Next Article in Journal
Emerging Management Approach for the Adverse Events of Immunotherapy of Cancer
Next Article in Special Issue
Indane-1,3-Dione: From Synthetic Strategies to Applications
Previous Article in Journal
Magnetic Self-Healing Composites: Synthesis and Applications
Previous Article in Special Issue
Synthesis of Functionalized Six-Membered-Ring Azahelicenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Asymmetric Michael Addition in Synthesis of β-Substituted GABA Derivatives

1
Jiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, College of Chemical Engineering, Nanjing Forestry University, Nanjing 210037, China
2
Departamento de Química Orgánica, Universidad de Valencia, 46100 Burjassot, Spain
3
Department of Organic Chemistry I, Faculty of Chemistry, University of the Basque Country UPV/EHU, Paseo Manuel Lardizábal 3, 20018 San Sebastián, Spain
4
IKERBASQUE, Basque Foundation for Science, Alameda Urquijo 36-5, Plaza Bizkaia, 48011 Bilbao, Spain
5
V.P. Kukhar Institute of Bioorganic Chemistry and Petrochemistry, The National Academy of Sciences of Ukraine, 1 Murmanska Str., 02094 Kyiv, Ukraine
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(12), 3797; https://doi.org/10.3390/molecules27123797
Submission received: 26 May 2022 / Revised: 7 June 2022 / Accepted: 7 June 2022 / Published: 13 June 2022
(This article belongs to the Special Issue Featured Reviews in Organic Chemistry)

Abstract

:
γ-Aminobutyric acid (GABA) represents one of the most prolific structural units widely used in the design of modern pharmaceuticals. For example, β-substituted GABA derivatives are found in numerous neurological drugs, such as baclofen, phenibut, tolibut, pregabalin, phenylpiracetam, brivaracetam, and rolipram, to mention just a few. In this review, we critically discuss the literature data reported on the preparation of substituted GABA derivatives using the Michael addition reaction as a key synthetic transformation. Special attention is paid to asymmetric methods featuring synthetically useful stereochemical outcomes and operational simplicity.

Graphical Abstract

1. Introduction

Tailor-made amino acids [1] play an indispensable role in the development of modern pharmaceuticals and drug formulations [2,3,4,5,6]. Thus, over 20% of newly approved small-molecule drugs contain structural fragments of AAs [7,8,9]. In particular, γ-aminobutyric-acid (GABA) derivatives bearing β-alkyl or β-aryl substituents, which include baclofen [10], phenibut [11,12,13], tolibut [14], and pregabalin [15,16,17], are recognized as an essential class of marketed pharmaceuticals for the treatment of neurological diseases (Figure 1). Similarly, piracetam-based GABA derivatives such as phenylpiracetam [18], brivaracetam [19,20,21], and rolipram [22] are also developed as pharmaceuticals. Introduction of β-alkyl or β-aryl substituent in the GABA backbone allows for the improvement of the lipophilic character of these compounds.
Significantly, biological activity of β-substituted GABA derivatives depends on their absolute configuration. For example, (R)-enantiomers of Baclofen (antispastic agent and muscle relaxant) and Phenibut (tranquilizer and anticonvulsant) are considerably more active than the corresponding (S)-enantiomers, while anticonvulsant activity of Pregabalin (anti-epilepsy drug) is primarily related to (S)-enantiomer [23]. Consequently, considerable efforts were devoted to developing asymmetric synthesis of β-substituted GABA derivatives, including chemical and biocatalytic resolution, asymmetric reduction, desymmetrization, aldol addition, and nucleophilic substitution [24,25,26]. For the past decade, the asymmetric Michael addition of carbon nucleophile to α,β-unsaturated compound-bearing nitro or carbonyl groups gained impressive progress, providing straightforward access to γ-nitrocarbonyl compounds as key chiral intermediates that can be converted into β-substituted GABA derivatives via subsequent transformation of functional groups. This review focuses on the application of the asymmetric Michael addition performed with the use of chiral auxiliaries as well as in the presence of chiral catalysts, especially chiral metal-free organocatalysts, in synthesis of β-substituted GABA derivatives.

2. Michael Addition of Carbonyl Compounds to α,β-Unsaturated Nitroalkenes

The proline-catalyzed asymmetric Michael addition reaction [27] of acetaldehyde with α,β-unsaturated nitroalkenes attracted considerable attention as synthetically useful routes to β-substituted derivatives of GABA. Thus, the addition of acetaldehyde to the nitroolefins (E)-1 and (E)-2 (Scheme 1) was carried out in the presence of enantiomerically pure (S)-diphenylprolinol silyl ether 3 as the catalyst [28,29]. After optimization of reaction conditions, the asymmetric Michael addition proceeded efficiently with 10–20 mol% of organocatalyst (S)-3 in such solvents as MeCN, DMF/i-PrOH, and 1,4-dioxane to afford γ-nitro aldehydes (S)-4 and (R)-5 in reasonable yield and excellent enantiomeric excess. Oxidation of γ-nitro aldehydes (S)-4 and (R)-5 was successfully performed in aqueous t-BuOH using NaClO2 and NaH2PO4 with 2-methyl-2-butene as a chlorine scavenger to afford carboxylic acids (S)-6 and (R)-7 in good to excellent yields [30]. The reduction of the nitro acid (S)-6 with Raney Ni in MeOH gave, after treatment with aqueous HCl, (S)-Baclofen 8 as hydrochloride salt in 91% yield. (R)-Pregabalin 9 was also synthesized by the reduction of the nitro group in (R)-7 under Pd/C in 93% yield. The mechanism of the asymmetric Michael addition reaction of acetaldehyde with nitroalkenes promoted by diphenylprolinol silyl ether involves the formation of the enamine as a nucleophile. Thus, the organocatalyst would react with the acetaldehyde forming anti-enamine with the double bond oriented away from the diphenylsiloxymethyl group. In this case, the (diphenylmethyl)trimethylsiloxy group provides the formation of anti-enamine and shielding one face of the enamine double bond. The anti-enamine would add stereoselectively to the nitroolefin via the acyclic synclinal transition state proposed by Seebach [31] as shown in Scheme 1.
In a similar way, the nitro olefin (E)-10 (Scheme 2), easily available from isovanillin via O-alkylation and Henry condensation, reacted with acetaldehyde in the presence of a catalytic amount of diphenylprolinol silyl ether (R)-3 (10 mol%) affording corresponding nitro aldehyde adduct which upon oxidation with oxone and esterification successfully transformed into ester derivative (R)-11 in 85% yield [32]. The nitroester (R)-11 underwent intramolecular reductive lactamization under H2 in presence of catalytic amount of Pd/C to furnish (R)-Rolipram 12 in 93% yield and >99% ee. The present method was also utilized to prepare enantiomerically pure (S)-Rolipram using (S)-diphenylprolinol silyl ether-mediated asymmetric Michael addition reaction as the key step. Recently, water-soluble diarylprolinol silyl ether containing the dimethylamine functionality was found to be very effective for the Michael additions of the acetaldehyde with nitroolefins. These reactions took place in brine with good yields and high enantioselectivities for a broad range of nitroolefins [33,34,35].
Enantioselective conjugate addition of malonates and their equivalents to nitroolefins promoted by bifunctional organocatalysts bearing a hydrogen-bonding donor group and Lewis base (tertiary amine) is considered as one of the most simple and efficient routes for constructing chiral β-substituted derivatives of GABA and their lactam analogs, especially from the perspective of green chemistry. The success of bifunctional organocatalysts was based on their ability to increase the reactivity of both nitroolefins and nucleophiles as well as control the approach of nucleophiles to nitroolefins in the transition state [36,37]. For example, the Michael reaction of nitroalkene (E)-1 (Scheme 3) with diethyl malonate was performed in the presence of Takemoto thiourea catalyst (R,R)-13 bearing 3,5-bis(trifluoromethyl)benzene and tertiary amino group [38]. The use of 2 equiv of diethyl malonate in toluene and 10 mol% catalyst loading was required to provide the Michael adduct (R)-14 in 80% yield with 94% ee. A single recrystallization made it possible to increase the enantiomeric purity of the product (R)-14 to 99% ee. Then, reductive cyclization of (R)-14 with NaBH4/NiCl2 in methanol gave γ-lactam (3S,4R)-15 as thermodynamically more stable anti-diastereomer [39,40], which after hydrolysis and decarboxylation was converted into lactam (R)-16 in 84%. Finally, acidic hydrolysis of (R)-16 gave (R)-Baclofen 8 as hydrochloride salt in 94% yield. A high level of enantioselectivity in organocatalytic Michael addition reaction was achieved as a result of dual activation through deprotonation of the acidic proton of diethylmalonate by tertiary amino group of the organocatalyst and the hydrogen-bond formation between the nitro group and the thiourea moiety [41].
Under solvent-free conditions reducing the amount of diethyl malonate from 2 to 1 equiv did not significantly influence both the yield and enantioselectivity of Michael addition to nitroalkenes promoted by thiourea (R,R)-13 catalysts. Thus, Michael addition of diethyl malonate to alkyl-substituted nitroalkene (E)-2 (Scheme 4) in the presence of 10 mol% of (R,R)-13 under solvent-free conditions produced the nitro ester (S)-17 in 73% yield and 88% enantiomeric excess [42]. Hydrogenation of (S)-17 over Raney Ni provided the pyrrolidin-2-one (3S,4S)-18 [43] in 72% after crystallization, which was subjected to ester hydrolysis followed by decarboxylation to give γ-lactam (S)-19 in 90% yield and 98% enantiomeric excess. Hydrolysis of γ-lactam (S)-19 with 6N HCl gave the enantiomerically pure (S)-Pregabalin 9 hydrochloride in 95% yield. Additionally, treatment of pyrrolidin-2-one derivative (3S,4S)-18 with 6N HCl at reflux also directly produced the (S)-Pregabalin 9 hydrochloride in 92% yield.
A series of Takamoto-type bifunctional organocatalysts were examined for the asymmetric Michael addition of malonate derivatives to nitroalkenes and showed very effective catalytic activity. For example, L-proline-derived bifunctional urea-pyrrolidine organocatalyst (S,R)-21 (Scheme 5) was demonstrated to catalyze the enantioselective Michael addition of diphenyl dithiomalonates to nitrostyrene (E)-20 in toluene as the best solvent at 25 °C [44]. The reaction performing with 5 mol% of (S,R)-21 was complete in 1.5 h, affording Michael adduct (R)-22 in 90% enantiomeric excess, which was improved to 98% after recrystallization from ethanol. Reduction of the nitro group using zinc in acetic acid and substoichiometric amounts of TiCl3 followed by intermolecular cyclization enabled the formation of the lactam (R)-23 in excellent yields of 90%. Hydrolysis of the lactam (R)-23 was finally achieved with 6N HCl and the resulting (R)-Phenibut 24 was isolated as hydrochloride in 85% yield. Additionally, mild basic hydrolysis-decarboxylation of Michael adduct (R)-22 provided γ-nitrothioester (R)-25 in 94% yield, which reduced and cyclized under the above described conditions (Zn/AcOH/TiCl3) to lactam (R)-26 that was isolated in 82% yield.
When the addition of cyclohexyl Meldrum’s acid to aliphatic nitroalkene (E)-2 (Scheme 6) was carried out with only 0.2 mol% of N-sulfinyl urea catalysis (Ss,R,R)-27 in cyclopentyl methyl ether (CPME), complete conversion of starting nitroalkene (E)-2 proceeded at 35 °C for 48 h providing the Michael adduct (S)-28 in 92% ee [45]. Using n-sulfinyl urea catalysis (Ss,R,R)-27 bearing cyclic tertiary amine was found to be essential for achieving high conversion and enantioselectivity. Direct hydrolysis/decarboxylation of addition product (S)-28 without purification led to one mole scale synthesis of γ-nitroacid (S)-29 in 90% overall yield from nitroalkene (E)-2. (S)-Pregabalin 9 could be provided by heterogeneous catalytic hydrogenation of (S)-29.
Another urea catalyst (S,S)-31 (Scheme 7) containing the pyrrolidine moiety was designed to catalyze in brine media [46] the enantioselective addition of dithiomalonate to otherwise unreactive β-CF3-β,β-disubstituted nitroalkenes 30 providing corresponding Michael adducts 32 with a β-trifluoromethylated quaternary stereocenter [47]. In brine media the use of 15 mol% of urea catalyst (S,S)-31 as well as the addition of o-xylene as additive was required for obtaining high yield and enantioselectivity of adducts 32 at 0 °C. In the absence of cosolvent, the achieved enantioselectivity of Michael adducts 32 was lower. These results could be explained by enforced hydrophobic interactions between catalysts and substrates when the reaction was carried out in brine media. Under optimal conditions, trifluoromethylated nitroalkenes 30 having substituted phenyl and iso-butyl groups were converted into the corresponding products 32 in 55–97% yields with 67–96% ee. So-obtained Michael products 32 were further subjected to reductive cyclization to furnish γ-lactam thioesters 33. Hydrolysis of γ-lactam thioesters 33 followed by a decarboxylation provided β-trifluoromethylated analogues of rolipram 34c, phenibut 35a, baclofen 35b, and pregabalin 35d. The enantiopurity of γ-lactam thioester 33d and γ-lactams 34a,b could be improved by recrystallization.
A three-component reaction of aromatic or aliphatic aldehydes, nitromethane and dimethyl malonate was found to catalyze by mesoporous siliceous material 36 (Scheme 8) incorporating urea-modified quinidine and propylamine groups in o-xylene at 70 °C leading to corresponding Michael adducts in reasonable yield [48,49]. The organic–inorganic hybrid catalyst 36 enabled access to Michael adducts with modest-to-good enantiomeric excess (50–70%). After removing the excess nitromethane by distillation, the o-xylene solution of Michael adducts was subjected to heterogeneous catalytic hydrogenation of the nitro group followed by cyclization and thermal decarboxylation to give γ-lactams with retention of enantiopurity. Thus, after column chromatography and recrystallization, (R)-Rolipram 12 and precursors of Phenibut (R)-26, Baclofen (R)-16 and Pregabalin (S)-19 were obtained in two-pot multicomponent operations with enantiopurity ranging from 81% ee to 97% ee. The solid hybrid catalyst could be reused three times without loss of activity.
Conformationally rigid cyclobutene-ring-derivative squaramide (R,S)-37 (Scheme 9) was shown to function in a complementary manner to thioureas as an effective hydrogen-bonding bifunctional organocatalyst for the Michael addition of dimethyl malonate to aromatic nitroalkene (E)-1 [50,51]. It was supposed that two H-bonds formed between the squaramide (R,S)-37 organocatalyst, incorporating tertiary nitrogen as Lewis base and chiral scaffold, and the nitroalkene substrate. Performing the reaction in dichloromethane as solvent with 5 mol% catalyst loading afforded the Michael product (R)-38 in 84% yield and 86% ee, albeit in a long reaction time. Polymer-immobilized squaramide was also effective for the Michael addition of dimethyl malonate to nitroalkene (E)-1, but significant loss of activity was observed in the second and third cycles. The adduct (R)-38 was further transformed to (R)-baclofen 8 hydrochloride in 56% overall yield through the procedure exemplified in Scheme 1.
Squaramide (R,R)-39 (Scheme 10), which possess an piperidine unit, was useful for the activation of Meldrum’s acid in enantioselective Michael addition to aliphatic nitroalkenes. The reaction of Meldrum’s acid and nitroalkene (E)-2 was carried out with catalyst loading of 5 mol% in dichloromethane as the best solvent to obtain the nitro derivative (S)-40 in 83% yield and 94% enantiomeric excess [52]. Squaramide (R,R)-39 showed higher catalytic activity and enantioselectivity compared to quinidine-based thioureas organocatalysts [53]. Single-step catalytic hydrogenation and acid-catalyzed deprotection of (S)-40 over Raney Ni in acetic acid than decarboxylation of the resulting malonic-acid derivative (S)-41 with 6N HCl afforded (S)-Pregabalin 9 hydrochloride in 52% yield for three steps.
The use of catalyst 42 (Scheme 11A) incorporating squaramide and hydroquinidine units allowed one to perform an enantioselective Michael addition of malonate to nitroolefin (E)-2 in brine due to the “hydrophobic hydration effect” [54,55]. Under these conditions, hydrophobic hydroquinidine-squaramide catalyst 42 directed the reaction towards the corresponding Michael adduct (S)-43 with high yield and enantioselectivity [43]. After the extraction of the Michael adduct (S)-43 with methylcyclohexane, catalyst 42 was recoveried in quantitative yield (>99%) by simple filtration permitted the catalyst recycling. Hydrogenation of the crude Michael adduct (S)-43 using Raney Ni followed by hydrolysis of resulting γ-lactam (3S,4R)-18 with 6N HCl afforded (S)-pregabalin 9 in enantiomerically pure form after simple recrystallization from 2-propanol/water. The developed procedure in brine media was also applied for the scalable synthesis of enantiomerically pure (S)-Rolipram 12 (Scheme 11B) from aryl nitroolefin (E)-10 using 10 mol% of hydroquinine-based squaramide catalyst 44.
The hydrophobic dihydroquinine-squaramide derivative 44 (Scheme 11B) also demonstrated its efficiency in brine media for enantioselective Michael addition of the diethyl dithiomalonates to (E)-α-methyl-β-nitrostyrenes 47 (Scheme 12) [56]. In these cases, a high loading of 44 (15 mol%) and lowering the reaction temperature to 0 °C were required for the reaction to proceed with high enantioselectivities. At the same time, no reaction was observed when catalyst 44 was used in organic solvents. The resulting adducts (S)-48 could further be converted into GABA analogs bearing an all-carbon quaternary stereocenter at the β-position. After the separation of the catalyst, the crude products (S)-48 were subjected to reduction affording γ-lactam thioesters 49, which were hydrolyzed with 6N HCl into β-methylated analogs of Phenibut (S)-50a and Baclofen (S)-50b as hydrochloride salts. At the same time, the β-methylated analog of Rolipram (S)-50c was prepared by basic hydrolysis and decarboxylation of corresponding γ-lactam thioester 49.
An enantioselective decarboxylative Michael addition between malonic-acid half-thioester [57] and β-nitroolefins providing access to γ-nitro thioester was demonstrated to catalyze with cinchona-based squaramide [58] and cinchona-based urea [59] organocatalysts under mild reaction conditions. Performing the reaction of malonic-acid half-thioester 51 (Scheme 13) and β-nitrostyrene (E)-1 with loading of 5 mol% of bifunctional quinine-based squaramide 52 as the most active and selective organocatalyst in methyl tert-butyl ether (MTBE) at 45 °C resulted in addition product (S)-53 in high yield and excellent enantioselecitivity within 22 h [58]. It is noteworthy that using E or Z isomers of the starting β-nitrostyrene 1 afforded the (S)-enantiomer of the Michael adduct 53 with the same ee value. The nitro thioester (S)-53 was converted with Raney Ni and H3PO4 in THF by intramolecular cyclization and recrystallization into enantiomerically pure γ-butyrolactam (S)-16. Finally, hydrolysis of resulting γ-butyrolactam (S)-16 could be performed with 6N HCl under reflux to deliver the (S)-Baclofen 8 as its hydrochloride salt in 78% yield.
The bifunctional bisalkaloid organocatalyst 54 (Scheme 14) was developed and successfully tested for the enantioselective conjugate addition of malonates to nitroalkenes [60]. The best result in terms of reactivity and selectivity was achieved using dimethyl malonate, 1 mol% of organocatalyst 54 in THF at room temperature. Under optimal conditions, the Michael reaction of dimethyl malonate with β-nitrostyrene (E)-1 proceeded well to give the corresponding adduct (R)-38 with excellent yield and enantioselectivity (after recrystallization). The obtained adduct (R)-38 was converted into (R)-Baclofen 8 hydrochloride according to the synthetic sequence demonstrated in Scheme 3.
The highly stereoselective (>99% ee) conjugate addition of acetophenone to β-nitrostyrenes (E)-1 and (E)-20 (Scheme 15) was completed using primary amine-thiourea organocatalyst [61] based on (S)-di-tert-butyl aspartate. With the optimal organocatalyst 55, derived from (1R,2R)-diphenylethylenediamine the Michael addition provided adducts (S)-56 and (S)-57 at low catalyst loading (5 mol%) in excellent yield [62]. After Bayer–Villiger oxidation of (S)-56 and (S)-57, corresponding γ-nitro esters (S)-58 and (S)-59 could be transformed into the (S)-Baclofen 8 and (S)-Phenibut 24 according to described above procedures. The enantiomer of 55 was also utilized as organocatalyst for the efficient synthesis of (R)-Baclofen 8.
Catalytic asymmetric version of the Michael addition reaction between malonates and nitroalkenes was also achieved by using chiral metal−ligand complexes [63,64]. For example, the highly enantioselective Michael addition of tert-butyl phenyl malonate to the β-nitrostyrene (E)-20 (Scheme 16) was developed in toluene at room temperature using the easily accessible chiral bis-(cyclohexyldiamine)-based Ni(II) complex 60 (2 mol%) [65]. This process offered a route for the synthesis of β-nitro derivative (S)-61 in 92% yield and 93% ee (1.8:1 diastereoisomeric ratio). Reduction and cyclization of (S)-61 with the more reactive ester group afforded the γ-lactam (3R,4S)-62 as one diastereomer. Final hydrolysis and decarboxylation of the γ-lactam (3R,4S)-62 with 6N HCl under reflux produced the (S)-Phenibut 24 as the hydrochloride salt in quantitative yield. Later, several heterogeneous catalysts incorporating chiral bis(cyclohexyldiamine)-based Ni(II) complexes were developed for the asymmetric Michael addition of malonates to both aromatic [66,67] and aliphatic [68,69,70] nitroalkenes. These catalysts demonstrated good activities, enantioselectivities, reusability, and applicability for the multistep sequential flow synthesis of the (S)-Pregabalin [69].
On the basis of mechanistic studies of conjugate addition reactions catalyzed by chiral nickel(II)-diamine complexes is proposed that the malonate displaces one diamine ligand of the catalyst generating the chiral enolate I (Scheme 17) [65]. Coordination of the nitroalkene to the nickel center of I leads to the intermediate II, and subsequent addition of enolate to the bound nitroalkene affords the 1,4-addition intermediate III. Then intermolecular proton transfer and displacement of the Michael product with another molecule of the malonate regenerates chiral enolate I.
The application of the chiral nickel(II)-diamine catalysts was extended to 1,4-selective asymmetric addition of malonates to nitroenynes [71]. The 1,4-addition of di-tert-butyl malonate to nitroenyne 63 (Scheme 18) in the presence of 2 mol% of Ni(II) complex 64 as the catalyst proceeded regioselectively under mild conditions, affording β-alkynyl nitro acid (R)-65 in good yield and high enantioselectivity (91% ee). Enantiomerical purity of product (R)-65 could be improved to 99% ee by single recrystallization. This protocol allowed to obtain β-alkynyl acid (R)-66 after decarboxylation of (R)-65 in the presence of TsOH under reflux. Reaction of β-alkynyl acid (R)-66 with oxalyl chloride and MeOH gave β-alkynyl ester (R)-67 in 64% yield, which by reduction of the nitro group followed by the treatment with di-tert-butyl dicarbonate (Boc)2O afforded the N-Boc β-alkynyl-γ-amino ester (R)-68. The ester (R)-68, as the common intermediate, was converted to the β-alkynyl-γ-amino acid (R)-69 and β-alkyny-γ-lactam (R)-70 using standard procedures without loss of enantiomeric purity.
The catalytic asymmetric decarboxylative 1,4-addition reaction of malonic-acid half-thioester to nitroalkene (E)-10 (Scheme 19) was realized by employing heterobimetallic system [72] with transition metal, rare-earth metal, and dinucleating amino-phenol ligand (S,S)-71. Screening of the transition metal and rare-earth metal combination revealed that Ni/La/ligand (S,S)-71 catalyst in the presence of phosphine oxide 72 as an achiral additive gave the best catalytic activity and selectivity delivering the addition product (S)-73 in 80% yield and 93% ee [73]. The synthetic utility of the asymmetric decarboxylative 1,4-addition reaction was demonstrated by the reduction and cyclization of (S)-73 with Zn and (CH3)3SiCl to (S)-Rolipram 12 in 83% yield.
Additionally, asymmetric 1,4-addition of dimethyl malonate to β-nitrostyrene (E)-1 catalyzed by polymer-supported CaCl2-pyridinebisoxazoline (Pybox) complex [74] was developed and implemented in continuous-flow synthesis of (S)-Baclofen precursor (3R,4S)-74 using a series of different heterogeneous catalysts (Figure 2) [75]. In the first step, amine-modified silica gel/molecular sieves 4Å column reactor was used for condensation of p-chlorobenzaldehyde and nitromethane into β-nitrostyrene (E)-1 in toluene. Then sequential-flow enantioselective asymmetric 1,4-addition of dimethyl malonate to β-nitrostyrene (E)-1 proceeded into the second column reactor with polymer-supported CaCl2-Pybox chiral catalyst to give adduct (S)-38 with high enantioselectivity (92% ee). Further, the third column reactor containing dimethylpolysilane (DMPS)-modified Pt catalyst supported on activated carbon (AC) and calcium phosphate (CP) was successfully employed for flow chemoselective hydrogenation and intramolecular cyclization of adduct (S)-38 to give 5 g of lactam (3R,4S)-73 in 93–96% overall yield based on nitromethane with 92% ee during the 69-h process. Finally, (S)-Baclofen 8 was obtained according to the standard procedure. Slightly modifying the continuous flow system using polymer-supported CaCl2-Pybox chiral catalyst was also employed in the synthesis of (R)-Phenibut [76] and (R)-Rolipram [77]. However, the polymer-supported CaCl2-Pybox chiral catalyst demonstrated poor enantioselectivities in reactions of nitroolefins bearing primary aliphatic substituents.

3. Michael Additions of Cyanide or Nitroalkanes to α,β-Unsaturated Carbonyl Compounds

Asymmetric conjugate addition of cyanide and nitroalkanes to α,β-unsaturated carbonyl substrates was another practical route for the synthesis of β-substituted GABA derivatives. For example, the Michael addition of diethylaluminum cyanide to substrate (R)-75 (Scheme 20) bearing oxazolidinone chiral auxiliary [78] was conducted as the key step in the synthesis of (S)-Pregabalin 9 from commercially available starting materials [79]. Conjugate addition was performed in toluene at 0 °C to produce addition adduct (4R,3′S)-76 in satisfactory yield and moderate dr (87:13). The diastereomerically pure (4R,3′S)-76 was provided after purification with column chromatography on silica gel in a 57% yield. The observed stereoselectivity was attributed to the approach of the diethylaluminum cyanide mainly from the less hindered Si face opposite to phenyl group in the oxazolidinone auxiliary. The removal of oxazolidinone chiral auxiliary by treating with LiOH and H2O2 in aqueous THF and reduction of the cyano group by hydrogenolysis under Raney Ni afforded (S)-pregabalin 9 in 95% yield for two steps. Moreover, acetone cyanohydrin was also effectively used as cyanide source for diastereoselective conjugate addition to α,β-unsaturated oxazolidinone (S)-77 (Scheme 20) [80]. The hydrocyanation of (S)-77 cleanly proceeded with two equivalents of acetone cyanohydrin and 10 mol% of Sm(III) isopropoxide as catalyst in toluene to give the addition adduct (4S,3′R)-78 in 75% yield and 88:12 dr. The diastereomerically pure product (4S,3′R)-78 was chromatographically isolated in 66% yield. The catalytic hydrogenation of the cyano group with simultaneous cleavage of oxazolidinone chiral auxiliary over platinum oxide afforded the lactam (R)-79 in 75% yield and 96% ee, which acidic hydrolysis led to the (R)-Pregabalin 9 as hydrochloride in 95% yield with retention of the enantiomeric purity. A similar route was employed for the synthesis of (S)-Baclofen 8 using aryl-substituted substrate (S)-80, acetone cyanohydrin and Sm(Oi-Pr)3 as catalyst (Scheme 20) [80]. Under standard conditions, the diastereomerically pure nitrile adduct (4S,3′S)-81 was obtained in 62% yield. Selective reduction of the cyano group with such reagents as NaBH4 and NiCl2 and hydrolysis of resulting lactam (S)-82 provided (S)-Baclofen 8 in excellent yield and enantiomeric purity.
Similarly, oxazolidinone as a chiral auxiliary was used to generate a new stereogenic center in the Michael addition between nitromethane and α,β-unsaturated oxazolidinone ®-75 with Cs2CO3 as a base providing a diastereomerically pure addition adduct (4R,3′S)-83 in 34% yield after two recrystallizations from isopropanol (Scheme 21) [81]. The oxazolidinone chiral auxiliary was removed by treating (4R,3′S)-83 with alkaline hydrogen peroxide to give the γ-nitroacid (S)-84. (S)-Pregabalin 9 was obtained after hydrogenation of γ-nitroacid (S)-84 using Pd/C in 26% overall yield for three steps from ®-75 with >99% ee.
Chiral α,β-unsaturated oxazolidino®(R)-85 (Scheme 22) was successfully used in tetramethylguanidine (TMG)-catalyzed diastereoselective conjugate addition of nitromethane with good stereoselectivity (93:7) [82]. The addition product (4R,3′R)-86, after crystallization from ethyl acetate, was obtained in 78% yield and 99% de. Subsequent synthetic operations involving elimination of the chiral auxiliary in basic solution, hydrogenation of NO2 group with Raney Ni, and recrystallization from water provided (R)-Baclofen 8 in 65% yield with 99% ee.
Synthesis of (S)-Pregabalin 9 (Scheme 23) starting from 2,3-O-isopropylidene-D-glyceraldehyde (R)-87 as the chiral source included highly diastereoselective Michael addition of nitromethane to α,β-unsaturated ester (S)-88 that produced chiral γ-nitroester derivative (S,S)-89 as crucial step [83]. Initially, Wadsworth–Emmons olefination of 2,3-O-isopropylidene-D-glyceraldehyde (R)-87 with triethyl phosphonoacetate afforded α,β-unsaturated ester (S)-88 as a mixture of E:Z isomers in ratio 9:1, which underwent conjugate addition of nitromethane in the presence of TBAF to give, after purification, γ-nitroester (S,S)-89 in 75% yield. Next, the γ-nitroester (S,S)-89 was treated with ammonium formate in the presence of Pd(OH)2/C, leading to the reduction of the nitro group with concomitant cyclization to the γ-lactam (S,S)-90 in 85% yield. The corresponding N-Boc protected γ-lactam (S,S)-91 was prepared by reaction with di-(tert-butyl) dicarbonate in the presence of DMAP and triethylamine. Then chemoselective cleavage of the ketal was achieved by using 90% aqueous acetic acid to afford diol (S,S)-92 in quantitative yield. Oxidative cleavage of diol (S,S)-92 with NaIO4 and Wittig olefination of resultant aldehyde (S)-93 with isopropylidenetriphenyl phosphorane afforded β-isobutenyl γ-lactam (R)-94 in 60% yield. Hydrolysis of the γ-lactam (R)-94 with 1M LiOH in THF at room temperature afforded acid (R)-95, which easily hydrogenated in the presence of catalytic Pd(OH)2/C and N-deprotected by acidolysis of the Boc-carbamate with aqueous HCl to provide the enantiomerically pure (S)-Pregabalin 9 in quantitative yield.
The enantioselective conjugate addition of nitromethane to α,β-unsaturated aldehyde (E)-96 (Scheme 24) was developed employing O-TMS-protected prolinol (R)-97 [84,85] as organocatalyst. The asymmetric reaction catalyzed with 20 mol% of (R)-97 in the presence of benzoic acid as an additive in EtOH furnished the Michael adduct (R)-98 in 73–83% yield with excellent 94–96% ee [30,86]. When the asymmetric addition of nitromethane to α,β-unsaturated aldehyde (E)-96 was carried out with 10 mol% loading of (R)-97 and lithium acetate as an additive in CH2Cl2/MeOH, the Michael adduct (R)-98 was obtained in 62% yield and 91% ee [87]. The mechanism of this process suggests that organocatalyst (R)-97 would react with the α,β-unsaturated aldehyde (E)-96 leading to the iminium ion A in which one of the enantiofaces is shielded by the large diphenyl(trimethylsilyloxy)methyl group. Thus, nucleophilic nitronate would attack iminium ion A from the Re face to form the observed (R)-configurated adduct 98. The presence of benzoic acid, as well as lithium acetate as additives, could accelerate the generation of the iminium ion A. Successive synthetic transformation of the γ-nitro aldehyde (R)-98 involving oxidation with NaClO2 in the presence of H2O2 and KHPO4 to γ-nitro acid (R)-99 and hydrogenation in the presence of Raney Ni efficiently provided (R)-Baclofen 8 hydrochloride. In the subsequent study preparation of (R)-Baclofen 8 from commercially available materials was also accomplished by employing a one-pot synthetic procedure that involved DBU catalyzed aldol condensation of p-chlorobenzaldehyde and acetaldehyde, O-TMS-protected prolinol mediated asymmetric Michael addition of nitromethane in the presence of formic acid as an additive, oxidation, and reduction [88]. The one-pot procedure afforded (R)-Baclofen 8 in 31% overall yield with high enantioselectivity (93% ee).
A one-pot procedure for the synthesis of γ-nitroester (S)-100 (Scheme 25) as a key intermediate of the (S)-Baclofen 8 was developed based on the Michael addition of nitromethane to α,β-unsaturated aldehyde (E)-96 catalyzed with diphenylprolinol silyl ether (S)-97 in MeOH, followed by the oxidative esterification using N-bromosuccinimide (NBS) as the oxidant [89]. The γ-nitroester (S)-100 was obtained in 94% enantiomeric excess. Next, the γ-nitroester (S)-100 was treated with NaBH4 in EtOH in the presence of NiCl2, leading to the reduction of the nitro group with accompanying cyclization to the γ-lactam (S)-82, which upon hydrolysis with 6M HCl afforded the (S)-Baclofen 8 as the hydrochloride in 94% yield.
Recently, a highly efficient process for continuous flow asymmetric synthesis [90,91,92] of Rolipram was developed. This three-step process implied (1) an initial organocatalytic flow synthesis based on telescoped flow asymmetric conjugate addition, (2) oxidative aldehyde esterification sequence using in situ-generated persulfuric acid [H2SO5] as oxidant, and (3) a final nitro reduction and lactamization. The process is initiated by an enantioselective Michael-type addition of nitromethane with the appropriately substituted cinnamaldehyde derivative (E)-101 (Scheme 26) in the presence of cross-linked polystyrene-supported cis-4-hydroxydiphenylprolinol tert-butyldimethylsilyl (TBS) ether as the chiral catalyst for the formation of the γ-nitroaldehyde (S)-102, which was obtained after adjusting the reaction conditions in 94% ee and 97% of conversion [93]. Taken into account that the initial chiral γ-nitroaldehyde (S)-102 was labile and could decompose during purification, it was oxidized efficiently towards the corresponding γ-nitroester (S)-103 by the action of the oxidizing agent persulfuric acid [H2SO5], in situ generated by the reaction of concentrated sulfuric acid with H2O2 in a continuous-flow oxidative esterification. To complete the flow synthesis of Rolipram, a nitro-reduction was necessary, which was carried out by treatment of the γ-nitroester (S)-103 with trichlorosilane (HSiCl3) in dry CH3CN. It was also observed that 4 equiv of N,N-diisopropylethylamine (DIEA) and 4 equiv of trichlorosilane were necessary to achieve quantitative and selective final nitro reduction/lactamization. In this manner, the enantioselective flow synthesis of (S)-Rolipram 12 was carried out after 4 h run in 83% yield and 94% ee. A similar continuous-flow process was developed for the asymmetric synthesis of substituted γ-nitrobutyric acids as key intermediates in the synthesis of Baclofen and other GABA derivatives [94].
Bifunctional squaramide catalyst (S)-104 (Scheme 27) possessing pyrrolidine moiety was employed in the enantioselective Michael additions of nitromethane to α,β-unsaturated aldehyde (E)-96 [50]. The organocatalytic reaction in the best conditions provided the adducts (R)-98 in synthetically valuable yield (64%) with high enantiomeric purity (92% ee). The enantioselectivity is explained through the formation of an enamine transition state TS allowing the Re-face attack of nitronate to yield the observed (R)-configurated adducts. Finally, compound (R)-98 was converted into (R)-Baclofen 8 in oxidation/reduction/acid hydrolysis sequence.
Primary amine-thiourea catalyst 106 (Scheme 28) synthesized from (1R,2R)-diaminocyclohexane and dehydroabietic amine allowed one to perform asymmetric Michael addition of nitroalkanes to α,β-unsaturated ketones providing high enantioselectivities for these Michael acceptors [95]. For example, the Michael addition of nitromethane to alkyl cinnamyl ketones 105 in toluene as the optimal solvent gave the γ-nitro ketones (R)-107 in 98% ee. It should be noted that excess nitromethane and the use of acetic acid as an additive shortened the reaction times and increased the yield of the desired products (R)-107 to 83–85%. Compound 105 is activated via hydrogen-bonding interactions between two NH moieties of the thiourea catalyst 106 and the carbonyl group. Moreover, the amine group of the catalyst 106 provides the Re face approach of nitromethane in TS leading to formation of Michael adducts 107 with R-configuration. Subsequent Baeyer–Villiger oxidation then reduction of the nitro group gave the γ-amino esters (R)-108 in an excellent yield. Finally, acid hydrolysis of (R)-108 provided the target (R)-Phenibut 24 and (R)-Baclofen 8 in 95% and 93% yield, respectively.
Enantioselective conjugate addition of nitromethane to α,β-unsaturated thioamide 109 (Scheme 29) was promoted by a mesitylcopper/(R)-DTBM-Segphos precatalyst exhibiting dual functions to activate both pronucleophile and electrophile. The reaction was carried out at 5 mol% precatalyst loading in n-hexane/toluene solvent system at room temperature affording γ-nitrothioamide (R)-110 in 92% yield with high enantioselectivity 99% ee [96]. Initially, deprotonation of nitromethane with chiral phosphine ligand/mesitylcopper delivering Cu-nitronate and subsequent coordination of thioamide 109 generated complex A (Scheme 29). Then, enantioselective C-C bond formation gave the intermediate Cu-thioamide enolate B, which as a soft Lewis acid/hard Brønsted base cooperative catalyst promoted proton exchange with nitromethane to give the addition adduct (R)-110 with simultaneous regeneration of complex A. Transformation of the γ-nitrothioamide (R)-110 to γ-nitrothioester (R)-111 was achieved in high yield by in situ S-alkylation and hydrolysis with MeI/TFA in wet THF. Finally, γ-nitrothioester (R)-111 was converted into (R)-Baclofen 8 by standard hydrolysis and reduction, highlighting the synthetic utility of the present protocol.

4. Michael Addition of Organometallic Reagents to α,β-Unsaturated Carbonyl Compounds

Γ-Lactams are widely used as starting materials for numerous syntheses, since this backbone can be readily converted into other functional groups through many conventional organic transformations. In particular, a large number of enantioselective syntheses have been reported that use γ-unsaturated lactams as substrates for obtaining compounds, possessing a C-4 chirality, of biological or synthetic interest. In this context, the group of Lin, Feng et al. in 2011 [97,98] described an enantioselective synthesis of β-substituted γ-lactams 113 (Scheme 30) by asymmetric 1,4-addition of arylboronic acids to N-Boc-protected α,β-unsaturated γ-lactams 112, a process that takes place in the presence of rhodium complexes (i.e., [RhCl(C2H4)2]2) as catalysts and using the chiral bicyclo [3.3.0]diene 114 as ligand. The designed chiral diene ligands were easily prepared by the same group in a three-step sequence [99]. An additive such as Et3N or KHF2, depending on the nature of the starting arylboronic acid, was necessary for this reaction. The process worked well, and excellent chemical yield and enantioselectivity were obtained in all the cases. However, in the case of arylboronic acids with electron-withdrawing substituents, a prolonged reaction time was required to reach the same high level of enantioselectivity and chemical yield.
This methodology was applied in the concise synthesis of some bioactive compounds such as the (R)-Baclofen 8 hydrochloride and the (R)-Rolipram 12 (Scheme 31). In the first synthesis, the corresponding chiral β-substituted γ-lactam (R)-113a was converted into (R)-Baclofen 8 in two-step by treatment with TFA, which underwent deprotection of Boc, followed by acid hydrolysis of (R)-115 with 6N HCl. In the second case, (R)-Rolipram 12 was isolated directly after deprotection of (R)-113b with TFA in CH2Cl2 at room temperature. A high chemical yield was obtained in either case.
In the same year, Liao et al. [100] reported a conceptually similar procedure in a process, which implies a rhodium-catalyzed addition of arylboronic acids to γ-phthalimido-substituted α,β-unsaturated esters 116 (easily prepared from 4-bromobut-2-enoate and phthalimide potassium salt) (Scheme 32) in the presence of chiral bis-sulfoxide ligand (R,R)-117 [101]. The process afforded β-substituted GABA derivatives 118 under mild conditions, with high yields and enantioselectivities (up to 96% ee), and using a Rh/(R,R)-1,2-bis(tert-butylsulfinyl)benzene complex [(R,R)-117/RhCl]2 as catalyst.
This methodology has been successfully applied in the synthesis of the optically pure (S)-Baclofen 8 and (S)-Rolipram 12. In the first case, a simple acidic hydrolysis (6N HCl) of (S)-118a (Scheme 33), afforded (S)-Baclofen 8 in 93% yield; and in the latter case, starting from the corresponding addition product (S)-118b, a two-step procedure that implies that treatment with aqueous hydrazine followed by reflux with Et3N in toluene provided the target (S)-Rolipram 12 in 78% yield.
Another variant for the enantioselective synthesis of these two important γ-aminobutyric acid analogues was reported by Helmchen and coworkers [102], who studied the 1,4-asymmetric addition of arylboronic acids to 4-amino-2,3-enoic acid ethyl ester derivatives 119 (Scheme 34) in a process catalyzed by Rh(I) complexes and in the presence of chiral BINAP ligands. Different reaction conditions were studied, in particular the nature of the protecting groups, the base, and the ligand for preparation of adducts 120. The best results were obtained when the base was Cs2CO3, the ligand (S)-BINAP, Boc as protecting group, a temperature of 100 °C, and [Rh(acac)(C2H4)2] as catalyst. Under these conditions and using the appropriate substituted arylboronic acid, (R)-Rolipram 12 and (R)-Baclofen 8 were obtained in good chemical yields and enantioselectivity.
In 1997, Alvarez-Builla et al. [103] developed a quite simple and stereoselective strategy for obtaining (R)-Rolipram 12 (Scheme 35) starting from L-glutamic acid. The process was initiated by a conjugate Michael addition of a suitable arylcuprate to a modified pyroglutamic derivative (S)-121. The authors indicated that the steric hindrance provided by the bulky silyloxy group in the protected alcohol (S)-121 (R = CH2OSiPh2t-Bu) was responsible for the final stereoselectivity observed. The addition product was transformed into (R)-Rolipram 12 by an initial deprotection of the adduct (2S,3R)-122 with triethylammonium fluoride (TEAF), followed by oxidation of the alcohol to carboxylic acid (2S,3R)-123 with the Jones reagent, and final simultaneous decarboxylation and deprotection of the Boc group.
The Michael reaction of unsaturated pyroglutamate derivative 124 (Scheme 36) protected as a 5-methyl-2,7,8-trioxabicyclo [3.2.1]octane ester with Grignard-cuprate containing 4-chlorophenyl group in the presence of trimethylsilyl chloride took place in good yield to give the adduct (4R,5S)-125. The stereoselectivity of the reaction can be explained by attacks of the organocuprate reagent at β-C of the C=C double bond from the less-hindered face [104]. The ester (4R,5S)-125 underwent HCl-promoted methanolysis accompanying deprotection of the Boc group to give corresponding methyl ester (2S,3R)-126 in good yield. Subsequent benzyl protection of lactam nitrogen followed by base hydrolysis of the methyl ester (2S,3R)-127 yielded pyroglutamic acid (2S,3R)-128. Decarboxylation and debenzylation produced pyrrolidin-2-one (R)-129, which finally hydrolyzed by treatment with 6N HCl into (R)-Baclofen 8 hydrochloride [105].
An earlier route reported by Meyers [106] in 1993 allowed the asymmetric synthesis of (R)-Rolipram, as part of a more general study, by conjugate addition of cyanocuprates to bicycle chiral α,β-unsaturated lactams. Once the appropriate starting aryl lactam 130 (Scheme 37) had been prepared in three steps from γ-ketocarboxylic acids [107], the subsequent treatment with the appropriate cyanocuprate (−78 °C, THF) gave 131 in good yield a product derived from the conjugate addition to the exo (β) face of the C=C double bond. Removal of the α-carboxy function was carried out easily by hydrogenolysis followed by heating in toluene, affording 132 in good yield (79%) and excellent diastereoselectivity. The cleavage of the bicyclic system was accomplished in two steps, first by treatment with an excess of calcium metal in liquid ammonia, followed by final reduction using acidic sodium cyanoborohydride to afford the target (R)-Rolipram 12 in good yield. Other examples, i.e., R = Me, n-Bu, Ph were also reported in this relatively general study.

5. Miscellaneous Types of Michael Addition

One of the first enantioselective syntheses of (R)-Rolipram was developed by Langlois et al. in 1997 [108], who reported a stereoselective conjugate addition of cyanide by means of AlEt2CN to an activated α,β-unsaturated oxazoline (R)-133 (Scheme 38), which was obtained in four steps starting from isovanillin and (R)-phenylglycinol as chiral auxiliary. In the process, strict control of the temperature (18 °C) was necessary to obtain an incomplete conversion (ca 50%), as at higher temperatures (i.e., 35 °C), the formation of by-products was mainly observed. Under these conditions, a mixture of two diastereoisomers, separated by flash chromatography, was obtained in 49% yield and a ratio of 63:37. The cyano group of the pure major diastereoisomer (R,R)-134 was selectively reduced with NaBH4∙NiCl2 to the amidine (R,R)-135, which after alkaline hydrolysis (2N NaOH, 95% EtOH) and recovery of the chiral auxiliary was converted into (R)-rolipram 12 in 55% yield from the nitrile (R,R)-134.
Among the different nucleophiles used for the enantioselective conjugate addition to unsaturated olefins, the α-carbanions derived from α,α-dithioacetals, which are known as synthetic chameleons, stand out for their relevance in the synthesis of a large number of products of biological importance. From the initial work of Koga and coworkers in 1985 [109], numerous studies have been reported emphasizing the significance of this strategy in the preparation of a large number of compounds of synthetic interest. On the other hand, electron-deficient alkenes such as nitroalkenes are important synthetic feedstocks in the construction of complex and chiral molecules in the presence of an array of stoichiometric and catalytic ligands. In this sense, in 2015 Benaglia, Gaggero et al. [110] reported the first enantioselective organocatalytic conjugate addition of 2-carboxythioesters-1,3-dithiane 136 (Scheme 39) to different nitroalkenes, particularly styrene derivatives, under mild conditions and in the presence of cinchona bifunctional derivatives as organocatalysts. Solvents such as toluene, room temperature, and cinchona derivative 137 (20 mol%) provided the best results in terms of chemical yield and enantioselectivity. Unfortunately, alkyl nitroalkenes did not react under the described reaction conditions. Conversion of the addition products 138 into β-nitro esters 139 was carried out in two steps by reaction first with NBS in acetone, followed by treatment with silver trifluoroacetate. This strategy was applied to the synthesis of the (S)-Baclofen 8 and other derivatives of synthetic value.
The stereochemical outcome was explained through the model depicted in Scheme 40, in which the nitroolefin is activated both by the hydrogen atom in the N-H section of the thiourea moiety and the ion pair formed between the charged quinuclidine nitrogen and the nucleophile. The thioester attack on the nitroalkene Si face occurs as shown in Scheme 40.
Quite recently, Nakamura et al. reported [111] a general and improved method for the catalytic enantioselective conjugate addition of α,α-dithioacetonitrile derivatives with aromatic and aliphatic β-nitroalkenes as electron-deficient alkenes, with the process being catalyzed by chiral bis(imidazoline)-palladium pincer-type complexes. In that study, the nature of the starting α,α-dithioacetal plays a crucial role in the success of the process. In addition, α,α-dithioacetonitriles are a recognized class of synthetic intermediates [112] as they can behave as cyanomethyl and cyanocarbonyl anion equivalents in their reactions with electron-deficient alkenes. After an exhaustive study with a variety of nitroolefins, 1,3-dithiane derivatives, catalysts, and reaction conditions, the authors concluded that the best results were obtained by reacting 1,3-dithiolane-2-carbonitrile 140 (Scheme 41) as α,α-dithioacetonitrile (1.5 equiv) with a variety of aromatic and aliphatic nitroalkenes in the presence of 5 mol% of the chiral palladium pincer complex Phebim-Pd complex 141 and Ag(acac) (5 mol%) at room temperature and ethyl acetate as solvent. The reaction afforded the conjugate addition products 142 with high yield and enantioselectivities (up to 98% ee).
The process is quite general, being effective with electron-withdrawing, electron-rich, heteroaryl, and naphthyl aromatic nitroalkenes. Moreover, the reaction with aliphatic nitroalkenes also gave the corresponding products in good yields and enantioselectivity. It is noteworthy that many other different α-acetonitrile carbanion equivalents such as acetonitrile, phenylthioacetonitrile, cyanoacetic acid, etc., were not effective, thus indicating the essential role of the nitrile and the dithiolane groups in the global process. The proposed mechanism for this reaction is depicted in Scheme 42. The palladium acetylacetonate-complex I was initially formed by an exchange reaction of Ag(acac) with Phebim-PdBr 141. This complex coordinated next with the cyano group of 140 to afford a new cationic complex II, which was deprotonated by acetylacetonate to give the Pd-ketenimide complex III. Then, the reaction of III with nitroalkenes took place to afford complex IV, which finally gave the protonated final addition product 142 and subsequent regeneration of the catalyst. This mechanism was corroborated by DFT calculations and ESI-MS spectroscopic analysis.
This methodology was applied to the synthesis of γ-lactams, and particularly to the (R)-Rolipram 12 (Scheme 43). Thus, the previously obtained addition product (R)-142a reacted first with InCl3∙4H2O and acetaldoxime to provide the corresponding amide (R)-143, whose nitro group was reduced with Fe/NH4Cl. Finally, the reductive desulfurination of the γ-lactam 144 with NiCl2∙6H2O/NaBH4 afforded (R)-Rolipram 12 [113].
In comparison with the enolate equivalents derived from α-carbon carbonyl compounds, the corresponding “inert” β-carbons of saturated carbonyl derivatives (β-sp3) have been far less studied in their behavior than nucleophiles [114]. In this context, Chi et al. [115] reported in 2013 an interesting β-carbon catalytic activation of simple saturated esters using N-heterocyclic carbenes (NHC) as organocatalysts in enantioselective reactions with various electrophiles such as enones, trifluoroketones, and hydrazones (Scheme 44).
After checking different reaction conditions including solvent, temperature, and achiral and chiral triazolium organocatalysts, the authors concluded that the best results in terms of efficiency (chemical yield and selectivity) were obtained by using the bulky chiral triazolium salt A (20 mol%) as organocatalyst in the reaction with enones as electrophiles, acetonitrile as solvent, room temperature, and in the presence of an excess of DBU as base, which was necessary to achieve the two key deprotonation steps. In this way, chiral cyclopentenes were obtained, as shown in Scheme 45. The usefulness of this strategy was also demonstrated in their reaction with other electrophiles such as trifluoroketones and hydrazones affording fluorinated γ-lactones and nonfluorinated γ-lactams, respectively. The nature of the solvent depends on the electrophile, with toluene being the solvent of choice in the case of trifluoroketones and ethyl acetate for hydrazones. The process took place, in general, with moderate-to-good chemical yields, moderate diastereoselectivity, and good enantioselectivity, obtaining the best results with β-(hetero)aryl substituted esters (Scheme 45).
This methodology was used for the synthesis of bioactive molecules and building blocks of synthetic interest. Thus, chiral cyclopentenes are starting materials for optically enriched 1,2-diols and amino alcohols. γ-Butyrolactones are the key unit in the synthesis of some natural products, and γ-lactams have been used in the preparation of some pharmaceuticals such as (S)-Baclofen 8 and (R)-Rolipram 12 as shown in Scheme 46.
To explain the above-mentioned results, the authors proposed the catalytic cycle outlined in Scheme 47. The initial addition of the NHC catalyst to the ester I provides a new NHC-bounded ester intermediate II, which after deprotonation forms the enolate intermediate III bearing a nucleophilic α-carbon. Deprotonation of the β-CHs of enolate III, favored among others by the electron-withdrawing nature of the triazolium moeity, leads to intermediate IV. Next, the β-carbon nucleophilic center of IV undergoes a Michael addition with enones to give the cyclopentane intermediate VII, via a cascade process involving an H+ transfer and an aldol reaction (V to VII). Finally, lactone formation and decarboxylation provided the target cyclopentene product IX.
Recently, the conjugate addition of NH2-unprotected tert-butyl glycinate 145 (Scheme 48) to α,β-unsaturated esters 146 using chiral pyridoxal (S)-147 as a carbonyl catalyst and DBU as the most appropriate base was found to produce chiral pyroglutamic-acid esters 148 after in situ lactamization with low diastereoselectivities but high enantioselectivities (81–97% ee) for both trans- and cis-diastereomers, which could be separated by column chromatography [116]. Moreover, a Lewis acid such as LiOTf was a necessary additive for the reaction. Under the optimal reaction conditions, aromatic, heteroaromatic, and alkyl α,β-unsaturated esters underwent asymmetric 1,4-conjugated addition and subsequent lactamization providing corresponding pyroglutamic-acid esters 148 in good yields with similar selectivities. Hydrolysis of the mixture of trans- and cis-diastereomers 148a (Scheme 48) using 2M HCl and subsequent decarboxylation afforded Rolipram 12 without loss of enantioselectivity.
The authors propose for conjugate addition of NH2-unprotected tert-butyl glycinate 145 to α,β-unsaturated esters 146 plausible pathway presented on Scheme 49. Condensation of chiral pyridoxal (S)-147 with tert-butyl glycinate 145 gives Schiff base 149. After deprotonation of Schiff base 149 with DBU asymmetric 1,4-conjugated addition of the carbon anion 150 to α,β-unsaturated ester 146 produces adduct 151, which is hydrolyzed to form γ-amino ester 152 and regenerates the pyridoxal catalyst (S)-147. Finally, γ-amino ester 152 undergoes in situ intramolecular cyclization to pyroglutamic-acid esters 148. Since N,N-propyl groups of the amide substituent in pyridoxal (S)-147 catalyst shield up the side of the pyridine ring, the α,β-unsaturated ester 146 approaches the enolate anion in TS from direction opposite to the amide side chain, affording (2R,3S) configuration of pyroglutamic-acid esters 148 (R = Ar). At the same time, the epimerization of trans- and cis-isomers occurring under the basic reaction conditions results in observed low diastereoselectivity of reaction.

6. Conclusions

This review highlights the role and significance of the asymmetric Michael addition reactions as key transformations in the synthesis of β-substituted GABA derivatives used as pharmaceuticals for the treatment of central nervous system disorders. As it follows from the discussed above data, preparation of β-substituted GABA derivatives via asymmetric Michael addition reactions is a rather mature science, allowing access to this class of compounds in desired structural variety and on a large scale. In particular, considerable achievements have been made in organocatalytic conjugate addition reactions of carbon nucleophiles (malonates, aldehydes, ketones, nitroalkanes, and amino-acid derivatives) to electron-deficient alkenes as well as their synthetic applications. Numerous chiral prolinol-, (thi)ourea-, squramide-, and cinchona alkaloid-based organocatalysts developed for Michael addition reactions allow enantioselective transformations with a variety of nucleophiles under mild reaction conditions, yielding γ-nitrocarbonyl compounds with the desired stereochemistry that can be readily reduced and hydrolyzed into β-substituted GABA derivatives. Bifunctional organocatalysts bearing a hydrogen-bonding donor group are frequently employed in these types of transformations, and in many cases high catalytic activity and enantioselectivity were achieved with low catalyst loading. Metal-catalyzed asymmetric Michael reactions are also widely applied in synthesis of β-substituted GABA derivatives, providing high reactivity and stereoselectivity. In addition, organocatalysts and chiral metal complexes can be immobilized and reused without loss of catalytic activity showing a great potential for syntheses on large scale and continuous-flow multi-step preparation of β-substituted GABA derivatives. Although much progress was made in the preparation of β-substituted GABA derivatives using organo- and metal-catalyzed asymmetric Michael addition reactions, there is still opportunities for improvement of existing catalytic systems and the introduction of new catalysts with the systematic studies on the reaction mechanisms. Furthermore, the reported stereochemical data of the enantioselective reactions discussed in this review still ignore the study of the SDE-properties [117,118] of enantiomerically enriched GABA derivatives. Regretfully, SDE properties of γ-amino acids have never been reported so far. On the other hand, it has been unambiguously demonstrated that chiral amines [119,120], α- [121,122], and β-amino acids [123] persistently show a significant magnitude of SDE under a variety of transformations. As a result of the SDE ignorance, the reported values of stereochemical outcome can be recorded with sizable mistakes leading to erroneous presentation synthetic value of the corresponding reactions. Studying the SDE-properties of chiral molecules is an important issue in laboratory practice and drug development [124,125,126]. In particular, the current standards in administering chiral drugs as pure enantiomers emphasize the importance of the SDE phenomenon in the reliable and safe preparation of chiral GABA-containing pharmaceuticals [127,128].
During the revision process of this review, two interesting related reports on synthesis of α-substituted GABA derivatives [129] and synthesis of (S)-(+)-homo-β-proline [130] were cited, which were suggested by the referee.

Author Contributions

Conceptualization, S.F., V.A.S. and A.S.; validation, J.H., J.E., A.L. and A.S.; writing—original draft preparation, J.H., J.E., S.F., A.L. and A.S.; writing—review and editing, J.H., S.F., V.A.S. and A.S.; supervision, V.A.S. and A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21761132021), and IKERBASQUE, Basque Foundation for Science. The financial support from the University of the Basque Country UPV/EHU (UFIQOSYC11/22), Basque Government (GVgrant IT1236-19), and Ministerio de Ciencia e Innovación (grant PID2019-109633GBC21) for. A.L. are also acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Soloshonok, V.A.; Cai, C.; Hruby, V.J.; Van Meervelt, L. Asymmetric synthesis of novel highly sterically constrained (2S, 3S)-3-methyl-3-trifluoromethyl-and (2S, 3S, 4R)-3-trifluoromethyl-4-methylpyroglutamic acids. Tetrahedron 1999, 55, 12045–12058. [Google Scholar] [CrossRef]
  2. Han, J.; Lyutenko, N.V.; Sorochinsky, A.E.; Okawara, A.; Konno, H.; White, S.; Soloshonok, V.A. Tailor-Made Amino Acids in Pharmaceutical Industry: Synthetic Approaches to Aza-Tryptophan Derivatives. Chem. Eur. J. 2021, 27, 17510–17528. [Google Scholar] [CrossRef]
  3. Han, J.; Konno, H.; Sato, T.; Soloshonok, V.A.; Izawa, K. Tailor-made amino acids in the design of small-molecule blockbuster drugs. Eur. J. Med. Chem. 2021, 220, 113448. [Google Scholar] [CrossRef]
  4. Han, J.; Konno, H.; Sato, T.; Izawa, K.; Soloshonok, V.A. Peptidomimetics and Peptide-Based Blockbuster Drugs. Curr. Org. Chem. 2021, 25, 1627–1658. [Google Scholar] [CrossRef]
  5. Liu, J.; Han, J.; Izawa, K.; Sato, T.; White, S.; Meanwell, N.A.; Soloshonok, V.A. Cyclic tailor-made amino acids in the design of modern pharmaceuticals. Eur. J. Med. Chem. 2020, 208, 112736. [Google Scholar] [CrossRef]
  6. Mei, H.; Han, J.; White, S.; Graham, D.J.; Izawa, K.; Sato, T.; Fustero, S.; Meanwell, N.A.; Soloshonok, V.A. Tailor-made amino acids and fluorinated motifs as prominent traits in modern pharmaceuticals. Chem. Eur. J. 2020, 26, 11349–11390. [Google Scholar] [CrossRef]
  7. Liu, A.; Han, J.; Nakano, A.; Konno, H.; Moriwaki, H.; Abe, H.; Izawa, K.; Soloshonok, V.A. New pharmaceuticals approved by FDA in 2020: Small-molecule drugs derived from amino acids and related compounds. Chirality 2022, 34, 86–103. [Google Scholar] [CrossRef]
  8. Yin, Z.; Hu, W.; Zhang, W.; Konno, H.; Moriwaki, H.; Izawa, K.; Han, J.; Soloshonok, V.A. Tailor-made amino acid-derived pharmaceuticals approved by the FDA in 2019. Amino Acids 2020, 52, 1227–1261. [Google Scholar] [CrossRef]
  9. Soloshonok, V.A.; Izawa, K. (Eds.) Asymmetric Synthesis and Application of α-Amino Acids; ACS Symposium Series #1009; Oxford University Press: Washington, DC, USA, 2009. [Google Scholar]
  10. Kent, C.N.; Park, C.; Lindsley, C.W. Classics in chemical neuroscience: Baclofen. ACS Chem. Neurosci. 2020, 11, 1740–1755. [Google Scholar] [CrossRef]
  11. Lapin, I. Phenibut (β-phenyl-GABA): A tranquilizer and nootropic drug. CNS Drug Rev. 2001, 7, 471–481. [Google Scholar] [CrossRef]
  12. Zvejniece, L.; Vavers, E.; Svalbe, B.; Veinberg, G.; Rizhanova, K.; Liepins, V.; Kalvinsh, I.; Dambrova, M. R-phenibut binds to the α2–δ subunit of voltage-dependent calcium channels and exerts gabapentin-like anti-nociceptive effects. Pharmacol. Biochem. Behav. 2015, 137, 23–29. [Google Scholar] [CrossRef]
  13. Dambrova, M.; Zvejniece, L.; Liepinsh, E.; Cirule, H.; Zharkova, O.; Veinberg, G.; Kalvinsh, I. Comparative pharmacological activity of optical isomers of phenibut. Eur. J. Pharmacol. 2008, 583, 128–134. [Google Scholar] [CrossRef]
  14. Tyurenkov, I.N.; Borodkina, L.E.; Bagmetova, V.V.; Berestovitskaya, V.M.; Vasil’eva, O.S. Comparison of nootropic and neuroprotective features of aryl-substituted analogs of gamma-aminobutyric acid. Bull. Exp. Biol. Med. 2016, 160, 465–469. [Google Scholar] [CrossRef]
  15. Silverman, R.B. From basic science to blockbuster drug: The discovery of Lyrica. Angew. Chem. Int. Ed. 2008, 47, 3500–3504. [Google Scholar] [CrossRef]
  16. Reinares, M.; Rosa, A.R.; Franco, C.; Goikolea, J.M.; Fountoulakis, K.; Siamouli, M.; Gonda, X.; Frangou, S.; Vieta, E. A systematic review on the role of anticonvulsants in the treatment of acute bipolar depression. Int. J. Neuropsychop. 2013, 16, 485. [Google Scholar] [CrossRef]
  17. Calandre, E.P.; Rico-Villademoros, F.; Slim, M. Alpha2delta ligands, gabapentin, pregabalin and mirogabalin: A review of their clinical pharmacology and therapeutic use. Expert Rev. Neurother. 2016, 16, 1263–1277. [Google Scholar] [CrossRef]
  18. Zvejniece, L.; Zvejniece, B.; Videja, M.; Stelfa, G.; Vavers, E.; Grinberga, S.; Svalbe, B.; Dambrova, M. Neuroprotective and anti-inflammatory activity of DAT inhibitor R-phenylpiracetam in experimental models of inflammation in male mice. Infammopharmacology 2020, 28, 1283–1292. [Google Scholar] [CrossRef]
  19. Malykh, A.G.; Sadaie, M.R. Piracetam and piracetam-like drugs. Drugs 2010, 70, 287–312. [Google Scholar] [CrossRef]
  20. Kenda, B.M.; Matagne, A.C.; Talaga, P.E.; Pasau, P.M.; Differding, E.; Lallemand, B.I.; Frycia, A.M.; Moureau, F.G.; Klitgaard, H.V.; Gillard, M.R.; et al. Discovery of 4-substituted pyrrolidone butanamides as new agents with significant antiepileptic activity. J. Med. Chem. 2004, 47, 530–549. [Google Scholar] [CrossRef]
  21. von Rosenstiel, P. Brivaracetam (ucb 34714). Neurotherapeutics 2007, 4, 84–87. [Google Scholar] [CrossRef]
  22. Zhu, J.; Mix, E.; Winblad, B. The antidepressant and antiinflammatory effects of rolipram in the central nervous system. CNS Drug Rev. 2001, 7, 387–398. [Google Scholar] [CrossRef]
  23. Gajcy, K.; Lochyński, S.; Librowski, T. A role of GABA analogues in the treatment of neurological diseases. Curr. Med. Chem. 2010, 17, 2338–2347. [Google Scholar] [CrossRef]
  24. Ordóñez, M.; Cativiela, C.; Romero-Estudillo, I. An update on the stereoselective synthesis of γ-amino acids. Tetrahedron Asymmetry 2016, 27, 999–1055. [Google Scholar] [CrossRef]
  25. Ramesh, P.; Suman, D.; Reddy, K.S.N. Asymmetric Synthetic Strategies of (R)-(–)-Baclofen: An Antispastic Drug. Synthesis 2018, 50, 211–226. [Google Scholar] [CrossRef]
  26. Martínez, C.A.; Hu, S.; Dumond, Y.; Tao, J.; Kelleher, P.; Tully, L. Development of a chemoenzymatic manufacturing process for pregabalin. Org. Process Res. Dev. 2008, 12, 392–398. [Google Scholar] [CrossRef]
  27. Shagufta, K.; Ameer, F.Z.; Sajjad, A.; Rabia, A.; Iqra, K.; Wajiha, Q.; Attia, M. Recent trends in the development of novel catalysts for asymmetric Michael reaction. Curr. Org. Chem. 2020, 24, 1397–1458. [Google Scholar]
  28. García-García, P.; Ladépêche, A.; Halder, R.; List, B. Catalytic asymmetric Michael reactions of acetaldehyde. Angew. Chem. Int. Ed. 2008, 47, 4719–4721. [Google Scholar] [CrossRef]
  29. Hayashi, Y.; Itoh, T.; Ohkubo, M.; Ishikawa, H. Asymmetric Michael reaction of acetaldehyde catalyzed by diphenylprolinol silyl ether. Angew. Chem. Int. Ed. 2008, 47, 4722–4724. [Google Scholar] [CrossRef]
  30. Gotoh, H.; Ishikawa, H.; Hayashi, Y. Diphenylprolinol silyl ether as catalyst of an asymmetric, catalytic, and direct Michael reaction of nitroalkanes with α, β-unsaturated aldehydes. Org. Lett. 2007, 9, 5307–5309. [Google Scholar] [CrossRef]
  31. Seebach, D.; Golinski, J. Synthesis of Open-Chain 2,3-Disubstituted 4-nitroketones by Diastereoselective Michael-addition of (E)-Enamines to (E)-Nitroolefins. A topological rule for C, C-bond forming processes between prochiral centres. Preliminary communication. Helv. Chim. Acta 1981, 64, 1413–1423. [Google Scholar] [CrossRef]
  32. Kaur, R.; Pandey, S.K. Efficient synthesis of (−)-(R)-and (+)-(S)-rolipram. Tetrahedron Lett. 2017, 58, 4333–4335. [Google Scholar] [CrossRef]
  33. Qiao, Y.; He, J.; Ni, B.; Headley, A.D. Asymmetric Michael reaction of acetaldehyde with nitroolefins catalyzed by highly water-compatible organocatalysts in aqueous media. Adv. Synth. Catal. 2012, 354, 2849–2853. [Google Scholar] [CrossRef]
  34. Nori, V.; Sinibaldi, A.; Giorgianni, G.; Pesciaioli, F.; Di Donato, F.; Cocco, E.; Biancolillo, A.; Landa, A.; Carlone, A. DoE-Driven Development of an Organocatalytic Enantioselective Addition of Acetaldehyde to Nitrostyrenes in Water. Chem. Eur. J. 2022, 28, e202104524. [Google Scholar] [CrossRef]
  35. Palomo, C.; Landa, A.; Mielgo, A.; Oiarbide, M.; Puente, A.; Vera, S. Water-compatible iminium activation: Organocatalytic Michael reactions of carbon-centered nucleophiles with enals. Angew. Chem. Int. Ed. 2007, 46, 8431–8435. [Google Scholar] [CrossRef]
  36. Alonso, D.A.; Baeza, A.; Chinchilla, R.; Gómez, C.; Guillena, G.; Pastor, I.M.; Ramón, D.R. Recent advances in asymmetric organocatalyzed conjugate additions to nitroalkenes. Molecules 2017, 22, 895. [Google Scholar] [CrossRef] [Green Version]
  37. Odagi, M.; Nagasawa, K. Recent Advances in Natural Products Synthesis Using Bifunctional Organocatalysts bearing a Hydrogen-Bonding Donor Moiety. Asian J. Org. Chem. 2019, 8, 1766–1774. [Google Scholar] [CrossRef]
  38. Okino, T.; Hoashi, Y.; Furukawa, T.; Xu, X.; Takemoto, Y. Enantio-and diastereoselective Michael reaction of 1, 3-dicarbonyl compounds to nitroolefins catalyzed by a bifunctional thiourea. J. Am. Chem. Soc. 2005, 127, 119–125. [Google Scholar] [CrossRef]
  39. Perlikowska, R.; Piekielna, J.; Mazur, M.; Koralewski, R.; Olczak, J.; do Rego, J.C.; Fichna, J.; Modranka, J.; Janecki, T.; Janecka, A. Antinociceptive and antidepressant-like action of endomorphin-2 analogs with proline surrogates in position 2. Bioorg. Med. Chem. 2014, 22, 4803–4809. [Google Scholar] [CrossRef]
  40. Sukhanova, A.A.; Nelyubina, Y.V.; Zlotin, S.G. Asymmetric synthesis of 3-prenyl-substituted pyrrolidin-2-ones. Mendeleev Commun. 2016, 26, 471–473. [Google Scholar] [CrossRef]
  41. Steppeler, F.; Dominika Iwan, D.; Elżbieta Wojaczyńska, E.; Wojaczyński, J. Chiral thioureas—Preparation and significance in asymmetric synthesis and medicinal chemistry. Molecules 2020, 25, 401. [Google Scholar] [CrossRef] [Green Version]
  42. Liu, J.; Wang, X.; Ge, Z.; Sun, Q.; Cheng, T.; Li, R. Solvent-free organocatalytic Michael addition of diethyl malonate to nitroalkenes: The practical synthesis of Pregabalin and γ-nitrobutyric acid derivatives. Tetrahedron 2011, 67, 636–640. [Google Scholar] [CrossRef]
  43. Bae, H.Y.; Song, C.E. Unprecedented Hydrophobic Amplification in Noncovalent Organocatalysis “on Water”: Hydrophobic Chiral Squaramide Catalyzed Michael Addition of Malonates to Nitroalkenes. ACS Catal. 2015, 5, 3613–3619. [Google Scholar] [CrossRef]
  44. Jin, H.; Kim, S.T.; Hwang, G.S.; Ryu, D.H. l-Proline Derived Bifunctional Organocatalysts: Enantioselective Michael Addition of Dithiomalonates to trans-β-Nitroolefins. J. Org. Chem. 2016, 81, 3263–3274. [Google Scholar] [CrossRef]
  45. Kimmel, K.L.; Weaver, J.D.; Ellman, J.A. Enantio-and diastereoselective addition of cyclohexyl Meldrum’s acid to β-and α, β-disubstituted nitroalkenes via N-sulfinyl urea catalysis. Chem. Sci. 2012, 3, 121–125. [Google Scholar] [CrossRef]
  46. Mondal, A.; Bhowmick, S.; Ghosh, A.; Chanda, T.; Bhowmick, K.C. Advances on asymmetric organocatalytic 1,4-conjugate addition reactions in aqueous and semi-aqueous media. Tetrahedron Asymmetry 2017, 28, 849–875. [Google Scholar] [CrossRef]
  47. Sim, J.H.; Park, J.H.; Maity, P.; Song, C.E. Access to Chiral GABA Analogues Bearing a Trifluoromethylated All-Carbon Quaternary Stereogenic Center through Water-Promoted Organocatalytic Michael Reactions. Org. Lett. 2019, 21, 6715–6719. [Google Scholar] [CrossRef]
  48. Leyva-Pérez, A.; García-García, P.; Corma, A. Multisite organic–inorganic hybrid catalysts for the direct sustainable synthesis of GABAergic drugs. Angew. Chem. Int. Ed. 2014, 53, 8687–8690. [Google Scholar] [CrossRef]
  49. García-García, P.; Zagdoun, A.; Copèret, C.; Lesage, A.; Díaz, U.; Corma, A. In situ preparation of a multifunctional chiral hybrid organic–inorganic catalyst for asymmetric multicomponent reactions. Chem. Sci. 2013, 4, 2006–2012. [Google Scholar] [CrossRef]
  50. Veverková, E.; Bilka, S.; Baran, R.; Šebesta, R. Squaramide-Catalyzed Michael Addition as a Key Step for the Direct Synthesis of GABAergic Drugs. Synthesis 2016, 48, 1474–1482. [Google Scholar]
  51. Marchetti, L.A.; Kumawat, L.K.; Mao, N.; Stephens, J.C.; Elmes, R.B.P. The versatility of squaramides: From supramolecular chemistry to chemical biology. Chem 2019, 5, 1398–1485. [Google Scholar] [CrossRef]
  52. Baran, R.; Veverková, E.; Škvorcová, A.; Šebesta, R. Enantioselective Michael addition of 1, 3-dicarbonyl compounds to a nitroalkene catalyzed by chiral squaramides—A key step in the synthesis of pregabalin. Org. Biomol. Chem. 2013, 11, 7705–7711. [Google Scholar] [CrossRef] [PubMed]
  53. Bassas, O.; Huuskonen, J.; Rissanen, K.; Koskinen, A.M.P. A simple organocatalytic enantioselective synthesis of pregabalin. Eur. J. Org. Chem. 2009, 2009, 1340–1351. [Google Scholar] [CrossRef] [Green Version]
  54. Kitanosono, T.; Kobayashi, S. Synthetic Organic “Aquachemistry” that Relies on Neither Cosolvents nor Surfactants. ACS Cent. Sci. 2021, 7, 739–747. [Google Scholar] [CrossRef]
  55. Bae, H.Y.; Some, S.; Oh, J.S.; Lee, Y.S.; Song, C.E. Hydrogen bonding mediated enantioselective organocatalysis in brine: Significant rate acceleration and enhanced stereoselectivity in enantioselective Michael addition reactions of 1,3-dicarbonyls to β-nitroolefins. Chem. Commun. 2011, 47, 9621–9623. [Google Scholar] [CrossRef] [PubMed]
  56. Sim, J.H.; Song, C.E. Water-Enabled Catalytic Asymmetric Michael Reactions of Unreactive Nitroalkenes: One-Pot Synthesis of Chiral GABA-Analogs with All-Carbon Quaternary Stereogenic Centers. Angew. Chem. Int. Ed. 2017, 56, 1835–1839. [Google Scholar] [CrossRef]
  57. Wang, Z.L. Recent advances in catalytic asymmetric decarboxylative addition reactions. Adv. Synth. Catal. 2013, 355, 2745–2755. [Google Scholar] [CrossRef]
  58. Bae, H.Y.; Some, S.; Lee, J.H.; Kim, J.Y.; Song, M.J.; Lee, S.; Zhang, Y.J.; Song, C.E. Organocatalytic Enantioselective Michael-Addition of Malonic Acid Half-Thioesters to β-Nitroolefins: From Mimicry of Polyketide Synthases to Scalable Synthesis of γ-Amino Acids. Adv. Synth. Catal. 2011, 353, 3196–3202. [Google Scholar] [CrossRef]
  59. Lubkoll, J.; Wennemers, H. Mimicry of Polyketide Synthases—Enantioselective 1, 4-Addition Reactions of Malonic Acid Half-Thioesters to Nitroolefins. Angew. Chem. Int. Ed. 2007, 46, 6841–6844. [Google Scholar] [CrossRef]
  60. Li, F.; Li, Y.Z.; Jia, Z.S.; Xu, M.H.; Tian, P.; Lin, G.Q. Biscinchona alkaloids as highly efficient bifunctional organocatalysts for the asymmetric conjugate addition of malonates to nitroalkenes at ambient temperature. Tetrahedron 2011, 67, 10186–10194. [Google Scholar] [CrossRef]
  61. Serdyuk, O.V.; Heckel, C.M.; Tsogoeva, S.B. Bifunctional primary amine-thioureas in asymmetric organocatalysis. Org. Biomol. Chem. 2013, 11, 7051–7071. [Google Scholar] [CrossRef] [Green Version]
  62. Tsakos, M.; Kokotos, C.G.; Kokotos, G. Primary Amine-Thioureas with Improved Catalytic Properties for “Difficult” Michael Reactions: Efficient Organocatalytic Syntheses of (S)-Baclofen, (R)-Baclofen and (S)-Phenibut. Adv. Synth. Catal. 2012, 354, 740–746. [Google Scholar] [CrossRef]
  63. Zheng, K.; Liu, X.; Feng, X. Recent advances in metal-catalyzed asymmetric 1,4-conjugate addition (ACA) of nonorganometallic nucleophiles. Chem. Rev. 2018, 118, 7586–7656. [Google Scholar] [CrossRef] [PubMed]
  64. Hui, C.; Pu, F.; Xu, J. Metal-Catalyzed Asymmetric Michael Addition in Natural Product Synthesis. Chem. Eur. J. 2017, 23, 4023–4036. [Google Scholar] [CrossRef]
  65. Evans, D.A.; Mito, S.; Seidel, D. Scope and mechanism of enantioselective Michael additions of 1,3-dicarbonyl compounds to nitroalkenes catalyzed by nickel (II)–diamine complexes. J. Am. Chem. Soc. 2007, 129, 11583–11592. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, K.; Jin, R.; Ceng, T.; Xu, X.; Gao, F.; Liu, G.; Li, H. Functionalized Periodic Mesoporous Organosilica: A Highly Enantioselective Catalyst for the Michael Addition of 1,3-Dicarbonyl Compounds to Nitroalkenes. Chem. Eur. J. 2012, 18, 15546–15553. [Google Scholar] [CrossRef]
  67. Bissessar, D.; Achard, T.; Bellemin-Laponnaz, S. Robust and Recyclable Self-Supported Chiral Nickel Catalyst for the Enantioselective Michael Addition. Adv. Synth. Catal. 2016, 358, 1982–1988. [Google Scholar] [CrossRef]
  68. Poe, S.L.; Kobašlija, M.; McQuade, D.T. Mechanism and application of a microcapsule enabled multicatalyst reaction. J. Am. Chem. Soc. 2007, 129, 9216–9221. [Google Scholar] [CrossRef]
  69. Ishitani, H.; Kanai, K.; Yoo, W.J.; Yoshida, T.; Kobayashi, S. A Nickel-Diamine/Mesoporous Silica Composite as a Heterogeneous Chiral Catalyst for Asymmetric 1, 4-Addition Reactions. Angew. Chem. Int. Ed. 2019, 58, 13313–13317. [Google Scholar] [CrossRef]
  70. Buendia, M.B.; Kegnæs, S.; Kramera, S. A Nickel-Bisdiamine Porous Organic Polymer as Heterogeneous Chiral Catalyst for Asymmetric Michael Addition to Aliphatic Nitroalkenes. Adv. Synth. Catal. 2020, 362, 5506–5512. [Google Scholar] [CrossRef]
  71. Li, X.; Li, X.; Peng, F.; Shao, Z. Mutually Complementary Metal- and Organocatalysis with Collective Synthesis: Asymmetric Conjugate Addition of 1,3-Carbonyl Compounds to Nitroenynes and Further Reactions of the Products. Adv. Synth. Catal. 2012, 354, 2873–2885. [Google Scholar] [CrossRef]
  72. Matsunaga, S.; Shibasaki, M. Recent advances in cooperative bimetallic asymmetric catalysis: Dinuclear Schiff base complexes. Chem. Commun. 2014, 50, 1044–1057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Furutachi, M.; Mouri, S.; Matsunaga, S.; Shibasaki, M. A Heterobimetallic Ni/La-salan Complex for Catalytic Asymmetric Decarboxylative 1,4-Addition of Malonic Acid Half-Thioester. Chem. Asian J. 2010, 5, 2351–2354. [Google Scholar] [CrossRef] [PubMed]
  74. Tsubogo, T.; Yamashita, Y.; Kobayashi, S. Toward efficient asymmetric carbon–carbon bond formation: Continuous flow with chiral heterogeneous catalysts. Chem. Eur. J. 2012, 18, 13624–13628. [Google Scholar] [CrossRef] [PubMed]
  75. Ishitani, H.; Furiya, Y.; Kobayashi, S. Enantioselective Sequential-Flow Synthesis of Baclofen Precursor via Asymmetric 1,4-Addition and Chemoselective Hydrogenation on Platinum/Carbon/Calcium Phosphate Composites. Chem. Asian. J. 2020, 15, 1688–1691. [Google Scholar] [CrossRef]
  76. Ishitani, H.; Saito, Y.; Tsubogo, T.; Kobayashi, S. Synthesis of Nitro-Containing Compounds through Multistep Continuous Flow with Heterogeneous Catalysts. Org. Lett. 2016, 18, 1346–1349. [Google Scholar] [CrossRef]
  77. Tsubogo, T.; Oyamada, H.; Kobayashi, S. Multistep continuous-flow synthesis of (R)-and (S)-rolipram using heterogeneous catalysts. Nature 2015, 520, 329–332. [Google Scholar] [CrossRef]
  78. Zadsirjan, V.; Heravi, M.M. Oxazolidinones as Chiral Auxiliaries in the Asymmetric 1,4-Conjugate Addition Reaction Applied to the Total Synthesis of Natural Products: A Supplemental Mini-Review. Curr. Org. Synth. 2018, 15, 3–20. [Google Scholar] [CrossRef]
  79. Tovar-Gudiño, E.; Morales-Nava, R.; Fernández-Zertuche, M. Diasteroselective conjugate addition of diethylaluminum cyanide to a conjugated N-enoyl system: An alternative synthesis of (S)-pregabalin. Can. J. Chem. 2014, 92, 45–48. [Google Scholar] [CrossRef]
  80. Armstrong, A.; Convine, N.J.; Popkin, M.E. Diastereoselective conjugate addition of cyanide to α, β-unsaturated oxazolidinones: Enantioselective synthesis of ent-pregabalin and baclofen. Synlett 2006, 2006, 1589–1591. [Google Scholar] [CrossRef]
  81. He, C.; Zhai, Z.; Zhou, Y.; Li, J.; Wang, G. A new synthetic route for the preparation of pregabalin. Synth. Commun. 2021, 51, 2034–2040. [Google Scholar] [CrossRef]
  82. Kuo, J.H.; Wong, W.C. Method for Preparing Chiral Baclofen. U.S. Patent 20,120,029,230, 25 September 2012. [Google Scholar]
  83. Izquierdo, S.; Aguilera, J.; Buschmann, H.H.; García, M.; Torrens, A.; Ortuño, R.M. Stereoselective and efficient synthesis of (S)-pregabalin from d-mannitol. Tetrahedron Asymmetry 2008, 19, 651–653. [Google Scholar] [CrossRef]
  84. Hayashi, Y. Domino and one-pot syntheses of biologically active compounds using diphenylprolinol silyl ether. Phys. Sci. Rev. 2020, 5, 20180088. [Google Scholar] [CrossRef]
  85. Donslund, B.S.; Johansen, T.K.; Poulsen, P.H.; Halskov, K.S.; Jørgensen, K.A. The diarylprolinol silyl ethers: Ten years after. Angew. Chem. Int. Ed. 2015, 54, 13860–13874. [Google Scholar] [CrossRef]
  86. Zu, L.; Xie, H.; Li, H.; Wang, J.; Wang, W. Highly enantioselective organocatalytic conjugate addition of nitromethane to α, β-unsaturated aldehydes: Three-step synthesis of optically active baclofen. Adv. Synth. Catal. 2007, 349, 2660–2664. [Google Scholar] [CrossRef]
  87. Wang, Y.; Li, P.; Liang, X.; Zhang, T.Y.; Ye, J. An efficient enantioselective method for asymmetric Michael addition of nitroalkanes to α, β-unsaturated aldehydes. Chem. Commun. 2008, 1232–1234. [Google Scholar] [CrossRef] [PubMed]
  88. Hayashi, Y.; Sakamoto, D.; Okamura, D. One-Pot Synthesis of (S)-Baclofen via Aldol Condensation of Acetaldehyde with Diphenylprolinol Silyl Ether Mediated Asymmetric Michael Reaction as a Key Step. Org. Lett. 2016, 18, 4–7. [Google Scholar] [CrossRef] [PubMed]
  89. Jensen, K.L.; Poulsen, P.H.; Donslund, B.S.; Morana, F.; Jorgensen, K.A. Asymmetric synthesis of γ-nitroesters by an organocatalytic one-pot strategy. Org. Lett. 2012, 14, 1516–1519. [Google Scholar] [CrossRef]
  90. Ötvös, S.B.; Kappe, C.O. Continuous flow asymmetric synthesis of chiral active pharmaceutical ingredients and their advanced intermediates. Green Chem. 2021, 23, 6117–6138. [Google Scholar] [CrossRef]
  91. Jiao, J.; Nie, W.; Yu, T.; Yang, F.; Zhang, Q.; Aihemaiti, F.; Yang, T.; Liu, X.; Wang, J.; Li, P. Multi-Step Continuous-Flow Organic Synthesis: Opportunities and Challenges. Chem. Eur. J. 2021, 27, 4817–4838. [Google Scholar] [CrossRef]
  92. Bloemendal, V.R.L.J.; Janssen, M.A.C.H.; van Hest, J.C.M.; Rutjes, F.P.J.T. Continuous one-flow multi-step synthesis of active pharmaceutical ingredients. React. Chem. Eng. 2020, 5, 1186–1197. [Google Scholar] [CrossRef]
  93. Nagy, B.S.; Llanes, P.; Pericas, M.A.; Kappe, O.; Ötvös, S.B. Enantioselective Flow Synthesis of Rolipram Enabled by a Telescoped Asymmetric Conjugate Addition–Oxidative Aldehyde Esterification Sequence Using in Situ-Generated Persulfuric Acid as Oxidant. Org. Lett. 2022, 24, 1066–1071. [Google Scholar] [CrossRef] [PubMed]
  94. Ötvös, S.B.; Llanes, P.; Pericás, M.A.; Kappe, O. Telescoped continuous flow synthesis of optically active γ-nitrobutyric acids as key intermediates of baclofen, phenibut, and fluorophenibut. Org. Lett. 2020, 22, 8122–8126. [Google Scholar] [CrossRef] [PubMed]
  95. Guo, X.T.; Shen, J.; Sha, F.; Wu, X.Y. Highly enantioselective Michael addition of nitroalkanes to enones and its application in syntheses of (R)-Baclofen and (R)-Phenibut. Synthesis 2015, 47, 2063–2072. [Google Scholar] [CrossRef]
  96. Ogawa, T.; Mouri, S.; Yazaki, R.; Kumagai, N.; Shibasaki, M. Intermediate as catalyst: Catalytic asymmetric conjugate addition of nitroalkanes to α, β-unsaturated thioamides. Org. Lett. 2012, 14, 110–113. [Google Scholar] [CrossRef]
  97. Shao, C.; Yu, H.J.; Wu, N.Y.; Tian, P.; Wang, R.; Feng, C.G.; Lin, G.Q. Asymmetric Synthesis of β-Substituted γ-Lactams via Rhodium/Diene-Catalyzed 1,4-Additions: Application to the Synthesis of (R)-Baclofen and (R)-Rolipram. Org. Lett. 2011, 13, 788–791. [Google Scholar] [CrossRef]
  98. Wang, Z.Q.; Feng, C.G.; Zhang, S.S.; Xu, M.H.; Lin, G.Q. Rhodium-Catalyzed Asymmetric Conjugate Addition of Organoboronic Acids to Nitroalkenes Using Chiral Bicyclo [3.3.0] Diene Ligands. Angew. Chem. Int. Ed. 2010, 49, 5780–5783. [Google Scholar] [CrossRef]
  99. Wang, Z.Q.; Feng, C.G.; Xu, M.H.; Lin, G.Q. Design of C2-Symmetric Tetrahydropentalenes as New Chiral Diene Ligands for Highly Enantioselective Rh-Catalyzed Arylation of N-Tosylarylimines with Arylboronic Acids. J. Am. Chem. Soc. 2007, 129, 5336–5337. [Google Scholar] [CrossRef]
  100. Han, F.; Chen, J.; Zhang, X.; Liu, J.; Cun, L.; Zhu, J.; Deng, J.; Liao, J. Rhodium-catalyzed asymmetric addition of arylboronic acids to γ-phthalimido-substituted-α, β-unsaturated carboxylic acid esters: An approach to γ-amino acids. Tetrahedron Lett. 2011, 52, 830–833. [Google Scholar] [CrossRef]
  101. Chen, J.; Chen, J.; Lang, F.; Zhang, X.; Cun, L.; Zhu, J.; Deng, J.; Liao, J. A C2-Symmetric Chiral Bis-Sulfoxide Ligand in a Rhodium-Catalyzed Reaction: Asymmetric 1,4-Addition of Sodium Tetraarylborates to Chromenones. J. Am. Chem. Soc. 2010, 132, 4552–4553. [Google Scholar] [CrossRef]
  102. Becht, J.M.; Meyer, O.; Helmchen, G. Enantioselective syntheses of (-)-(R)-rolipram,(-)-(R)-baclofen and other GABA analogues via rhodium-catalyzed conjugate addition of arylboronic acids. Synthesis 2003, 18, 2805–2810. [Google Scholar] [CrossRef]
  103. Diaz, A.; Siro, J.G.; García-Navío, J.L.; Vaquero, J.J.; Alvarez-Builla, J. A stereoselective synthesis of (R)-(-)-rolipram from L-glutamic acid. Synthesis 1997, 1997, 559–562. [Google Scholar] [CrossRef] [Green Version]
  104. Oba, M.; Saegusa, T.; Nishiyama, N.; Nishiyama, K. Synthesis of non-proteinogenic amino acids using Michael addition to unsaturated orthopyroglutamate derivative. Tetrahedron 2009, 65, 128–133. [Google Scholar] [CrossRef]
  105. Chang, M.Y.; Sun, P.P.; Chena, S.T.; Chang, N.C. A facile synthesis of 3-aryl pyroglutamic acid. Facile synthesis of baclofen and chlorpheg. Tetrahedron Lett. 2003, 44, 5271–5273. [Google Scholar] [CrossRef]
  106. Meyers, A.I.; Snyder, L. The synthesis of aracemic 4-substituted pyrrolidinones and 3-substituted pyrrolidines. An asymmetric synthesis of (-)-rolipram. J. Org. Chem. 1993, 58, 36–42. [Google Scholar] [CrossRef]
  107. Meyers, A.I.; Snyder, L. Asymmetric conjugate additions to chiral bicyclic lactams. Synthesis of aracemic trans-2,3-disubstituted pyrrolidines. J. Org. Chem. 1992, 57, 3814–3819. [Google Scholar] [CrossRef]
  108. Langlois, N.; Wang, H.S. Synthesis of the Novel Antidepressant (R)-(-)-Rolipram. Synth. Commun. 1997, 27, 3133–3144. [Google Scholar] [CrossRef]
  109. Momioka, K.; Sudani, M.; Shinmi, Y.; Koga, K. Enantioselective conjugate addition reaction mediated by chiral ligands. Chem. Lett. 1985, 14, 329–332. [Google Scholar]
  110. Massolo, E.; Benaglia, M.; Genoni, A.; Annunziata, R.; Celentano, G.; Gaggero, N. Stereoselective reaction of 2-carboxythioesters-1,3-dithiane with nitroalkenes: An organocatalytic strategy for the asymmetric addition of a glyoxylate anion equivalent. Org. Biomol. Chem. 2015, 13, 5591–5596. [Google Scholar] [CrossRef] [Green Version]
  111. Nakamura, S.; Tokunaga, A.; Saito, H.; Kondo, M. Enantioselective conjugate addition of an α, α-dithioacetonitrile with nitroalkenes using chiral bis (imidazoline)–Pd complexes. Chem. Commun. 2019, 55, 5391–5394. [Google Scholar] [CrossRef]
  112. Kondo, M.; Saito, H.; Nakamura, S. Direct catalytic enantioselective Mannich-type reaction of α, α-dithioacetonitriles with imines using chiral bis (imidazoline)–Pd complexes. Chem. Commun. 2017, 53, 6776–6779. [Google Scholar] [CrossRef]
  113. Baures, P.W.; Eggleston, D.S.; Erhard, K.F.; Cieslinski, L.B.; Torphy, T.J.; Christensen, S.B. Crystal structure, absolute configuration, and phosphodiesterase inhibitory activity of (+)-1-(4-bromobenzyl)-4-[(3-cyclopentyloxy)-4-methoxyphenyl]-2-pyrrolidinone. J. Med. Chem. 1993, 36, 3274–3277. [Google Scholar] [CrossRef] [PubMed]
  114. Pirnot, M.T.; Rankic, D.A.; Martin, D.B.C.; MacMillan, D.W.C. Photoredox activation for the direct β-arylation of ketones and aldehydes. Science 2013, 339, 1593–1596. [Google Scholar] [CrossRef]
  115. Fu, Z.; Xu, J.; Zhu, T.; Leong, W.W.Y.; Chi, Y.R. β-Carbon activation of saturated carboxylic esters through N-heterocyclic carbene organocatalysis. Nat. Chem. 2013, 5, 835–839. [Google Scholar] [CrossRef] [PubMed]
  116. Ma, J.; Zhou, Q.; Song, G.; Song, Y.; Zhao, G.; Ding, K.; Zhao, B. Enantioselective synthesis of pyroglutamic acid esters from glycinate via carbonyl catalysis. Angew. Chem. Int. Ed. 2021, 60, 10588–10592. [Google Scholar] [CrossRef]
  117. Sorochinsky, A.E.; Aceña, J.L.; Soloshonok, V.A. Self-Disproportionation of Enantiomers of Chiral, Non-Racemic Fluoroorganic Compounds: Role of Fluorine as Enabling Element. Synthesis 2013, 45, 141–152. [Google Scholar] [CrossRef]
  118. Sorochinsky, A.E.; Katagiri, T.; Ono, T.; Wzorek, A.; Aceña, J.L.; Soloshonok, V.A. Optical purifications via Self-Disproportionation of Enantiomers by achiral chromatography; Case study of a series of α-CF3-containing secondary alcohols. Chirality 2013, 25, 365–368. [Google Scholar] [CrossRef]
  119. Nakamura, T.; Tateishi, K.; Tsukagoshi, S.; Hashimoto, S.; Watanabe, S.; Soloshonok, V.A.; Aceña, J.L.; Kitagawa, O. Self-Disproportionation of Enantiomers of Non-racemic Chiral Amine Derivatives Through Achiral Chromatography. Tetrahedron 2012, 68, 4013–4017. [Google Scholar] [CrossRef]
  120. Suzuki, Y.; Han, J.; Kitagawa, O.; Aceña, J.L.; Klika, K.D.; Soloshonok, V.A. A comprehensive examination of the self-disproportionation of enantiomers (SDE) of chiral amides via achiral, laboratory-routine, gravity-driven column chromatography. RSC Adv. 2015, 5, 2988–2993. [Google Scholar] [CrossRef]
  121. Han, J.; Wzorek, A.; Kwiatkowska, M.; Soloshonok, V.A.; Klika, K.D. The self-disproportionation of enantiomers (SDE) of amino acids and their derivatives. Amino Acids 2019, 51, 865–889. [Google Scholar] [CrossRef]
  122. Hosaka, T.; Imai, T.; Wzorek, A.; Marcinkowska, M.; Kolbus, A.; Kitagawa, O.; Soloshonok, V.A.; Klika, K.D. The self-disproportionation of enantiomers (SDE) of α-amino acid derivatives; facets of steric and electronic properties. Amino Acids 2019, 51, 283–294. [Google Scholar] [CrossRef]
  123. Wzorek, A.; Sato, A.; Drabowicz, J.; Soloshonok, V.A.; Klika, K.D. Remarkable magnitude of the self-disproportionation of enantiomers (SDE) via achiral chromatography; application to the practical-scale enantiopurification of β-amino acid esters. Amino Acids 2016, 48, 605–613. [Google Scholar] [CrossRef] [PubMed]
  124. Han, J.; Kitagawa, O.; Wzorek, A.; Klika, K.D.; Soloshonok, V.A. The self-disproportionation of enantiomers (SDE): A menace or an opportunity? Chem. Sci. 2018, 9, 1718–1739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Soloshonok, V.A.; Wzorek, A.; Klika, K.D. A question of policy: Should tests for the self-disproportionation of enantiomers (SDE) be mandatory for reports involving scalemates? Tetrahedron Asymmetry 2017, 28, 1430–1434. [Google Scholar] [CrossRef]
  126. Han, J.; Nelson, D.J.; Sorochinsky, A.E.; Soloshonok, V.A. Self-Disproportionation of Enantiomers via Sublimation; New and Truly Green Dimension in Optical Purification. Curr. Org. Synth. 2011, 8, 310–317. [Google Scholar] [CrossRef]
  127. Han, J.; Wzorek, A.; Klika, K.D.; Soloshonok, V.A. Recommended Tests for the Self-Disproportionation of Enantiomers (SDE) to Ensure Accurate Reporting of the Stereochemical Outcome of Enantioselective Reactions. Molecules 2021, 26, 2757. [Google Scholar] [CrossRef]
  128. Han, J.; Dembinski, R.; Soloshonok, V.A.; Klika, K.D. A Call for a Change in Policy Regarding the Necessity for SDE Tests to Validate the Veracity of the Outcome of Enantioselective Syntheses, the Inherent Chiral State of Natural Products, and Other Cases Involving Enantioenriched Samples. Molecules 2021, 26, 3994. [Google Scholar] [CrossRef]
  129. Chi, Y.; Guo, L.; Kopf, N.A.; Gellman, S.H. Enantioselective Organocatalytic Michael Addition of Aldehydes to Nitroethylene: Efficient Access to γ2-Amino Acids. J. Am. Chem. Soc. 2008, 130, 5608–5609. [Google Scholar] [CrossRef] [Green Version]
  130. Ye, W.; Jiang, Z.; Zhao, Y.; Goh, S.L.M.; Leow, D.; Soh, Y.T.; Tan, C.H. Chiral Bicyclic Guanidine as a Versatile Brønsted Base Catalystfor the Enantioselective Michael Reactions of Dithiomalonatesandb-Ket oThi oesters. Adv. Synth. Catal. 2007, 349, 2454–2458. [Google Scholar] [CrossRef]
Figure 1. Representative β-substituted GABA derivatives with clinical applications.
Figure 1. Representative β-substituted GABA derivatives with clinical applications.
Molecules 27 03797 g001
Scheme 1. Organocatalytic asymmetric Michael reaction of acetaldehyde with nitroalkenes (E)-1 and (E)-2.
Scheme 1. Organocatalytic asymmetric Michael reaction of acetaldehyde with nitroalkenes (E)-1 and (E)-2.
Molecules 27 03797 sch001
Scheme 2. Synthetic approach of (R)-Rolipram 12 employing the organocatalyzed asymmetric Michael addition acetaldehyde to nitro olefin (E)-10.
Scheme 2. Synthetic approach of (R)-Rolipram 12 employing the organocatalyzed asymmetric Michael addition acetaldehyde to nitro olefin (E)-10.
Molecules 27 03797 sch002
Scheme 3. Enantioselective Michael reaction of (E)-1 with diethyl malonate in the presence of thiourea (R,R)-13.
Scheme 3. Enantioselective Michael reaction of (E)-1 with diethyl malonate in the presence of thiourea (R,R)-13.
Molecules 27 03797 sch003
Scheme 4. Enantioselective Michael reaction of (E)-2 with diethyl malonate in the presence of thiourea (R,R)-13 under solvent-free conditions.
Scheme 4. Enantioselective Michael reaction of (E)-2 with diethyl malonate in the presence of thiourea (R,R)-13 under solvent-free conditions.
Molecules 27 03797 sch004
Scheme 5. Enantioselective Michael addition of diphenyl dithiomalonates to β-nitrostyrene (E)-20 catalyzed by L-proline-derived urea (S,R)-21.
Scheme 5. Enantioselective Michael addition of diphenyl dithiomalonates to β-nitrostyrene (E)-20 catalyzed by L-proline-derived urea (S,R)-21.
Molecules 27 03797 sch005
Scheme 6. Catalytic enantioselective addition of cyclohexyl Meldrum’s acid to nitroalkene (E)-2 using N-sulfinyl urea catalysis (Ss,R,R)-27.
Scheme 6. Catalytic enantioselective addition of cyclohexyl Meldrum’s acid to nitroalkene (E)-2 using N-sulfinyl urea catalysis (Ss,R,R)-27.
Molecules 27 03797 sch006
Scheme 7. Synthesis of GABA derivatives with a trifluoromethylated quaternary stereogenic center.
Scheme 7. Synthesis of GABA derivatives with a trifluoromethylated quaternary stereogenic center.
Molecules 27 03797 sch007
Scheme 8. Asymmetric three-component reaction of aldehydes, nitromethane, and malonate.
Scheme 8. Asymmetric three-component reaction of aldehydes, nitromethane, and malonate.
Molecules 27 03797 sch008
Scheme 9. Michael addition of dimethyl malonate to β-nitrostyrene (E)-1 using squaramide organocatalysis (R,S)-37.
Scheme 9. Michael addition of dimethyl malonate to β-nitrostyrene (E)-1 using squaramide organocatalysis (R,S)-37.
Molecules 27 03797 sch009
Scheme 10. The Michael addition of Meldrum’s acid to nitroalkene (E)-2 catalyzed by squaramide (R,R)-39 catalyst.
Scheme 10. The Michael addition of Meldrum’s acid to nitroalkene (E)-2 catalyzed by squaramide (R,R)-39 catalyst.
Molecules 27 03797 sch010
Scheme 11. Enantioselective syntheses of (S)-Pregabalin 9 (A) and (S)-Rolipram 12 (B) under “on water” conditions.
Scheme 11. Enantioselective syntheses of (S)-Pregabalin 9 (A) and (S)-Rolipram 12 (B) under “on water” conditions.
Molecules 27 03797 sch011
Scheme 12. Enantioselective Michael addition reaction of β,β-disubstituted nitroalkenes (E)-47 with dithiomalonate using dihydroquinine-squaramide catalyst.
Scheme 12. Enantioselective Michael addition reaction of β,β-disubstituted nitroalkenes (E)-47 with dithiomalonate using dihydroquinine-squaramide catalyst.
Molecules 27 03797 sch012
Scheme 13. Enantioselective Michael addition of malonic-acid half-thioester 51 to β-nitrostyrene (E)-1. Synthesis of (S)-baclofen 8.
Scheme 13. Enantioselective Michael addition of malonic-acid half-thioester 51 to β-nitrostyrene (E)-1. Synthesis of (S)-baclofen 8.
Molecules 27 03797 sch013
Scheme 14. Asymmetric conjugate addition of dimethyl malonate to β-nitrostyrene (E)-1. Synthesis of (R)-Baclofen 8 hydrochloride.
Scheme 14. Asymmetric conjugate addition of dimethyl malonate to β-nitrostyrene (E)-1. Synthesis of (R)-Baclofen 8 hydrochloride.
Molecules 27 03797 sch014
Scheme 15. Michael reaction between acetophenone and β-nitrostyrenes using primary amine-thiourea organocatalyst 55.
Scheme 15. Michael reaction between acetophenone and β-nitrostyrenes using primary amine-thiourea organocatalyst 55.
Molecules 27 03797 sch015
Scheme 16. Enantioselective Michael additions of tert-butyl phenyl malonate to β-nitrostyrene (E)-20 catalyzed by diamine−Ni(II) complex 60.
Scheme 16. Enantioselective Michael additions of tert-butyl phenyl malonate to β-nitrostyrene (E)-20 catalyzed by diamine−Ni(II) complex 60.
Molecules 27 03797 sch016
Scheme 17. Proposed mechanism reaction.
Scheme 17. Proposed mechanism reaction.
Molecules 27 03797 sch017
Scheme 18. Asymmetric synthesis of β-alkynyl-γ-amino acid (R)-69 and β-alkyny-γ-lactam (R)-70.
Scheme 18. Asymmetric synthesis of β-alkynyl-γ-amino acid (R)-69 and β-alkyny-γ-lactam (R)-70.
Molecules 27 03797 sch018
Scheme 19. Ni–La system for decarboxylative 1,4-addition of malonic-acid half-thioester to nitroalkene (E)-10.
Scheme 19. Ni–La system for decarboxylative 1,4-addition of malonic-acid half-thioester to nitroalkene (E)-10.
Molecules 27 03797 sch019
Figure 2. Sequential flow synthesis of Baclofen precursor (3R,4S)-74.
Figure 2. Sequential flow synthesis of Baclofen precursor (3R,4S)-74.
Molecules 27 03797 g002
Scheme 20. Diastereoselective Michael addition of cyanide to α,β-unsaturated oxazolidinones.
Scheme 20. Diastereoselective Michael addition of cyanide to α,β-unsaturated oxazolidinones.
Molecules 27 03797 sch020
Scheme 21. Preparation of (S)-Pregabalin 9.
Scheme 21. Preparation of (S)-Pregabalin 9.
Molecules 27 03797 sch021
Scheme 22. Tetramethylguanidine-catalyzed diastereoselective conjugate addition of nitromethane to α,β-unsaturated oxazolidinone (R)-85.
Scheme 22. Tetramethylguanidine-catalyzed diastereoselective conjugate addition of nitromethane to α,β-unsaturated oxazolidinone (R)-85.
Molecules 27 03797 sch022
Scheme 23. Diastereoselective synthesis of (S)-Pregabalin 9 from 2,3-O-isopropylidene-D-glyceraldehyde (R)-87.
Scheme 23. Diastereoselective synthesis of (S)-Pregabalin 9 from 2,3-O-isopropylidene-D-glyceraldehyde (R)-87.
Molecules 27 03797 sch023
Scheme 24. O-TMS-Protected prolinol (R)-97 catalyzed enantioselective Michael addition of nitromethane with α,β-unsaturated aldehyde (E)-96.
Scheme 24. O-TMS-Protected prolinol (R)-97 catalyzed enantioselective Michael addition of nitromethane with α,β-unsaturated aldehyde (E)-96.
Molecules 27 03797 sch024
Scheme 25. Application of the one-pot process for the formation of γ-nitroester (S)-100 in the synthesis of (S)-Baclofen 8.
Scheme 25. Application of the one-pot process for the formation of γ-nitroester (S)-100 in the synthesis of (S)-Baclofen 8.
Molecules 27 03797 sch025
Scheme 26. Continuous-flow enantioselective synthesis of (S)-Rolipram 12.
Scheme 26. Continuous-flow enantioselective synthesis of (S)-Rolipram 12.
Molecules 27 03797 sch026
Scheme 27. Squaramide (S)-104-catalyzed enantioselective Michael addition of nitromethane with α,β-unsaturated aldehyde (E)-96.
Scheme 27. Squaramide (S)-104-catalyzed enantioselective Michael addition of nitromethane with α,β-unsaturated aldehyde (E)-96.
Molecules 27 03797 sch027
Scheme 28. Enantioselective Michael addition of nitromethane to alkyl cinnamyl ketones and its application in syntheses of (R)-Phenibut 24 and (R)-Baclofen 8.
Scheme 28. Enantioselective Michael addition of nitromethane to alkyl cinnamyl ketones and its application in syntheses of (R)-Phenibut 24 and (R)-Baclofen 8.
Molecules 27 03797 sch028
Scheme 29. Catalytic asymmetric conjugate addition of nitromethane to α,β-unsaturated thioamide 109.
Scheme 29. Catalytic asymmetric conjugate addition of nitromethane to α,β-unsaturated thioamide 109.
Molecules 27 03797 sch029
Scheme 30. Rhodium-catalyzed asymmetric 1,4-addition reaction of arylboronic acids to α,β-unsaturated γ-lactams 112.
Scheme 30. Rhodium-catalyzed asymmetric 1,4-addition reaction of arylboronic acids to α,β-unsaturated γ-lactams 112.
Molecules 27 03797 sch030
Scheme 31. Synthesis of (R)-Baclofen 8 and (R)-Rolipram 12.
Scheme 31. Synthesis of (R)-Baclofen 8 and (R)-Rolipram 12.
Molecules 27 03797 sch031
Scheme 32. Rh-catalyzed asymmetric 1,4-addition of arylboronic acids to alkyl γ-phatalimidocrotonates 116.
Scheme 32. Rh-catalyzed asymmetric 1,4-addition of arylboronic acids to alkyl γ-phatalimidocrotonates 116.
Molecules 27 03797 sch032
Scheme 33. Synthesis of (S)-Rolipram 12 and (S)-Baclofen 8.
Scheme 33. Synthesis of (S)-Rolipram 12 and (S)-Baclofen 8.
Molecules 27 03797 sch033
Scheme 34. Enantioselective synthesis of (R)-Rolipram 12 and (R)-Baclofen 8 via Rh-catalyzed conjugate addition of arylboronic acids.
Scheme 34. Enantioselective synthesis of (R)-Rolipram 12 and (R)-Baclofen 8 via Rh-catalyzed conjugate addition of arylboronic acids.
Molecules 27 03797 sch034
Scheme 35. Stereoselective synthesis of (R)-Rolipram 12 from L-glutamic acid.
Scheme 35. Stereoselective synthesis of (R)-Rolipram 12 from L-glutamic acid.
Molecules 27 03797 sch035
Scheme 36. Preparation of (R)-Baclofen 8 using Michael addition to unsaturated pyroglutamate ester (S)-124.
Scheme 36. Preparation of (R)-Baclofen 8 using Michael addition to unsaturated pyroglutamate ester (S)-124.
Molecules 27 03797 sch036
Scheme 37. Asymmetric route to (R)-Rolipram 12 by addition of cuprates to chiral bicyclic lactam 130.
Scheme 37. Asymmetric route to (R)-Rolipram 12 by addition of cuprates to chiral bicyclic lactam 130.
Molecules 27 03797 sch037
Scheme 38. Synthesis of (R)-rolipram 12 using oxazolidines as chiral auxiliary.
Scheme 38. Synthesis of (R)-rolipram 12 using oxazolidines as chiral auxiliary.
Molecules 27 03797 sch038
Scheme 39. Organocatalytic 1,3-dithiane addition to styrene nitroalkenes. Synthesis of β-nitro esters and (S)-Baclofen 8.
Scheme 39. Organocatalytic 1,3-dithiane addition to styrene nitroalkenes. Synthesis of β-nitro esters and (S)-Baclofen 8.
Molecules 27 03797 sch039
Scheme 40. Proposed stereoselection model.
Scheme 40. Proposed stereoselection model.
Molecules 27 03797 sch040
Scheme 41. Enantioselective reaction of α,α-dithioacetonitrile 140 with nitroalkenes.
Scheme 41. Enantioselective reaction of α,α-dithioacetonitrile 140 with nitroalkenes.
Molecules 27 03797 sch041
Scheme 42. Plausible reaction mechanism for the reaction of α,α-dithioacetonitrile 140 with nitroalkenes using Phebim-PdBr 141.
Scheme 42. Plausible reaction mechanism for the reaction of α,α-dithioacetonitrile 140 with nitroalkenes using Phebim-PdBr 141.
Molecules 27 03797 sch042
Scheme 43. Synthesis of optically active (R)-Rolipram 12.
Scheme 43. Synthesis of optically active (R)-Rolipram 12.
Molecules 27 03797 sch043
Scheme 44. β-Carbon activation of saturated carboxylic esters.
Scheme 44. β-Carbon activation of saturated carboxylic esters.
Molecules 27 03797 sch044
Scheme 45. Reactivity of the ester β-carbon with different electrophiles.
Scheme 45. Reactivity of the ester β-carbon with different electrophiles.
Molecules 27 03797 sch045
Scheme 46. Application to the synthesis of bioactive molecules.
Scheme 46. Application to the synthesis of bioactive molecules.
Molecules 27 03797 sch046
Scheme 47. Plausible mechanism for the catalytic ester β-activation.
Scheme 47. Plausible mechanism for the catalytic ester β-activation.
Molecules 27 03797 sch047
Scheme 48. Enantioselective synthesis of pyroglutamic-acid esters from NH2-unprotected glycinate using carbonyl catalysis.
Scheme 48. Enantioselective synthesis of pyroglutamic-acid esters from NH2-unprotected glycinate using carbonyl catalysis.
Molecules 27 03797 sch048
Scheme 49. Proposed reaction mechanism.
Scheme 49. Proposed reaction mechanism.
Molecules 27 03797 sch049
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, J.; Escorihuela, J.; Fustero, S.; Landa, A.; Soloshonok, V.A.; Sorochinsky, A. Asymmetric Michael Addition in Synthesis of β-Substituted GABA Derivatives. Molecules 2022, 27, 3797. https://doi.org/10.3390/molecules27123797

AMA Style

Han J, Escorihuela J, Fustero S, Landa A, Soloshonok VA, Sorochinsky A. Asymmetric Michael Addition in Synthesis of β-Substituted GABA Derivatives. Molecules. 2022; 27(12):3797. https://doi.org/10.3390/molecules27123797

Chicago/Turabian Style

Han, Jianlin, Jorge Escorihuela, Santos Fustero, Aitor Landa, Vadim A. Soloshonok, and Alexander Sorochinsky. 2022. "Asymmetric Michael Addition in Synthesis of β-Substituted GABA Derivatives" Molecules 27, no. 12: 3797. https://doi.org/10.3390/molecules27123797

Article Metrics

Back to TopTop