Next Article in Journal
Effects of Depth-Based Object Isolation in Simulated Retinal Prosthetic Vision
Next Article in Special Issue
An Application of the Eigenproblem for Biochemical Similarity
Previous Article in Journal
Novel Coordination Mode in the Potassium Mefenamate Trihydrate Polymeric Structure
Previous Article in Special Issue
The Kinetic Model of Diffusion and Reactions in Powder Catalysts during Temperature Programmed Oxygen Isotopic Exchange Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Selected Asymmetric Reactions Promoted by Chiral Catalysts: Cyclopropanations, Friedel–Crafts, Mannich, Michael and Other Zinc-Mediated Processes—An Update

by
Michał Rachwalski
*,
Aleksandra Buchcic-Szychowska
and
Stanisław Leśniak
Department of Organic and Applied Chemistry, University of Lodz, Tamka 12, PL-91-403 Lodz, Poland
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(10), 1762; https://doi.org/10.3390/sym13101762
Submission received: 24 August 2021 / Revised: 15 September 2021 / Accepted: 16 September 2021 / Published: 22 September 2021
(This article belongs to the Collection Feature Papers in Chemistry)

Abstract

:
The main purpose of this review article is to present selected asymmetric synthesis reactions in which chemical and stereochemical outcomes are dependent on the use of an appropriate chiral catalyst. Optically pure or enantiomerically enriched products of such transformations may find further applications in various fields. Among an extremely wide variety of asymmetric reactions catalyzed by chiral systems, we are interested in: asymmetric cyclopropanation, Friedel–Crafts reaction, Mannich and Michael reaction, and other stereoselective processes conducted in the presence of zinc ions. This paper describes the achievements of the above-mentioned asymmetric transformations in the last three years. The choice of reactions is related to the research that has been carried out in our laboratory for many years.

1. Introduction

The synthesis of optically pure or enantiomerically enriched compounds has become one of the most important areas of interest in recent decades in modern synthetic organic chemistry [1]. Substances of high optical purity are still used in many industrial sectors, including the pharmaceutical and food industries. Chemists working in both academia and industry must agree with the fact that single chirality plays a huge role in nature and that biological properties are closely related to the absolute configuration of the product [2]. The opposite enantiomers act differently on living organisms and may exhibit different activities. Some of the differences are huge, ranging from different tastes and smells to even teratogenic (fetal-damaging) effects [3]. Between 1958 and 1962, thalidomide-racemic sedative drug, taken by pregnant women, damaged more than 10,000 fetuses worldwide [4]. This drastic example seems to be sufficient proof that the synthesis of optically pure substances used especially in the pharmaceutical and food industries is still gaining more and more importance.
There are three main strategies for the synthesis of enantiomerically pure substances: the first one is a methodology based on the use of naturally occurring substances with defined absolute configuration, the second resolution of racemic mixtures (kinetic resolution and deracemization techniques) [5], and finally, asymmetric synthesis in which one or more chiral centers are generated in the substrate. The subject of this review article is closely related to the last methodology, i.e., asymmetric synthesis. There are several approaches used in asymmetric synthesis; however, we focused on the use of chiral catalysts. The large number of asymmetric reactions studied and the dynamically developing synthesis of new chiral catalysts made the preparation of this review article quite difficult. However, we have used the knowledge derived from the research conducted in our laboratory. Thanks to our previous experience with asymmetric cyclopropanation reactions [6], Friedel–Crafts alkylation [7], Mannich [8,9] and Michael [10] reactions, and other asymmetric transformations performed in the presence of zinc ions [11], we decided to describe recent achievements in the field of these asymmetric transformations.

2. Asymmetric Cyclopropanation Reactions

Asymmetric cyclopropanation is currently one of the most used methods for creating new C-C bonds, as a cyclopropane ring having a unique steric and electronic properties is a structural motif present in many biologically active molecules, e.g., Cabozantinib, Simeprevir, Lumacaftor, etc. [12]. Some synthetic approaches with or without transition metals to achieve cyclopropyl ring have been reviewed in 2018 [13]. Moreover, some examples illustrating relationships between pharmacological activity and structure of various cyclopropyl derivatives have also been reported in 2018 [14].
Dictyopterenes are chemical compounds occurring in marine and freshwater environments and being sexual attractants or pheromones [15]. The asymmetric synthesis of Dictyopterene C’ 5 and its derivatives was performed using asymmetric cyclopropanation as one of the steps. α,β-Unsaturated aldehydes 1 and α-substituted α-diazoesters 2 took part in the reaction promoted by a chiral oxazaborolidinium ion catalyst (COBI) 3 (Scheme 1) [16]. Further transformations of cyclopropyl derivatives of type 4 including Julia–Kocienski reaction and Sonogashira and Suzuki coupling led to the corresponding dictyopterenes 5 [16].
Asymmetric Simmons–Smith cyclopropanation using diiodomethane and diethylzinc was applied in multi-step synthesis of the core of solanoeclepin A 7 (Scheme 2) which is a hatching agent of potato cyst nematodes [17]. Tricyclo [5.2.1.0] decane skeleton of solanoeclepin A was constructed starting from (R)-seudenol 6 (Scheme 2) [17].
Two further literature reports on the asymmetric cyclopropanation reaction include both experimental and theoretical studies [18,19]. In the first one, α-fluoroacrylates 8 were subjected to enantioselective cyclopropanation promoted by chiral rhodium catalysts 9 (Scheme 3) [18]. The corresponding cyclopropyl junctions 10 were formed with excellent enantio- and diastereoselectivities [18]. Monofluorinated cyclopropane derivatives have inter alia, antiviral (anti-herpetic) and antibacterial properties; the presence of fluorine and cyclopropane increases the bioavailability, selectivity and similarity to the respective receptors.
Chiral rhodium catalysts were also applied in stereoselective cyclopropanation of N-phenoxylsulfonamides 11 with cyclopropenyl secondary alcohols 12 (Scheme 4) [19]. A very wide spectrum of substrates was tested and, in most cases, the desired products were obtained in high chemical yields with excellent enantioselectivities and diastereoselectivity [19]. This type of cyclopropane moiety is the structural basis of drugs such as milnacipran and (+)-Tranylcypromine.
An original and interesting approach to the cyclopropanation process was reported by Breinbauer et al. The enzymes ene reductases were used as catalysts in the reaction of reduction in carbon-carbon double bonds in α,β-unsaturated compounds 13 with electron-withdrawing group [20]. It turned out that this type of enzyme also catalyzes the process of creating carbon-carbon bonds; thus, the aforementioned α,β-unsaturated systems underwent reductive cyclization. This new enzymatic transformation allows access to chiral cyclopropyl derivatives with excellent ee (up to 99%) (Scheme 5) [20].

3. Asymmetric Friedel–Crafts Reactions

Asymmetric Friedel–Crafts reaction is another widely used method for the enantioselective construction of carbon-carbon bonds [21]. The title transformation is often used in the synthesis of biologically active compounds [22] and some organocatalytic approaches have been compiled in a review article in 2019 [23]. Among the extremely wide variety of chiral sources used in asymmetric Friedel–Crafts reaction, in this review, special attention is paid to systems such as chiral derivatives of urea [24], thiourea [24], squaramide, chiral phosphoric acids and various complexes of N-heterocycles with metals.
Herrera et al. studied the activation of urea- and thiourea-derived catalysts in an asymmetric Friedel–Crafts reaction of indole 14 and nitrostyrene 15 as model substrates (Scheme 6) [25,26].
In the first contribution authors showed that the addition of an external Brönsted acid (rac-mandelic acid) enhanced the catalytic activity of the urea 16 system, increasing the enantioselectivity to 68% ee [25]. In the second study, chiral thiourea derivatives were activated with metals, leading to the best results in terms of enantioselectivity for the gold (I) complex of thiourea 17 [26].
An asymmetric alkylation of sulfonylindoles of type 18 with naphthol and its derivatives was performed by Han and co-workers using bifunctional catalysts bearing urea or a thiourea motif [27]. The corresponding products of type 20 were formed in the highest yields and enantiomeric excesses using thiourea 19 (Scheme 7). Additionally, among many solvents tested, the authors showed that the two-phase water–toluene system is the best one [27]. Taking into account the structure of compound 18, the proposed method may be useful in the synthesis of such natural compounds as (+)-gliocladin C, (+)-bionectin A or leptosin D.
A similar Friedel–Crafts transformation consisting in arylation of 1-azadienes of type 21 was reported in 2020 [28]. The desired product 22 was obtained in very high chemical yield (95%), and with an excellent enantiomeric ratio (97:3), when the urea-analog of system 19 (19a) was applied as catalyst (Scheme 8) [28]. The same reaction was investigated using thiourea bifunctional catalyst 19 which led to the appropriate chiral product with a moderate yield (72%) and enantioselectivity (62% of ee) [29]. It should be stressed that chiral triarylmethanes are the structural skeleton of many systems that exhibit antiviral and antibacterial effects, as well as a drug for tuberculosis or skin diseases.
Friedel–Crafts alkylation/cyclization of 4-hydroxyindole 23 promoted by bifunctional tertiary amine-urea catalyst 24 was reported by Lin and Duan et al. [30]. The corresponding spirooxindole-pyranoindole products of type 25 were formed very efficiently in terms of yield and enantioselectivity (Scheme 9) [30]. Indole derivatives are found in many drugs, while systems structurally similar to compound 25 have predominantly anti-malarial activity (e.g., NID609).
Friedel–Crafts functionalization of 4-hydroxyindole 23 with (E)-2-nitroallylic acetate 26 was successfully carried out using bifunctional squaramide catalyst 27 (Scheme 10) [31]. Chiral product 28 was formed in 86% yield, with over 99% of ee and diastereomeric ratio over 20:1 when system 27 was applied in DCM at 50 °C in the presence of dipotassium hydrogen phosphate and 4 Å molecular sieves [31]. It should be mentioned that thiourea catalysts showed weaker catalytic activity in this process [31]. As with other indole systems, it is present in many drugs and natural compounds.
Furo [2,3-b] benzofuranones were synthesized via efficient asymmetric Friedel–Crafts alkylation/hemiketalization/lactonization cascade reaction [32]. The search for new methods of the synthesis of chiral furo [2,3-b] benzofuranones is desirable due to the high biological (e.g., antibacterial, anti-thrombotic) activity of such systems. In the model reaction, 3-ylidene oxindole 29 was treated by 2-naphthol. The corresponding product 31 was obtained in the highest yield (83%) and enantioselectivity (90% of ee) in toluene when squaramide 30 was used as chiral promoter (Scheme 11). Thiourea derivatives were also tested but the results were not as significant as for squaramide [32].
Porous carbon nanosheet-supported squaramide 30 catalysts were successfully used for highly enantioselective Friedel–Crafts reaction between isatin ketimine 32 and pyrazolone 33 (Scheme 12) [33]. The best results in terms of yield (94%) and ee (up to 99%) were achieved after 15 min in DCM in the presence of potassium carbonate and N-fluorobenzenesulfonimide (NFSI) [33].
Chiral phosphoric acids acted as catalysts in asymmetric Friedel–Crafts alkylation of indoles with trifluoromethyl ketone hydrates 34 bearing a benzothiazole moiety (Scheme 13) [34]. The most successful transformation was achieved in case of the use of 35 as the chiral catalyst [34]. In this synthesis, special emphasis was placed on the chiral α-trifluoromethyl tertiary alcohol, which is present in many Anti-HIV drugs.
Similar catalysts were examined in the asymmetric Friedel–Crafts reaction of indoles with in situ generated N-acyl imines from alanine-derived N-(acyloxy) phthalimide 36 (Scheme 14) [35]. Reactions were performed under blue LED light in the presence of ruthenium or iridium photocatalysts [35]. The most efficient model reaction (86% yield, 90% ee) was carried out in the presence lithium salt 37.
Enantioselective Friedel–Crafts reaction of cyclic N-sulfimines 38 with 2-methoxyfurane 39 catalyzed by chiral BINOL-derived phosphoric acids was described in 2019 (Scheme 15) [36]. The application of chiral derivative 40 in o-xylene at room temperature gave the best results in terms of chemical yield and enantiomeric ratio [36]. The asymmetric FC reaction in which furan is the substrate is less understood as compared to indole. The furan ring is present in many natural compounds: (-)-furodysinin, fraxinellone (can be used as a pesticide but is non-toxic to mammals) or lophotoxin (it is an antagonist of the nicotinic acetylcholine receptor).
BINOL-derived chiral phosphoric acids are among the most widely used phosphorus derivatives to catalyze the asymmetric Friedel–Crafts reaction [37]. In addition to the reaction cited above, other selected examples of application of such catalysts are: alkylation of indoles with β-substituted cyclopenteneimines leading to β-indolyl cyclopentanones [38], asymmetric arylation of 2,2,2-trifluoroacetophenones [39], reaction of indoles with α-iminophosphonates [40], tandem three-component reaction of indole with aldehydes and p-toluenesulfonamide [41] and reaction of indoles with 3-indolylsulfamidates [42].
Beletskaya et al. described asymmetric Friedel–Crafts reactions of indoles with arylidene malonates 41 in the presence of magnesium iodide [43] and with phthaloyl-protected aminomethylenemalonate 43 in the presence of Cu(OTf)2 [44]. The processes were catalyzed by PyBox [43] or iPrBox [44], respectively (Scheme 16).
The asymmetric reaction of indoles with α,β-unsaturated carbonyl compounds 45 (enones) was carried out using three similar catalytic systems (Scheme 17) [45,46,47]. In the first contribution, reaction was promoted by imidazoline-oxazoline 46 complex with Cu(OTf)2 (acetonitrile, room temperature), leading to indole derivatives acting as novel α-glucosidase inhibitors in vitro [45]. The second approach comprises a utilization of chiral N,N-dioxide 47-scandium(III) complexes in dichloromethane at 35 °C [46], and finally, the third work relies on the application of cationic aqua complex of 2,2′-bypiridine 48 with palladium(II) in water at room temperature [47].
Chiral bis (3-indolyl) methanes 51 were synthesized in high yields and in a highly enantioselective manner via asymmetric Friedel–Crafts alkylation of indoles with trifluoromethylated nitroalkenes 49 (Scheme 18) [48]. Reactions were promoted by a nickel(II) complex of imidazoline oxazoline 50. Systems of type 51 have antibacterial and antifungal properties, inhibit the growth of cancer cells, and are also colorimetric sensors.
Asymmetric Friedel–Crafts reactions of indoles with trifluoromethyl pyruvate 52 were independently described in 2018 [49,50]. Kitamura et al. applied complexes of BOX-type 53 with Cu(OTf)2 in cyclopentyl methyl ether (CPME) at 0 °C [49]. In turn, Wang and Yang et al. carried out this process using Trost’s dinuclear zinc-ProPhenol catalyst 54 in dichloromethane at 10 °C in the presence of diethylzinc (Scheme 19) [50].
Similar dinuclear zinc catalytic systems were also used in the synthesis of 2,5-pyrrolidinyl dispirooxindoles [51] and tetrahydrofuran spirooxindoles [52] via cascade reactions, where one of the steps is a Friedel–Crafts process. Moreover, this type of chiral catalyst was successfully utilized for Friedel–Crafts reaction of pyrrole with chalcones [53].
The organocatalytic Friedel–Crafts alkylation of phloroglucinol derivatives 55 with enals 56 was described by Zu et al. [54]. The corresponding chiral products were afforded in high yields and enantioselectivities (Scheme 20) when diphenylprolinol TMS ether 57 was applied as the catalyst. This reaction opens up access to total synthesis of (+)-aflatoxin B2 [54].
The asymmetric propargylation of 3-substituted indoles via Friedel–Crafts process catalyzed by Ni(cod)2 and (R)-BINAP afforded the corresponding products bearing internal alkyne group with excellent enantio- and diastereoselectivities (Scheme 21) [55].
At the end of this chapter, it is worth mentioning a few more non-obvious approaches to the asymmetric Friedel–Crafts reaction. They are for sure the use of N-heterocyclic carbenes as catalysts in the synthesis of indole-fused polycyclic alcohols [56], guanosine-based self-assembly as catalytic system [57], and various DNA-employing catalytic systems [58,59,60].

4. Asymmetric Mannich Reactions

Asymmetric Mannich reaction is another process of enantioselective formation of carbon-carbon bonds, which can result in the formation of useful building blocks and compounds exhibiting biological activity.
Bifunctional thiourea systems were employed to the reaction between 3-indolinone-2-carboxylates 58 to N-Boc-benzaldimines generated in situ from α-amidosulfones 59 [61]. The best results in terms of chemical yield, enantio- and diastereoselectivity were obtained with the use of thiourea 60 (Scheme 22) [61]. Chiral derivatives of isatins occur in the nature; these are compounds such as (+)-isatisine A, trigonoliimine or mersicarpine. All three exhibit antiviral properties.
4-Substituted isoxazolidin-5-ones 61 were reacted with isatin-derived imines in the presence of (DHQD)2PYR ligand 62 (Scheme 23) [62]. Reactions performed in methyl tert-butyl ether (MTBE) at 0 °C gave the corresponding Mannich products in almost quantitative yields, and excellent enantioselectivity (99%) and diastereoselectivity (20:1 of d.r.). The obtained products are building blocks for the synthesis of systems with bactericidal properties.
Catalytic asymmetric Mannich reaction of N-Boc-aldimines with α,β-unsaturated pyrazoleamides 63 was successfully catalyzed by the complex of copper(I)-(R)-DTBM-SEGPHOS 64 in the presence of triethylamine leading to the corresponding syn-vinylogous products in a highly enantioselective and diastereoselective manner (Scheme 24) [63].
One of the most common chiral catalysts used in the asymmetric Mannich reaction, namely L-proline, was combined with the corresponding photocatalysts and used in enantioselective Mannich reaction of dihydroquinoxalinones 65 with ketones [64] and in the synthesis of C2-quaternary indolin-3-ones [65]. The dihydroquinoxalinone skeleton is found in many natural and synthetic compounds, which are drugs and plant protection products. In the first work, chiral quinoxaline junctions 66 were obtained in high yields (up to 94%) and very high enantioselectivities (up to 99% of ee) (Scheme 25) [64]. Dihydroquinoxalin-2-one 65 under the irradiation of visible light and in the presence of Eosin Y and oxygen is oxidized to quinoxalin-2(1H)-one (imine).
In the second case, Guan and He et al. also showed excellent catalytic activity of L-proline in the presence of ruthenium–bipirydyl complex (Scheme 26) [65].
Chiral phosphoric acid 69 was employed as catalysts in asymmetric Mannich reaction of cyclic C-acylimines 67 with ketones (Scheme 27) [66]. C2-quaternary indolin-3-ones 68 were formed with excellent values of chemical yield and stereoselectivity [66].

5. Asymmetric Michael Reactions

The Michael addition reaction constitutes one of the most powerful tools of construction of C-C bonds in modern synthetic organic chemistry. The number of different Michael donors and acceptors is large and constantly increasing; therefore, this reaction is still the subject of numerous studies. The Michael reaction in the asymmetric version is often a key step in the synthesis of many natural products [67]. The search for new chiral catalysts for this asymmetric transformation is still a challenge for various research groups [68]. Many recent reports inform about the use of bifunctional organocatalysts, such as thiourea–tertiary amine systems [69]. Moreover, Michael additions performed in the presence of zinc ions also enjoy quite a lot of interest [70]. Currently, target materials in the asymmetric Michael addition reaction are very often isoxazole-5-ones and isoxazolidin-5-ones [71], and pyrazolones [72].
Enantioselective 1,6-aza-Michael addition reaction of 4(H)-isoxazol-5-ones 70 to p-quinone methides 71 promoted by various organocatalysts was described by Blay and Pedro et al. [73]. The isoxazol-5-one ring is present in many natural compounds, and there are many examples of compounds from this group that are inhibitors of various enzymes. The properties of this ring mean that some derivatives are successfully used in photonics. The desired isoxazolin-5-one derivatives 72 were formed with the greatest efficiency in dichloroethane (DCE) at room temperature in the presence of 3Å molecular sieves under action of thiourea system 73 (Scheme 28) [73].
Spirooxindole γ-lactones bearing the CF3 substituent were synthesized using organocatalytic asymmetric Michael/lactonization cascade [74]. Spiroxoindole lactones have activities such as antibacterial, anti-biofilm or inhibition of (THF)-α-induced apoptosis. 3-Hydroxyindoles 74 reacted with 3-trifluoroethylidene oxindoles 75 in DCM/MeCN mixture at room temperature giving the desired lactones 76 in up to 97% yield, with up to 98:2 of d.r., and with up to 98% ee under action of chiral squaramide derivative 77 (Scheme 29) [74].
Highly efficient asymmetric Michael addition of pyrazol-5-ones 78 to β-trifluoromethyl-α,β-unsaturated ketones 79 was reported by Chang et al. (Scheme 30) [75]. Many systems containing a trifluoromethyl group at the stereogenic center are pharmacophores, e.g., Befloxatone is a monoamine oxidase inhibitor, odanacat-ib is a potential drug for osteoporosis by inhibiting the enzyme cathepsin K leading to bone reduction. The highest levels of enantioselectivity and diastereoselectivity of the reaction were achieved using dipeptide-based urea-amide catalyst 80 in chlorobenzene at 25 °C (Scheme 30) [75].
Cinchona-derived aminocatalysts of types 81–82 (Figure 1) were successfully applied in the four-component cycloaddition reaction of 3-substituted 2-cyclopentenones with isoxazole-5-one derivatives [76], and in the Michael reaction of arylidene-isoxazol-5-one with 1,3-diesters [77], respectively.
Meldrum’s acid constitutes a very useful tool in asymmetric organocatalysis including Michael reactions [78]. Chiral P,N-ligands of type 85 were used in the copper-catalyzed enantioselective difluoromethylation of Meldrum’s acid derivative 83 using (difluoromethyl)zinc reagents (Scheme 31) [79]. The appropriate β-difluoromethylated carbonyl compounds 84 were formed with enantioselectivities around 80% of ee in high to excellent chemical yields [79].
Chiral P,N-ligands 88 were also used in copper-mediated tandem double Michael reactions of diethylzinc to α,β-unsaturated ketones 86 followed by trapping with nitroalkenes 87 (Scheme 32) [80]. The corresponding tandem products were in high yields and with excellent enantioselectivities [80].
Diarylprolinol silyl ether 92 was successfully employed in the asymmetric cyclization of o-hydroxycinnamaldehydes 89 with diphenylphosphine oxide 90 (Scheme 33) [81]. Organophosphorus compounds are widespread in the natural environment and in many cases have important biological properties. The corresponding 4-diphenylphosphinyl chroman-2-ols 91 were obtained with 84–99% ee and 7:1 to 20:1 of diastereomeric ratio [81].
Silyl ethers of diarylprolinol were also used in the highly stereoselective synthesis of spiro pyrazolones via Michael/Conia-ene cascade reaction [82] and in the synthesis of biaryl atropoisomers via domino Michael–Henry reaction affording the substituted nitrocyclohexanecarbaldehydes [83].

6. Asymmetric Reactions in the Presence of Zinc Ions

Among the very wide range of asymmetric reactions taking place in the presence of zinc ions, the most exploited but still valid and simple reactions are asymmetric additions of diethylzinc to carbonyl compounds, especially aldehydes. In 2018, Wang described an enantioselective analysis towards the rational design of chiral ligands efficiently catalyzing the asymmetric addition of diethylzinc to benzaldehyde [84].
The effect of regioisomerism on the efficiency of amino alcohol ligands in the asymmetric addition of diethylzinc to benzaldehyde was investigated by Mátravölgyi et al. [85]. The aforementioned reactions performed in the presence of atropisomeric 1-phenylpyrroles 93 and 94 led to the opposite enantiomers of desired alcohol with high enantioselectivity (Scheme 34) [85].
Recent developments towards efficient, highly enantioselective addition of diethylzinc to aldehydes include the use of chiral derivatives such as (Figure 2): 5-cis-substituted proline derivatives (prolinols 95 and prolinamines 96) [86], proline-based N,N-dioxides 97 [87], pinane-based tridentate ligands 98 [88], axially chiral tridentate isoquinoline-derived ligands 99 [89], chiral oxazoline-based systems 100 [90], thiophene-derived amino alcohols 101 [91], amino alcohols 102 [92], noscapine-derived β-amino alcohols 103 [93,94], chiral ferrocene and ruthenocene substituted aminomethylnaphthols 104 [95], and chiral P,N-ligands 105 [96].
Enantioselective copper(II)-promoted 1,4-conjugate addition of Et2Zn to nitrodienes 106 was reported by Wu et al. [97]. The corresponding 1,4-adducts 107 were obtained in 81–98% yield and with 87–97% ee when a chiral amidophosphine ligand 108 was applied (Scheme 35) [97].
Enantioselective aldol-type reaction of trifluorodiazomethane 109 with various aldehydes was efficiently catalyzed by quinine in the presence of diethylzinc in THF (Scheme 36) [98]. The corresponding chiral β-trifluoromethyl alcohols 110 were constructed in very satisfactory levels of yield and enantioselectivity [98].

7. Summary

This review article presented selected examples of recent achievements in the field of asymmetric synthesis with chiral catalysts. The authors’ attention was focused on asymmetric transformations such as the cyclopropanation, Friedel–Crafts, Mannich, Michael reactions, and reactions in the presence of zinc ions. All the examples given relate to reactions with high chemical yields and high stereoselectivity.
The research related to asymmetric synthesis with the use of chiral organic catalysts is extremely well studied and is constantly developed by many research groups around the world. The asymmetric reactions presented in this review are well known and exploited in the literature. Our suggestion for future research directions in this field is that researchers should focus on less obvious asymmetric transformations, such as Morita–Baylis–Hillman or Rauhut–Currier reactions, whose chiral products have potential applications in many fields. Moreover, perhaps it is worth trying to design chiral catalysts taking into account new structural motifs, such as the aziridine ring.

Author Contributions

Conceptualization, M.R. and S.L.; software, A.B.-S. and M.R.; formal analysis, M.R. and S.L.; writing—original draft preparation, M.R. and A.B.-S.; writing—review and editing, M.R. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. Eliel, E.L.; Wilen, S.H. Stereochemistry of Organic Compounds; John Wiley & Sons, Ltd.: New York, NY, USA, 1994; pp. 1–1296. [Google Scholar]
  2. Ariëns, E.J. Stereochemistry: A source of problems in medicinal chemistry. Med. Res. Rev. 1986, 6, 451–466. [Google Scholar] [CrossRef] [PubMed]
  3. Stephens, T.D.; Bunde, C.J.W.; Fillmore, B.J. Mechanism of action in thalidomide teratogenesis. Biochem. Pharmacol. 2000, 59, 1489–1499. [Google Scholar] [CrossRef]
  4. Botting, J. The history of thalidomide. Drug News Perspect. 2002, 15, 604–611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rachwalski, M.; Vermue, N.; Rutjes, F.P.J.T. Recent advances in enzymatic and chemical deracemisation of racemic compounds. Chem. Soc. Rev. 2013, 42, 926. [Google Scholar] [CrossRef] [Green Version]
  6. Rachwalski, M.; Kaczmarczyk, S.; Leśniak, S.; Kiełbasiński, P. Highly Efficient Asymmetric Simmons–Smith cyclopropanation promoted by chiral heteroorganic aziridinyl ligands. ChemCatChem 2014, 6, 873–875. [Google Scholar] [CrossRef]
  7. Buchcic, A.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Asymmetric friedel-crafts alkylation of indoles catalyzed by chiral aziridine-phosphines. Catalysts 2020, 10, 971. [Google Scholar] [CrossRef]
  8. Rachwalski, M.; Leenders, T.; Kaczmarczyk, S.; Kiełbasiński, P.; Leśniak, S.; Rutjes, F.P.J.T. Efficient catalysts for asymmetric Mannich reactions. Org. Biomol. Chem. 2013, 11, 4207–4213. [Google Scholar] [CrossRef]
  9. Buchcic, A.; Zawisza, A.; Leśniak, S.; Adamczyk, J.; Pieczonka, A.M.; Rachwalski, M. Enantioselective mannich reaction promoted by chiral phosphinoyl-aziridines. Catalysts 2019, 9, 837. [Google Scholar] [CrossRef] [Green Version]
  10. Wujkowska, Z.; Zawisza, A.; Leśniak, S.; Rachwalski, M. Phosphinoyl-aziridines as a new class of chiral catalysts for enantioselective Michael addition. Tetrahedron 2019, 75, 230–235. [Google Scholar] [CrossRef]
  11. Leśniak, S.; Rachwalski, M.; Pieczonka, A.M. Optically pure aziridinyl ligands as useful catalysts in the stereocontrolled synthesis. Curr. Org. Chem. 2014, 18, 3045–3065. [Google Scholar] [CrossRef]
  12. Časar, Z. Synthetic approaches to contemporary drugs that contain the cyclopropyl moiety. Synthesis 2020, 52, 1315–1345. [Google Scholar] [CrossRef]
  13. Wu, W.; Lin, Z.; Jiang, H. Recent advances in the synthesis of cyclopropanes. Org. Biomol. Chem. 2018, 16, 7315–7329. [Google Scholar] [CrossRef]
  14. Novakov, I.A.; Babushkin, A.S.; Yablokov, A.S.; Nawrozkij, M.B.; Vostrikova, O.V.; Shejkin, D.S.; Mkrtchyan, A.S.; Balakin, K.V. Synthesis and structure—Activity relationships of cyclopropane-containing analogs of pharmacologically active compounds. Russ. Chem. Bull. Int. Ed. 2018, 67, 395–418. [Google Scholar] [CrossRef]
  15. Jüttner, F.; Wurster, K. Evidence of ectocarpene and dictyopterenes A and C’ in the water of a freshwater lake. Limnol. Oceanogr. 1984, 29, 1322–1324. [Google Scholar] [CrossRef]
  16. Kim, T.; Kim, J.Y.; Park, K.Y.; Ryu, D.H. Asymmetric synthesis of (-)-Dictyopterene C’ and its derivatives via catalytic enantioselective cyclopropanation. Bull. Korean Chem. Soc. 2021, 42, 675–678. [Google Scholar] [CrossRef]
  17. Sun, M.; Li, W.-D.Z.; Qiu, F.G. Asymmetric synthesis of the DEFG rings of solanoeclepin A. Org. Lett. 2019, 21, 644–647. [Google Scholar] [CrossRef]
  18. Pons, A.; Tognetti, V.; Joubert, L.; Poisson, T.; Pannecoucke, X.; Charette, A.B.; Jubault, P. Catalytic enantioselective cyclopropanation of α-fluoroacrylates: An experimental and theoretical study. ACS Catal. 2019, 9, 2594–2598. [Google Scholar] [CrossRef]
  19. Zheng, G.; Zhou, Z.; Zhu, G.; Zhai, S.; Xu, H.; Duan, X.; Yi, W.; Li, X. Rhodium(III)-catalyzed enantio- and diastereoselective C-H Cyclopropylation of N-Phenoxylsulfonamides: Combined experimental and computational studies. Angew. Chem. Int. Ed. 2020, 59, 2890–2896. [Google Scholar] [CrossRef]
  20. Heckenbichler, K.; Schweiger, A.; Brandner, L.A.; Binter, A.; Toplak, M.; Macheroux, P.; Gruber, K.; Breinbauer, R. Asymmetric reductive carbocyclization using engineered ene reductases. Angew. Chem. Int. Ed. 2018, 57, 7240–7244. [Google Scholar] [CrossRef]
  21. Poulsen, T.B.; Jörgensen, K.A. Catalytic asymmetric friedel-crafts alkylation reactions—Copper showed the way. Chem. Rev. 2008, 108, 2903–2915. [Google Scholar] [CrossRef]
  22. Maltsev, O.V.; Beletskaya, I.P.; Zlotin, S.G. Organocatalytic Michael and Friedel-Crafts reactions in enantioselective synthesis of biologically active compounds. Russ. Chem. Rev. 2011, 80, 1067–1113. [Google Scholar] [CrossRef]
  23. Heravi, M.M.; Zadsirjan, V.; Heydari, M.; Masoumi, B. Organocatalyzed asymmetric Friedel-Crafts reaction: An update. Chem. Rec. 2019, 19, 2236–2340. [Google Scholar] [CrossRef]
  24. Concepción Gimeno, M.; Herrera, R.P. Hydrogen bonding and internal or external Lewis or Brønsted acid assisted (Thio)urea catalysts. Eur. J. Org. Chem. 2020, 1057–1068. [Google Scholar] [CrossRef]
  25. Sonsona, I.G.; Marqués-López, E.; Häring, M.; Díaz Díaz, D.; Herrera, R.P. Urea activation by an external Brønsted acid: Breaking self-association and tuning catalytic performance. Catalysts 2018, 8, 305. [Google Scholar] [CrossRef] [Green Version]
  26. Izaga, A.; Herrera, R.P.; Concepción Gimeno, M. Gold(I)-Mediated Thiourea Organocatalyst Activation: A synergic effect for asymmetric catalysis. ChemCatChem 2017, 9, 1313–1321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Han, J.-L.; Liao, Y.-T.; Chang, C.-H. Asymmetric organocatalytic conjugate addition of electron-rich phenols and 1,3-dicarbonyls to arylsulfonyl indoles in an oil-water biphasic system. Eur. J. Org. Chem. 2019, 5815–5823. [Google Scholar] [CrossRef]
  28. Wang, C.-J.; Yang, Q.-Q.; Wang, M.-X.; Shang, Y.-H.; Tong, X.-Y.; Deng, Y.-H.; Shao, Z. Catalytic asymmetric 1,4-type Friedel-Crafts (hetero)arylations of 1-azadienes: The highly enantioselective syntheses of chiral hetero-triarylmethanes. Org. Chem. Front. 2020, 7, 609–616. [Google Scholar] [CrossRef]
  29. Zhang, C.; Cheng, Y.; Li, F.; Luan, Y.; Li, P.; Li, W. Organocatalytic enantioselective regiodivergent C-H bond functionalization of 1-Naphthols with 1-Azadienes. Adv. Synth. Catal. 2020, 362, 1286–1291. [Google Scholar] [CrossRef]
  30. Gao, Y.; Wang, X.; Wie, Z.; Cao, J.; Liang, D.; Lin, Y.; Duan, H. Asymmetric synthesis of spirooxindole-pyranoindole products via Friedel-Crafts alkylation/cyclization of the indole carbocyclic ring. New J. Chem. 2020, 44, 9788–9792. [Google Scholar] [CrossRef]
  31. Liu, J.-Y.; Yang, X.-C.; Lu, H.; Gu, Y.-C.; Xu, P.-F. Organocatalytic, enantioselective friedel-crafts reaction of indoles in the carbocyclic ring and electron-rich phenols. Org. Lett. 2018, 20, 2190–2194. [Google Scholar] [CrossRef]
  32. Xu, K.; Chen, W.; Chen, X.; Wang, B.; Huang, J.; Tian, X. Organocatalytic asymmetric Friedel-Crafts alkylation/hemiketalization/lactonization cascade reactions: Highly enantioselective synthesis of furo[2,3-b]benzofuranones. Org. Chem. Front. 2020, 7, 1679–1684. [Google Scholar] [CrossRef]
  33. Zhao, L.; Bao, X.; Hu, Q.; Wang, B.; Lu, A.-H. Porous carbon nanosheet-supported chiral squaramide for highly enantioselective Friedel-crafts reaction. ChemCatChem 2018, 10, 1248–1252. [Google Scholar] [CrossRef]
  34. Wang, W.; Xiong, W.; Wang, J.; Wang, Q.-A.; Yang, W. Brønsted acid-catalyzed asymmetric Friedel-crafts alkylation of indoles with benzothiazole-bearing trifluoromethyl ketone hydrates. J. Org. Chem. 2020, 85, 4398–4407. [Google Scholar] [CrossRef]
  35. Shen, M.-L.; Shen, Y.; Wang, P.-S. Merging visible-light photoredox and chiral phosphate catalysis for asymmetric friedel-crafts reaction with in situ generation of N-acyl imines. Org. Lett. 2019, 21, 2993–2997. [Google Scholar] [CrossRef]
  36. Lee, J.; Kim, S.-G. Asymmetric phosphoric acid-catalyzed aza-friedel crafts reaction of furan with cyclic N-sulfimines. Bull. Korean Chem. Soc. 2019, 40, 606–609. [Google Scholar] [CrossRef]
  37. Ma, J.; Kass, S.R. Electrostatically enhanced phosphoric acids and their applications in asymmetric Friedel-crafts alkylations. J. Org. Chem. 2019, 84, 11125–11134. [Google Scholar] [CrossRef]
  38. Liu, W.; Rajkumar, S.; Wu, W.; Huang, Z.; Yang, X. Asymmetric synthesis of β-indolyl cyclopentanones and cyclopentylamides with an all-carbon quaternary stereocenter via chiral phosphoric acid catalyzed friedel-crafts alkylation reactions. Org. Lett. 2019, 21, 3563–3567. [Google Scholar] [CrossRef]
  39. Ma, J.; Kass, S.R. Asymmetric arylation of 2,2,2-trifluoroacetophenones catalyzed by chiral electrostatically-enhanced phosphoric acids. Org. Lett. 2018, 20, 2689–2692. [Google Scholar] [CrossRef]
  40. Maestro, A.; Martinez de Marigorta, E.; Palacios, F.; Vicario, J. Enantioselective α-aminophosphonate functionalization of indole ring through an organocatalyzed friedel-crafts reaction. J. Org. Chem. 2019, 84, 1094–1102. [Google Scholar] [CrossRef]
  41. Gong, W.; Chen, X.; Jiang, H.; Chu, D.; Cui, Y.; Liu, Y. Highly stable Zr(IV)-based metal-organic frameworks with chiral phosphoric acids for catalytic asymmetric tandem reactions. J. Am. Chem. Soc. 2019, 141, 7498–7508. [Google Scholar] [CrossRef]
  42. Kim, Y.; Lee, J.; Jung, J.; Kim, S.-G. Chiral Brønsted acid-catalyzed Friedel-Crafts reaction of 3-indolylsulfamidates with indoles: Synthesis of enantioenriched bisindolylmethane sulfamates. Tetrahedron Lett. 2019, 60, 1625–1630. [Google Scholar] [CrossRef]
  43. Anokhin, M.V.; Feofanov, M.N.; Averin, A.D.; Beletskaya, I.P. The asymmetric Friedel-Crafts reaction of indoles with arylidenemalonates catalyzed by MgI2/PyBox complexes. ChemistrySelect 2018, 3, 1388–1391. [Google Scholar] [CrossRef]
  44. Tarasenko, E.A.; Shestakov, I.V.; Rybakov, V.B.; Beletskaya, I.P. Enantioselective Copper(II)/box-catalyzed synthesis of chiral β3-tryptophan derivatives. ChemCatChem 2019, 11, 3913–3918. [Google Scholar] [CrossRef] [Green Version]
  45. Shahidul Islam, M.; Barakat, A.; Al-Majid, A.M.; Ali, M.; Yousuf, S.; Iqbal Choudhary, M.; Khalid, R.; Ul-Haq, Z. Catalytic asymmetric synthesis of indole derivatives as novel α-glucosidase inhibitors in vitro. Bioorg. Chem. 2018, 79, 350–354. [Google Scholar] [CrossRef] [PubMed]
  46. Zuo, Y.; Yang, N.; Huang, X.; Hu, C.; Su, Z. Mechanism and origins of stereoinduction in an asymmetric Friedel-Crafts alkylation reaction of chalcone catalyzed by chiral N,N′-dioxide-Sc(III) complex. J. Org. Chem. 2018, 83, 4628–4640. [Google Scholar] [CrossRef]
  47. Kitanosono, T.; Hisada, T.; Yamashita, Y.; Kobayashi, S. Hydrogen-bonding-assisted cationic aqua Palladium(II) complex enables highly efficient asymmetric reactions in water. Angew. Chem. Int. Ed. 2021, 60, 3407–3411. [Google Scholar] [CrossRef] [PubMed]
  48. Zhu, W.-J.; Gong, J.-F.; Song, M.-P. Synthesis of chiral Bis(3-indolyl)methanes bearing a trifluoromethylated all-carbon quaternary stereocenter via nickel-catalyzed asymmetric Friedel-crafts alkylation reaction. J. Org. Chem. 2020, 85, 9525–9537. [Google Scholar] [CrossRef] [PubMed]
  49. Phuc Le, T.; Higashita, K.; Tanaka, S.; Yoshimura, M.; Kitamura, M. Revisiting the Cuii-catalyzed asymmetric Friedel-Crafts reaction of indole with trifluoropyruvate. Org. Lett. 2018, 20, 7149–7153. [Google Scholar] [CrossRef]
  50. Hua, Y.-Z.; Chen, J.-W.; Yang, H.; Wang, M.-C. Asymmetric Friedel-Crafts alkylation of indoles with trifluoromethyl pyruvate catalyzed by a dinuclear zinc catalyst. J. Org. Chem. 2018, 83, 1160–1166. [Google Scholar] [CrossRef]
  51. Yang, X.-C.; Liu, M.-M.; Mathey, F.; Yang, H.; Hua, Y.-Z. Access to chiral 2,5-pyrrolidinyl dispirooxindoles via dinuclear zinc-catalyzed asymmetric cascade reactions. J. Org. Chem. 2019, 84, 7762–7775. [Google Scholar] [CrossRef]
  52. Guo, Y.-J.; Guo, X.; Kong, D.-Z.; Lu, H.-J.; Liu, L.-T.; Hua, Y.-Z.; Wang, M.-C. Catalytic asymmetric synthesis of tetrahydrofuran spirooxindoles via a dinuclear zinc catalyst. J. Org. Chem. 2020, 85, 4195–4206. [Google Scholar] [CrossRef]
  53. Hua, Y.-Z.; Han, X.-W.; Huang, L.-H.; Wang, M.-C. Asymmetric Friedel-Crafts alkylation of pyrrole with chalcones catalyzed by a dinuclear zinc catalyst. Chin. J. Org. Chem. 2018, 38, 237–245. [Google Scholar] [CrossRef]
  54. Wang, Z.; Zu, L. Organocatalytic enantioselective direct alkylation of phloroglucinol derivatives: Asymmetric total synthesis of (+)-aflatoxin B2. Chem. Commun. 2019, 55, 5171–5174. [Google Scholar] [CrossRef] [PubMed]
  55. Miyazaki, Y.; Zhou, B.; Tsuji, H.; Kawatsura, M. Nickel-catalyzed asymmetric friedel-crafts propargylation of 3-substituted indoles with propargylic carbonates bearing an internal alkyne group. Org. Lett. 2020, 22, 2049–2053. [Google Scholar] [CrossRef]
  56. Anwar, M.; Yang, S.; Xu, W.; Liu, J.-G.; Perveen, S.; Kong, X.-W.; Tazeen Zehra, S.; Fang, X.-Q. Carbene-catalyzed asymmetric Friedel-Crafts alkylation-annulation sequence and rapid synthesis of indole-fused polycyclic alkaloids. Commun. Chem. 2019, 2, 85. [Google Scholar] [CrossRef] [Green Version]
  57. Bai, J.; Sun, X.; Wang, H.-S.; Li, C.; Qiao, R.-Z. Guanosine-based sel-assembly as an enantioselective catalyst scaffold. J. Org. Chem. 2020, 85, 2010–2018. [Google Scholar] [CrossRef]
  58. Bai, J.K.; Chen, D.; Li, C.; Wang, H.S.; Qiao, R.Z. PNA as hybrid catalyst scaffold catalyzed asymmetric friedel-crafts alkylation. Catal. Lett. 2020, 150, 2082–2090. [Google Scholar] [CrossRef]
  59. Mansot, J.; Aubert, S.; Duchemin, N.; Vasseur, J.-J.; Arseniyadis, S.; Smietana, M. A rational quest for selectivity through precise ligand-positioning in tandem DNA-catalysed Friedel-Crafts alkylation/asymmetric protonation. Chem. Sci. 2019, 10, 2875–2881. [Google Scholar] [CrossRef] [Green Version]
  60. Dey, S.; Jäschke, A. Covalently functionalized DNA duplexes and quadruplexes as hybrid catalysts in an enantioselective friedel-crafts reaction. Molecules 2020, 25, 3121. [Google Scholar] [CrossRef]
  61. Ravi Kumar, G.; Ramesh, B.; Yarlagadda, S.; Sridhar, B.; Subba Reddy, B.V. Organocatalytic enantioselective Mannich reaction: Direct access to chiral β-amino esters. ACS Omega 2019, 4, 2168–2177. [Google Scholar] [CrossRef] [Green Version]
  62. Yu, J.-S.; Noda, H.; Shibasaki, M. Exploiting β-amino acid enolates in direct catalytic diastereo-and enantioselective C-C bond-forming reactions. Chem. Eur. J. 2018, 24, 15796–15800. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, H.-J.; Zhong, F.; Xie, Y.-C.; Yin, L. Catalytic asymmetric Mannich-type reaction enabled by efficient dienolization of α,β-unsaturated pyrazoleamides. Chin. J. Chem. 2021, 39, 55–61. [Google Scholar] [CrossRef]
  64. Rostoll-Berenguer, J.; Blay, G.; Carmen-Muñoz, M.; Pedro, J.R.; Vila, C. A Combination of visible-light organophotoredox catalysis and asymmetric organocatalysis for the enantioselective Mannich reaction of dihydroquinoxalinones with ketones. Org. Lett. 2019, 21, 6011–6015. [Google Scholar] [CrossRef] [PubMed]
  65. Dong, C.-L.; Ding, X.; Huang, L.-Q.; He, Y.-H.; Guan, Z. Merging visible light photocatalysis and l-/d-proline catalysis: Direct asymmetric oxidative dearomatization of 2-arylindoles to access C2-quaternary indolin-3-ones. Org. Lett. 2020, 22, 1076–1080. [Google Scholar] [CrossRef]
  66. Li, J.-S.; Liu, Y.-J.; Li, S.; Ma, J.-A. Chiral phosphoric acid-catalyzed direct asymmetric mannich reaction of cyclic C-acylimines with simple ketones: Facile access to C2-quaternary indolin-3-ones. Chem. Commun. 2018, 54, 9151–9154. [Google Scholar] [CrossRef]
  67. 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]
  68. Malkar, R.S.; Jadhav, A.L.; Yadav, G.D. Innovative catalysis in Michael addition reactions for C-X bond formation. Mol. Catal. 2020, 485, 110814. [Google Scholar] [CrossRef]
  69. Gandhi, S.; Sivadas, V.; Baire, B. Thiourea-tertiary amine promoted cascade catalysis: A tool for complexity generation. Eur. J. Org. Chem. 2021, 220–234. [Google Scholar] [CrossRef]
  70. Rohit, K.R.; Ujwaldev, S.M.; Krishnan, K.K.; Anilkumar, G. Recent developments and perspectives in the zinc-catalysed Michael addition. Asian, J. Org. Chem. 2018, 7, 85–102. [Google Scholar] [CrossRef]
  71. Macchia, A.; Eitzinger, A.; Brière, J.-F. Asymmetric synthesis of isoxazol-5-ones and isoxazolidin-5-ones. Synthesis 2021, 53, 107–122. [Google Scholar] [CrossRef]
  72. Carceller-Ferrer, L.; Blay, G.; Pedro, J.R.; Vila, C. Recent advances in catalytic enantioselective synthesis of pyrazolones with a tetrasubstituted stereogenic center at the 4-position. Synthesis 2020, 53, 215–237. [Google Scholar] [CrossRef]
  73. Torán, R.; Vila, C.; Sanz-Marco, A.; Carmen-Muñoz, M.; Pedro, J.R.; Blay, G. Organocatalytic enantioselective 1,6-aza-Michael addition of isoxazolin-5-ones to p-quinone methides. Eur. J. Org. Chem. 2020, 627–630. [Google Scholar] [CrossRef]
  74. Yang, Z.-T.; Zhao, J.; Yang, W.-L.; Deng, W.-P. Enantioselective construction of CF3-Containing Spirooxindole γ-Lactones via organocatalytic asymmetric michael/lactonization. Org. Lett. 2019, 21, 1015–1020. [Google Scholar] [CrossRef]
  75. Xu, X.; He, Y.; Zhou, J.; Li, X.; Zhu, B.; Chang, J. Organocatalytic asymmetric michael addition of pyrazol-5-ones to β-trifluoromethyl-α,β-unsaturated ketones: Stereocontrolled construction of vicinal quaternary and tertiary stereocenters. J. Org. Chem. 2020, 85, 574–584. [Google Scholar] [CrossRef]
  76. Xiao, W.; Zhou, Z.; Yang, Q.-Q.; Du, W.; Chen, Y.-C. Organocatalytic asymmetric four-component [5+1+1+1] cycloadditions via a quintuple cascade process. Adv. Synth. Catal. 2018, 360, 3526–3533. [Google Scholar] [CrossRef]
  77. Macchia, A.; Dafnae Cuomo, V.; Di Mola, A.; Pierri, G.; Tedesco, C.; Palombi, L.; Massa, A. On the necessity of one-pot tautomer trapping in asymmetric michael reactions of arylideneisoxazol-5-ones. Eur. J. Org. Chem. 2020, 2264–2270. [Google Scholar] [CrossRef]
  78. Pair, E.; Cadart, T.; Levacher, V.; Brière, J.-F. Meldrum’s acid: A useful platform in asymmetric organocatalysis. ChemCatChem 2016, 8, 1882–1890. [Google Scholar] [CrossRef]
  79. Endo, Y.; Ishii, K.; Mikami, K. Chiral copper-catalyzed enantioselective Michael difluoromethylation of arylidene meldrum’s acids with (difluoromethyl)zinc reagents. Tetrahedron 2019, 75, 4099–4103. [Google Scholar] [CrossRef]
  80. Wang, Q.; Li, S.; Hou, C.-J.; Chu, T.-T.; Hu, X.-P. Copper-catalyzed asymmetric tandem double Michael reactions of diethylzinc to α,β-unsaturated ketones followed by trapping with nitroolefins. Tetrahedron 2019, 75, 3943–3950. [Google Scholar] [CrossRef]
  81. Sun, H.; Li, Y.; Liu, W.; Zheng, Y.; He, Z. Organocatalytic asymmetric cascade cyclization reaction of o-hydroxycinnamaldehydes with diphenylphosphine oxide. Chin. Chem. Lett. 2018, 29, 1625–1628. [Google Scholar] [CrossRef]
  82. Putatunda, S.; Alegre-Requena, J.V.; Meazza, M.; Franc, M.; Rohalová, D.; Vemuri, P.; Císařová, I.; Herrera, R.P.; Rios, R.; Veselý, J. Proline bulky substituents consecutively act as steric hindrances and directing groups in a Michael/Conia-ene cascade reaction under synergistic catalysis. Chem. Sci. 2019, 10, 4107–4115. [Google Scholar] [CrossRef] [Green Version]
  83. Hayashi, Y.; Takikawa, A.; Koshino, S.; Ishida, K. Asymmetric Synthesis of Biaryl Atropoisomers Using an Organocatalyst-Mediated Domino Reaction as the Key Step. Chem. Eur. J. 2019, 25, 10319–10322. [Google Scholar] [CrossRef]
  84. Wang, M.-C. Enantioselective analysis: Logic of chiral ligand design for asymmetric addition of diethylzinc to benzaldehyde. Chin. J. Org. Chem. 2018, 38, 162–170. [Google Scholar] [CrossRef]
  85. Mátravölgyi, B.; Deák, S.; Erdélyi, Z.; Hergert, T.; Ábrányi-Balogh, P.; Faigl, F. Effect of regioisomerism on the efficiency of 1-phenylpyrrole-type atropisomeric amino alcohol ligands in enantioselective organometallic reactions. Synlett 2018, 29, 2171–2175. [Google Scholar] [CrossRef] [Green Version]
  86. Prause, F.; Wagner, S.; Breuning, M. Enantioselective addition of diethylzinc to aldehydes catalyzed by 5-cis-substituted proline derivatives. Tetrahedron 2019, 75, 94–101. [Google Scholar] [CrossRef]
  87. Gao, E.; Li, Q.; Duan, L.; Li, L.; Li, Y.-M. Isosterically designed chiral catalysts: Rationale, optimization and their application in enantioselective nucleophilic addition to aldehydes. Tetrahedron 2020, 76, 131648. [Google Scholar] [CrossRef]
  88. Raji, M.; Minh Le, T.; Fülöp, F.; Szakonyi, Z. Synthesis and investigation of pinane-based chiral tridentate ligands in the asymmetric addition of diethylzinc to aldehydes. Catalysts 2020, 10, 474. [Google Scholar] [CrossRef]
  89. Sweetman, B.A.; Guiry, P.J. Axially chiral tridentate isoquinoline derived ligands for diethylzinc addition to aldehydes. Tetrahedron 2018, 74, 5567–5581. [Google Scholar] [CrossRef]
  90. Aydin, A.E. Enantioselective addition of diethylzinc to aromatic aldehydes using chiral oxazoline-based ligands. Russ. J. Org. Chem. 2020, 56, 1304–1312. [Google Scholar] [CrossRef]
  91. Aydin, A.E. Enantioselective addition of diethylzinc to aromatic aldehydes using novel thiophene-based chiral ligands. Russ. J. Org. Chem. 2020, 56, 901–909. [Google Scholar] [CrossRef]
  92. Sappino, C.; Mari, A.; Mantineo, A.; Moliterno, M.; Palagri, M.; Tatangelo, C.; Suber, L.; Bovicelli, P.; Ricelli, A.; Righi, G. New chiral amino alcohol ligands for catalytic enantioselective addition of diethylzinc to aldehydes. Org. Biomol. Chem. 2018, 16, 1860–1870. [Google Scholar] [CrossRef] [PubMed]
  93. Mohebbi, M.; Salehi, P.; Bararjanian, M.; Ebrahimi, S.N. Noscapine-derived β-amino alcohols as new organocatalysts for enantioselective addition of diethylzinc to aldehydes. J. Iran. Chem. Soc. 2018, 15, 47–53. [Google Scholar] [CrossRef]
  94. Mohebbi, M.; Bararjanian, M.; Ebrahimi, S.N.; Smieško, M.; Salehi, P. Noscapine derivatives as new chiral catalysts in asymmetric synthesis: Highly enantioselective addition of diethylzinc to aldehydes. Synthesis 2018, 50, 1841–1848. [Google Scholar] [CrossRef] [Green Version]
  95. Dikova, K.; Kostova, K.; Simova, S.; Linden, A.; Chimov, A.; Dimitrov, V. Synthesis and crystal structures of chiral ferrocene and ruthenocene substituted aminomethylnaphthols obtained through Betti-condensation. Polyhedron 2019, 165, 177–187. [Google Scholar] [CrossRef]
  96. Wang, Q.; Li, S.; Hou, C.-J.; Chu, T.-T.; Hu, X.-P. Chiral P,N-ligands for the highly enantioselective addition of diethylzinc to aromatic aldehydes. Appl. Organometal. Chem. 2019, 33, e5108. [Google Scholar] [CrossRef]
  97. Rexiti, R.; Zhang, Z.-G.; Lu, J.; Sha, F.; Wu, X.-Y. Regioselective and enantioselective Cu(II)-catalyzed 1,4-conjugate addition of diethylzinc reagent to nitrodienes. J. Org. Chem. 2019, 84, 1330–1338. [Google Scholar] [CrossRef] [PubMed]
  98. Rong, M.-Y.; Yang, L.; Nie, J.; Zhang, F.-G.; Ma, J.-A. Construction of chiral β-trifluoromethyl alcohols enabled by catalytic enantioselective aldol-type reaction of CF3CHN2. Org. Lett. 2019, 21, 4280–4283. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Asymmetric synthesis of Dictyopterene C’.
Scheme 1. Asymmetric synthesis of Dictyopterene C’.
Symmetry 13 01762 sch001
Scheme 2. Synthesis of solanoeclepin A core starting from (R)-seudenol.
Scheme 2. Synthesis of solanoeclepin A core starting from (R)-seudenol.
Symmetry 13 01762 sch002
Scheme 3. Enantioselective cyclopropanation of α-fluoroacrylates.
Scheme 3. Enantioselective cyclopropanation of α-fluoroacrylates.
Symmetry 13 01762 sch003
Scheme 4. Stereoselective cyclopropanation of N-phenoxylsulfonamides.
Scheme 4. Stereoselective cyclopropanation of N-phenoxylsulfonamides.
Symmetry 13 01762 sch004
Scheme 5. Enzyme-promoted C-C bond formation via reductive cyclization.
Scheme 5. Enzyme-promoted C-C bond formation via reductive cyclization.
Symmetry 13 01762 sch005
Scheme 6. Reaction of indole and nitrostyrene promoted by urea or thiourea catalysts.
Scheme 6. Reaction of indole and nitrostyrene promoted by urea or thiourea catalysts.
Symmetry 13 01762 sch006
Scheme 7. Reaction of sulfonylindole with naphthol catalyzed by thiourea 19.
Scheme 7. Reaction of sulfonylindole with naphthol catalyzed by thiourea 19.
Symmetry 13 01762 sch007
Scheme 8. Urea-promoted arylation of 1-azadienes.
Scheme 8. Urea-promoted arylation of 1-azadienes.
Symmetry 13 01762 sch008
Scheme 9. Alkylation/cyclization of 4-hydroxyindole in the presence of urea system 24.
Scheme 9. Alkylation/cyclization of 4-hydroxyindole in the presence of urea system 24.
Symmetry 13 01762 sch009
Scheme 10. Friedel–Crafts reaction of indole catalyzed by squaramide derivative 27.
Scheme 10. Friedel–Crafts reaction of indole catalyzed by squaramide derivative 27.
Symmetry 13 01762 sch010
Scheme 11. Synthesis of furo [2,3-b] benzofuranones promoted by squaramide 30.
Scheme 11. Synthesis of furo [2,3-b] benzofuranones promoted by squaramide 30.
Symmetry 13 01762 sch011
Scheme 12. Friedel–Crafts reaction of pyrazolone with isatine ketimine.
Scheme 12. Friedel–Crafts reaction of pyrazolone with isatine ketimine.
Symmetry 13 01762 sch012
Scheme 13. Acid-catalyzed Friedel–Crafts alkylation of indoles.
Scheme 13. Acid-catalyzed Friedel–Crafts alkylation of indoles.
Symmetry 13 01762 sch013
Scheme 14. Friedel–Crafts reaction of indoles with in situ generated N-acyl imines.
Scheme 14. Friedel–Crafts reaction of indoles with in situ generated N-acyl imines.
Symmetry 13 01762 sch014
Scheme 15. Enantioselective reaction of N-sulfimines 38 with 2-methoxyfurane.
Scheme 15. Enantioselective reaction of N-sulfimines 38 with 2-methoxyfurane.
Symmetry 13 01762 sch015
Scheme 16. Reactions of indole with malonate derivatives promoted by PyBox and iPrBox.
Scheme 16. Reactions of indole with malonate derivatives promoted by PyBox and iPrBox.
Symmetry 13 01762 sch016
Scheme 17. Reaction of indole with enones.
Scheme 17. Reaction of indole with enones.
Symmetry 13 01762 sch017
Scheme 18. Synthesis of chiral bis (3-indolyl) methanes.
Scheme 18. Synthesis of chiral bis (3-indolyl) methanes.
Symmetry 13 01762 sch018
Scheme 19. Alkylation of indoles with trifluoromethyl pyruvates.
Scheme 19. Alkylation of indoles with trifluoromethyl pyruvates.
Symmetry 13 01762 sch019
Scheme 20. Enantioselective alkylation of phloroglucinol derivatives.
Scheme 20. Enantioselective alkylation of phloroglucinol derivatives.
Symmetry 13 01762 sch020
Scheme 21. Asymmetric Friedel–Crafts propargylation of indoles.
Scheme 21. Asymmetric Friedel–Crafts propargylation of indoles.
Symmetry 13 01762 sch021
Scheme 22. Asymmetric Mannich reaction leading to chiral β-amino esters.
Scheme 22. Asymmetric Mannich reaction leading to chiral β-amino esters.
Symmetry 13 01762 sch022
Scheme 23. Mannich reaction of isoxazolidin-5-ones with isatin-derived imines.
Scheme 23. Mannich reaction of isoxazolidin-5-ones with isatin-derived imines.
Symmetry 13 01762 sch023
Scheme 24. Asymmetric Mannich reaction of N-Boc aldimines with pyrazoleamides.
Scheme 24. Asymmetric Mannich reaction of N-Boc aldimines with pyrazoleamides.
Symmetry 13 01762 sch024
Scheme 25. Mannich reaction of dihydroquinoxalinones with ketones.
Scheme 25. Mannich reaction of dihydroquinoxalinones with ketones.
Symmetry 13 01762 sch025
Scheme 26. Synthesis of C2-quaternary indolin-3-ones.
Scheme 26. Synthesis of C2-quaternary indolin-3-ones.
Symmetry 13 01762 sch026
Scheme 27. Mannich reaction catalyzed by chiral phosphoric acid.
Scheme 27. Mannich reaction catalyzed by chiral phosphoric acid.
Symmetry 13 01762 sch027
Scheme 28. Enantioselective Michael addition of isoxazolin-5-ones to p-quinone methides.
Scheme 28. Enantioselective Michael addition of isoxazolin-5-ones to p-quinone methides.
Symmetry 13 01762 sch028
Scheme 29. Synthesis of spirooxindoles γ-lactones.
Scheme 29. Synthesis of spirooxindoles γ-lactones.
Symmetry 13 01762 sch029
Scheme 30. Michael addition of pyrazol-5-ones to α,β-unsaturated ketones.
Scheme 30. Michael addition of pyrazol-5-ones to α,β-unsaturated ketones.
Symmetry 13 01762 sch030
Figure 1. Cinchona-derived catalysts for asymmetric Michael reactions.
Figure 1. Cinchona-derived catalysts for asymmetric Michael reactions.
Symmetry 13 01762 g001
Scheme 31. Asymmetric Michael difluoromethylation of Meldrum’s acids.
Scheme 31. Asymmetric Michael difluoromethylation of Meldrum’s acids.
Symmetry 13 01762 sch031
Scheme 32. Copper-promoted tandem double Michael reactions.
Scheme 32. Copper-promoted tandem double Michael reactions.
Symmetry 13 01762 sch032
Scheme 33. Cascade cyclization of o-hydroxycinnamaldehydes with diphenylphosphine oxide.
Scheme 33. Cascade cyclization of o-hydroxycinnamaldehydes with diphenylphosphine oxide.
Symmetry 13 01762 sch033
Scheme 34. Addition of Et2Zn to benzaldehyde in the presence of atropisomeric 1-phenylpyrroles.
Scheme 34. Addition of Et2Zn to benzaldehyde in the presence of atropisomeric 1-phenylpyrroles.
Symmetry 13 01762 sch034
Figure 2. Catalysts for asymmetric diethylzinc addition to aldehydes.
Figure 2. Catalysts for asymmetric diethylzinc addition to aldehydes.
Symmetry 13 01762 g002
Scheme 35. Enantioselective addition of diethylzinc to nitrodienes.
Scheme 35. Enantioselective addition of diethylzinc to nitrodienes.
Symmetry 13 01762 sch035
Scheme 36. Enantioselective aldol-type reaction of CF3CHN2 with aldehydes.
Scheme 36. Enantioselective aldol-type reaction of CF3CHN2 with aldehydes.
Symmetry 13 01762 sch036
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rachwalski, M.; Buchcic-Szychowska, A.; Leśniak, S. Recent Advances in Selected Asymmetric Reactions Promoted by Chiral Catalysts: Cyclopropanations, Friedel–Crafts, Mannich, Michael and Other Zinc-Mediated Processes—An Update. Symmetry 2021, 13, 1762. https://doi.org/10.3390/sym13101762

AMA Style

Rachwalski M, Buchcic-Szychowska A, Leśniak S. Recent Advances in Selected Asymmetric Reactions Promoted by Chiral Catalysts: Cyclopropanations, Friedel–Crafts, Mannich, Michael and Other Zinc-Mediated Processes—An Update. Symmetry. 2021; 13(10):1762. https://doi.org/10.3390/sym13101762

Chicago/Turabian Style

Rachwalski, Michał, Aleksandra Buchcic-Szychowska, and Stanisław Leśniak. 2021. "Recent Advances in Selected Asymmetric Reactions Promoted by Chiral Catalysts: Cyclopropanations, Friedel–Crafts, Mannich, Michael and Other Zinc-Mediated Processes—An Update" Symmetry 13, no. 10: 1762. https://doi.org/10.3390/sym13101762

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

Article Metrics

Back to TopTop