.80–1.85 (m, 1H), 3.25 (m, 1H), 3.46 (brd, J = 12.6 Hz, 1H), 3.55 (m, 1H), 4.67 (brd, J = 8.0 Hz, 1H), 7.13 (brs, 1H); 13C-NMR (125 MHz, CDCl3): δ = 18.0, 19.2, 28.2, 30.1, 48.4, 55.7, 80.8, 108.0–113.0 (complex signals of –CF2 and –CF3), 157.6; HRMS (ESI-TOF): calcd for C18H21F17N2O4SNa (M+Na)+: 707.0848, Found: 707.0873.
(S)-N-(2-Amino-3-methylbutyl)-perfluorooctanesulfonamide (7). To a solution of 10 (570 mg, 0.833 mmol) in EtOAc (3.5 mL) was added a 4M solution of hydrochloric acid in EtOAc (3.5 mL) at 0 °C. After stirring for 2 h at room temperature, the reaction mixture was evaporated. The residue was added to saturated aqueous NaHCO3 and extracted three times with EtOAc. The EtOAc layers were combined, washed with brine, dried over anhydrous MgSO4, and evaporated. The residue was purified by flash column chromatography on silica gel with a 20:1 mixture of CHCl3 and MeOH to give the pure 7 (444 mg, 91%) as a colorless powder. Mp = 144–145 °C; Molecules 18 14529 i050 = + 6.9° (c = 1.01 in MeOH); 1H-NMR (500 MHz, CD3OD): δ = 1.01 (d, J = 6.8 Hz, 3H), 1.03 (d, J = 6.8 Hz, 3H), 1.90–1.97 (m, 1H), 2.82–2.86 (m, 1H), 3.15 (dd, J = 8.5, 13.1 Hz, 1H), 3.41 (dd, J = 4.0, 13.1 Hz, 1H); 13C-NMR (125 MHz, CD3OD): δ = 18.9, 19.0, 30.1, 47.5, 60.7, 109.7–121.5 (complex signals of –CF2 and –CF3); Anal. Calcd for C13H13F17N2O2S: C, 26.72; H, 2.24; N, 4.79. Found: C, 26.75; H, 2.41; N, 4.86.

3.4. Typical Procedure for Michael Addition (Table 3)

A typical procedure of the Michael additions using 6 is as follows: To a solution of 11 (30.8 mg, 0.100 mmol) and organocatalyst 6 (3.8 mg, 0.010 mmol) in m-xylene (1.0 mL) was added 2-phenylpropanal (26.8 µL, 0.200 mmol) and trifluororacetic acid (0.7 µL, 0.010 mmol) at room temperature. After stirring at room temperature for 2 h, the reaction mixture was directly purified by flash column chromatography on silica gel with a 3:1 mixture of hexane and EtOAc to afford the pure 13a (42.0 mg, 95%) as a colorless powder. All the Michael addition products 13 in the paper are known compounds that exhibited spectroscopic data identical to those reported in the literature [36,37].
(R)-2-Methyl-2-phenyl-4,4-bis(phenylsulfonyl)butanal (13a). Molecules 18 14529 i052 = −25.6° (c = 1.00, CHCl3); 95% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 21.7 min, tminor = 25.9 min.
(R)-2-(4-Bromophenyl)-2-methyl-4,4-bis(phenylsulfonyl)butanal (13b). Molecules 18 14529 i053 = −15.2° (c = 1.00, CHCl3); 89% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 27.1 min, tminor = 38.5 min.
(R)-2-(4-Fluorophenyl)-2-methyl-4,4-bis(phenylsulfonyl)butanal (13c). Molecules 18 14529 i052 = +23.5° (c = 1.00, CHCl3); 91% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 25.5 min, tminor = 32.1 min.
(R)-2-Methyl-4,4-bis(phenylsulfonyl)-2-p-tolylbutanal (13d). Molecules 18 14529 i052 = +25.4° (c = 1.00, CHCl3); 92% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 21.3 min, tminor = 29.7 min.
(R)-2-(3-Methoxyphenyl)-2-methyl-4,4-bis(phenylsulfonyl)butanal (13e). Molecules 18 14529 i054 = +10.6° (c = 1.00, CHCl3); 92% ee; enantiomeric excess was determined by HPLC with Chiralcel AD-H column (hexane/2-propanol = 80:20), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 27.4 min, tminor = 38.6 min.
(R)-2-(2-Methoxyphenyl)-2-methyl-4,4-bis(phenylsulfonyl)butanal (13f). Molecules 18 14529 i052 = −70.8° (c = 1.00, CHCl3); 83% ee; enantiomeric excess was determined by HPLC with Chiralcel AD-H column (hexane/2-propanol = 80:20), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 18.8 min, tminor = 25.9 min.
(R)-2-(3-Bromophenyl)-2-methyl-4,4-bis(phenylsulfonyl)butanal (13g). Molecules 18 14529 i052 = −71.7° (c = 1.00, CHCl3); 91% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 80:20), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 41.3 min, tminor = 47.1 min.
(R)-2-Methyl-2-(naphthalen-2-yl)-4,4-bis(phenylsulfonyl)butanal (13h). Molecules 18 14529 i052 = +13.0° (c = 1.00, CHCl3); 92% ee; enantiomeric excess was determined by HPLC with Chiralpak AS-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 32.9 min, tminor = 39.1 min.
(R)-2-Methyl-2-(naphthalen-1-yl)-4,4-bis(phenylsulfonyl)butanal (13i). Molecules 18 14529 i053 = −31.8° (c = 1.00, CHCl3); 92% ee; enantiomeric excess was determined by HPLC with Chiralcel AD-H column (hexane/2-propanol = 70:30), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 16.7 min, tminor = 22.4 min.
(R)-2-Methoxy-2-phenyl-4,4-bis(phenylsulfonyl)butanal (13j). 68% ee; enantiomeric excess was determined by HPLC with Chiralcel AD-H column (hexane/2-propanol = 5:1), flow rate = 1.0 mL/min; λ = 220 nm; tmajor = 37.1 min, tminor = 44.5 min.
(R)-tert-Butyl 1-oxo-2-phenyl-4,4-bis(phenylsulfonyl)butan-2-ylcarbamate (13k). Molecules 18 14529 i055 = +10.0° (c = 1.00, CHCl3) 64% ee; enantiomeric excess was determined by HPLC with Chiralcel OD-H column (hexane/2-propanol = 90:10), flow rate = 0.5 mL/min; λ = 220 nm; tmajor = 19.9 min, tminor = 22.2 min.

4. Conclusions

Novel organocatalysts 6 and 7 can easily be prepared from l-valine, an inexpensive and commercially available natural amino acid. Organocatalysts 6 and 7, which are simple β-aminosulfonamides with only one stereogenic center, efficiently catalyze the conjugate additions of various branched aldehydes to vinyl sulfone 11 with a short reaction time at room temperature to give the corresponding addition products possessing all-carbon quaternary stereocenters with high enantioselectivities. The excellent performance is probably due to the carbon skeleton of l-valine and the electron-withdrawing effect of the perfluoroalkyl groups on 6 and 7. Moreover, fluorous organocatalyst 7 bearing a perfluorooctyl group was readily recovered by simple solid phase extraction using fluorous silica gel and was immediately reusable without further purification for the first cycle. Further application of these organocatalysts in the synthesis of bioactive compounds is currently being investigated in our laboratory.

Acknowledgments

This work was supported in part by Grants-in-Aid for Scientific Research (C) (No. 22590007) from the Japan Society for the Promotion of Science.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fuji, K. Asymmetric creation of quaternary carbon centers. Chem. Rev. 1993, 93, 2037–2066. [Google Scholar] [CrossRef]
  2. Corey, E.J.; Guzman-Perez, A. The catalytic enantioselective construction of molecules with quaternary carbon stereocenters. Angew. Chem. Int. Ed. 1998, 37, 388–401. [Google Scholar] [CrossRef]
  3. Christoffers, J.; Mann, A. Enantioselective construction of quaternary stereocenters. Angew. Chem. Int. Ed. 2001, 40, 4591–4597. [Google Scholar] [CrossRef]
  4. Christoffers, J.; Baro, A. Construction of quaternary stereocenters: New perspectives through enantioselective Michael reactions. Angew. Chem. Int. Ed. 2003, 42, 1688–1690. [Google Scholar] [CrossRef]
  5. Douglas, C.J.; Overman, L.E. Catalytic asymmetric synthesis of all-carbon quaternary stereocenters. Proc. Natl. Acad. Sci. USA 2004, 101, 5363–5367. [Google Scholar] [CrossRef]
  6. Christoffers, J.; Baro, A. Stereoselective construction of quaternary stereocenters. Adv. Synth. Catal. 2005, 347, 1473–1482. [Google Scholar] [CrossRef]
  7. Trost, B.M.; Jiang, C. Catalytic enantioselective construction of all-carbon quaternary stereocenters. Synthesis 2006, 2006, 369–396. [Google Scholar] [CrossRef]
  8. Cozzi, P.G.; Hilgraf, R.; Zimmermann, N. Enantioselective catalytic formation of quaternary stereogenic centers. Eur. J. Org. Chem. 2007, 2007, 5969–5994. [Google Scholar]
  9. Marek, I.; Sklute, G. Creation of quaternary stereocenters in carbonyl allylation reactions. Chem. Commun. 2007, 2007, 1683–1691. [Google Scholar] [CrossRef]
  10. Riant, O.; Hannedouche, J. Asymmetric catalysis for the construction of quaternary carbon centres: nucleophilic addition on ketones and ketimines. Org. Biomol. Chem. 2007, 5, 873–888. [Google Scholar] [CrossRef]
  11. Hawner, C.; Alexakis, A. Metal-catalyzed asymmetric conjugate addition reaction: Formation of quaternary stereocenters. Chem. Commun. (Camb) 2010, 46, 7295–7306. [Google Scholar] [CrossRef]
  12. Shimizu, M. Construction of asymmetric quaternary carbon centers with high enantioselectivity. Angew. Chem. Int. Ed. 2011, 50, 5998–6000. [Google Scholar] [CrossRef]
  13. Das, J.P.; Marek, I. Enantioselective synthesis of all-carbon quaternary stereogenic centers in acyclic systems. Chem. Commun. (Camb) 2011, 47, 4593–4623. [Google Scholar] [CrossRef]
  14. Bella, M.; Gasperi, T. Organocatalytic formation of quaternary stereocenters. Synthesis 2009, 2009, 1583–1614. [Google Scholar] [CrossRef]
  15. Figueiredo, R.M.; Christmann, M. Organocatalytic synthesis of drugs and bioactive natural products. Eur. J. Org. Chem. 2007, 2007, 2575–2600. [Google Scholar] [CrossRef]
  16. Mossé, S.; Alexakis, A. First organocatalyzed asymmetric Michael addition of aldehydes to Vinyl sulfones. Org. Lett. 2005, 7, 4361–4364. [Google Scholar] [CrossRef]
  17. Liu, T.-Y.; Long, J.; Li, B.-J.; Jiang, L.; Li, R.; Wu, Y.; Ding, L.-S.; Chen, Y.-C. Enantioselective construction of quaternary carbon centre catalysed by bifunctional organocatalyst. Org. Biomol. Chem. 2006, 4, 2097–2099. [Google Scholar] [CrossRef]
  18. Zhu, Q.; Lu, Y. Organocatalytic michael addition of aldehydes to vinyl sulfones: Enantioselective α-alkylations of aldehydes and their derivatives. Org. Lett. 2008, 10, 4803–4806. [Google Scholar] [CrossRef]
  19. Zhu, Q.; Cheng, L.; Lu, Y. Asymmetric organocatalytic Michael addition of ketones to vinyl sulfone. Chem. Commun. 2008, 2008, 6315–6317. [Google Scholar]
  20. Quintard, A.; Bournaud, C.; Alexakis, A. Diversity-oriented synthesis towards conceptually new highly modular aminal–pyrrolidine organocatalysts. Chem. Eur. J. 2008, 14, 7504–7507. [Google Scholar] [CrossRef]
  21. Zhu, Q.; Lu, Y. Enantioselective conjugate addition of nitroalkanes to vinyl sulfone: An organocatalytic access to chiral amines. Org. Lett. 2009, 11, 1721–1724. [Google Scholar] [CrossRef]
  22. Landa, A.; Maestro, M.; Masdeu, C.; Puente, A.; Vera, S.; Oiarbide, M.; Palomo, C. Highly enantioselective conjugate additions of aldehydes to vinyl sulfones. Chem. Eur. J. 2009, 15, 1562–1565. [Google Scholar] [CrossRef]
  23. Quintard, A.; Alexakis, A. Organocatalytic addition on 1,2-bis(sulfone)vinylenes leading to an unprecedented rearrangement. Chem. Eur. J. 2009, 15, 11109–11113. [Google Scholar] [CrossRef]
  24. Xiao, J.; Liu, Y.-L.; Loh, T.-P. The organocatalytic enantioselective Michael addition of aldehydes to vinyl sulfones in water. Synlett 2010, 2010, 2029–2032. [Google Scholar]
  25. Quintard, A.; Alexakis, A. Asymmetric addition of α-hetero-disubstituted aldehydes to vinyl sulfones; formation of highly functionalized tetrasubstituted carbon centres. Chem. Commun. (Camb) 2010, 46, 4085–4087. [Google Scholar] [CrossRef]
  26. Alba, A.-N.R.; Conpanyó, X.; Valero, G.; Moyano, A.; Rios, R. Enantioselective organocatalytic addition of oxazolones to 1,1-bis(phenylsulfonyl)ethylene: A convenient asymmetric synthesis of quaternary α-amino acids. Chem. Eur. J. 2010, 16, 5354–5361. [Google Scholar] [CrossRef]
  27. Chua, P.J.; Tan, B.; Yang, L.; Zeng, X.; Zhu, D.; Zhong, G. Highly stereoselective synthesis of indanes with four stereogenic centers via sequential Michael reaction and [3+2] cycloaddition. Chem. Commun. 2010, 46, 7611–7613. [Google Scholar] [CrossRef]
  28. Zhu, Q.; Lu, Y. Stereocontrolled creation of all-carbon quaternary stereocenters by organocatalytic conjugate addition of oxindoles to vinyl sulfone. Angew. Chem. Int. Ed. 2010, 49, 7753–7756. [Google Scholar] [CrossRef]
  29. Quintard, A.; Alexakis, A. 1,2-Sulfone rearrangement in organocatalytic reactions. Org. Biomol. Chem. 2011, 9, 1407–1418. [Google Scholar] [CrossRef]
  30. Quintard, A.; Alexakis, A. Highly enantioselective organocascade intermolecular iminium/enamine Michael addition on enals. Chem. Commun. (Camb) 2011, 47, 7212–7214. [Google Scholar] [CrossRef]
  31. Xiao, J.; Lu, Y.-P.; Liu, Y.-L.; Wong, P.-S.; Loh, T.-P. A New class of structurally rigid tricyclic chiral secondary amine organocatalyst: Highly enantioselective organocatalytic michael addition of aldehydes to vinyl sulfones. Org. Lett. 2011, 13, 876–879. [Google Scholar] [CrossRef]
  32. Kang, J.Y.; Carter, R.G. Primary Amine, Thiourea-based dual catalysis motif for synthesis of stereogenic, all-carbon quaternary center-containing cycloalkanones. Org. Lett. 2012, 14, 3178–3181. [Google Scholar] [CrossRef]
  33. Dou, X.; Lu, Y. Enantioselective Conjugate addition of 3-fluorooxindoles to vinyl sulfone: An organocatalytic access to chiral 3-fluoro-3-substituted oxindoles. Org. Biomol. Chem. 2013, 11, 5217–5221. [Google Scholar] [CrossRef]
  34. Mossé, S.; Alexakis, A.; Mareda, J.; Bollot, G.; Bernardinelli, G.; Filinchuk, Y. Enantioselective organocatalytic conjugate addition of aldehydes to vinyl sulfones and vinyl phosphonates as challenging michael acceptors. Chem. Eur. J. 2009, 15, 3204–3220. [Google Scholar] [CrossRef]
  35. Quintard, A.; Belot, S.; Marchal, E.; Alexakis, A. Aminal–pyrrolidine organocatalysts—highly efficient and modular catalysts for α-functionalization of carbonyl compounds. Eur. J. Org. Chem. 2010, 2010, 927–936. [Google Scholar] [CrossRef]
  36. Zhu, Q.; Lu, Y. Chiral primary amine mediated conjugate addition of branched aldehydes to vinyl sulfone: Asymmetric generation of quaternary carbon centers. Chem. Commun. (Camb) 2010, 46, 2235–2237. [Google Scholar] [CrossRef]
  37. Moteki, S.A.; Xu, S.; Arimitsu, S.; Maruoka, K. Design of structurally rigid trans-diamine-based tf-amide organocatalysts with a dihydroanthracene framework for asymmetric conjugate additions of heterosubstituted aldehydes to vinyl sulfones. J. Am. Chem. Soc. 2010, 132, 17074–17076. [Google Scholar] [CrossRef]
  38. Kanada, Y.; Yuasa, H.; Nakashima, K.; Murahashi, M.; Tada, N.; Itoh, A.; Koseki, Y.; Miura, T. Asymmetric conjugate addition of aldehydes to vinyl sulfone using a diaminomethylenemalononitrile organocatalyst. Tetrahedron Lett. 2013, 54, 4896–4899. [Google Scholar] [CrossRef]
  39. The Handbook of Fluorous Chemistry; Gladysz, J.A.; Horváth, I.T.; Curran, D.P. (Eds.) Wiley-VCH: Weinheim, Germany, 2004.
  40. Zhang, W.; Cai, C. New chemical and biological applications of fluorous technologies. Chem. Commun. 2008, 2008, 5686–5694. [Google Scholar] [CrossRef]
  41. Miura, T.; Imai, K.; Ina, M.; Tada, N.; Imai, N.; Itoh, A. Direct asymmetric aldol reaction with recyclable fluorous organocatalyst. Org. Lett. 2010, 12, 1620–1623. [Google Scholar] [CrossRef]
  42. Miura, T.; Imai, K.; Kasuga, H.; Ina, M.; Tada, N.; Imai, N.; Itoh, A. Direct asymmetric aldol reactions in brine with recyclable fluorous β-aminosulfonamide organocatalysts. Tetrahedron 2011, 67, 6340–6346. [Google Scholar] [CrossRef]
  43. Miura, T.; Imai, K.; Kasuga, H.; Ina, M.; Tada, N.; Imai, N.; Itoh, A. Highly efficient asymmetric aldol reaction in brine using fluorous sulfonamide organocatalyst. Org. Biomol. Chem. 2012, 10, 2209–2213. [Google Scholar] [CrossRef]
  44. Miura, T.; Nishida, S.; Masuda, A.; Tada, N.; Itoh, A. Asymmetric Michael additions of aldehydes to maleimides using a recyclable fluorous thiourea organocatalyst. Tetrahedron Lett. 2011, 52, 4158–4160. [Google Scholar] [CrossRef]
  45. Miura, T.; Nakashima, K.; Tada, N.; Itoh, A. An Effective and catalytic oxidation using recyclable fluorous IBX. Chem. Commun. 2011, 47, 1875–1877. [Google Scholar] [CrossRef]
  46. Miura, T.; Ina, M.; Imai, K.; Nakashima, K.; Masuda, A.; Imai, N.; Tada, N.; Itoh, A. β-Aminosulfonamide-catalyzed direct asymmetric Aldol reaction in brine. Synlett 2011, 2011, 410–414. [Google Scholar]
  47. Miura, T.; Yasaku, Y.; Koyata, N.; Murakami, Y.; Imai, N. Direct asymmetric aldol reactions in brine using novel sulfonamide catalyst. Tetrahedron Lett. 2009, 50, 2632–2635. [Google Scholar] [CrossRef]
  48. Miura, T.; Ina, M.; Imai, K.; Nakashima, K.; Yasaku, Y.; Koyata, N.; Murakami, Y.; Imai, N.; Tada, N.; Itoh, A. Direct asymmetric aldol reactions in water with a β-aminosulfonamide organocatalyst. Tetrahedron Asymmetry 2011, 22, 1028–1034. [Google Scholar] [CrossRef]
  49. Miura, T.; Yuasa, H.; Murahashi, M.; Ina, M.; Nakashima, K.; Tada, N.; Itoh, A. Highly efficient asymmetric conjugate additions of aldehydes with vinyl sulfones using sulfonamide organocatalyst. Synlett 2012, 2012, 2385–2388. [Google Scholar]
  50. Imai, N.; Nokami, J.; Nomura, T.; Ninomiya, Y.; Shinobe, A.; Matsushiro, S. Convenient preparation of various chiral disulfonamides from α-amino acid and alkylation of cinnamaldehyde in the presence of the chiral disulfonamides. Bull. Okayama Univ. Sci. 2002, 47–52. [Google Scholar]
  • Sample Availability: Samples of the compounds are not available from the authors.

Share and Cite

MDPI and ACS Style

Nakashima, K.; Murahashi, M.; Yuasa, H.; Ina, M.; Tada, N.; Itoh, A.; Hirashima, S.-i.; Koseki, Y.; Miura, T. Perfluoroalkanesulfonamide Organocatalysts for Asymmetric Conjugate Additions of Branched Aldehydes to Vinyl Sulfones. Molecules 2013, 18, 14529-14542. https://doi.org/10.3390/molecules181214529

AMA Style

Nakashima K, Murahashi M, Yuasa H, Ina M, Tada N, Itoh A, Hirashima S-i, Koseki Y, Miura T. Perfluoroalkanesulfonamide Organocatalysts for Asymmetric Conjugate Additions of Branched Aldehydes to Vinyl Sulfones. Molecules. 2013; 18(12):14529-14542. https://doi.org/10.3390/molecules181214529

Chicago/Turabian Style

Nakashima, Kosuke, Miho Murahashi, Hiroki Yuasa, Mariko Ina, Norihiro Tada, Akichika Itoh, Shin-ichi Hirashima, Yuji Koseki, and Tsuyoshi Miura. 2013. "Perfluoroalkanesulfonamide Organocatalysts for Asymmetric Conjugate Additions of Branched Aldehydes to Vinyl Sulfones" Molecules 18, no. 12: 14529-14542. https://doi.org/10.3390/molecules181214529

Article Metrics

Back to TopTop