Next Article in Journal
Investigating Rubidium Density and Temperature Distributions in a High-Throughput 129Xe-Rb Spin-Exchange Optical Pumping Polarizer
Next Article in Special Issue
Enantioselective Synthesis of Atropisomers by Oxidative Aromatization with Central-to-Axial Conversion of Chirality
Previous Article in Journal
Comparison of Phytochemical Profiles of Wild and Cultivated American Ginseng Using Metabolomics by Ultra-High Performance Liquid Chromatography-High-Resolution Mass Spectrometry
Previous Article in Special Issue
Atroposelective Amination of Indoles via Chiral Center Induced Chiral Axis Formation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in the Asymmetric Synthesis of BINOL Derivatives

by
Everton Machado da Silva
,
Hérika Danielle Almeida Vidal
,
Marcelo Augusto Pereira Januário
and
Arlene Gonçalves Corrêa
*
Centre of Excellence for Research on Sustainable Chemistry, Department of Chemistry, Federal University of São Carlos, São Carlos 13565-905, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(1), 12; https://doi.org/10.3390/molecules28010012
Submission received: 30 November 2022 / Revised: 15 December 2022 / Accepted: 16 December 2022 / Published: 20 December 2022
(This article belongs to the Special Issue Atroposelective Synthesis of Novel Axially Chiral Molecules)

Abstract

:
BINOL derivatives have shown relevant biological activities and are important chiral ligands and catalysts. Due to these properties, their asymmetric synthesis has attracted the interest of the scientific community. In this work, we present an overview of the most efficient methods to obtain chiral BINOLs, highlighting the use of metal complexes and organocatalysts as well as kinetic resolution. Further derivatizations of BINOLs are also discussed.

Graphical Abstract

1. Introduction

The chirality resulting from restricted rotation around a single bond is called atropisomerism (axial chirality). This phenomenon was first described by Christie and Kenner [1] in 1922 when investigating the biaryl 6,6’-dinitro-2,2’-diphenic acid (Figure 1), and the term “atropisomer”, derived from the Greek where “a” means “not” and “tropes” means “turn”, was created by Kuhn. Atropisomers belong to the class of axially chiral compounds; however, in this case, the enantiomers exist due to restricted rotation around a single bond [2].
Axial chirality has also been considered as an important structural element of many natural products [3] and bioactive compounds, whose enantiomers generally exhibit different pharmacological activities and metabolic processes in vivo and in vitro [4]. Examples of natural products include viriditoxin, produced by fungi with antibacterial activity, and vancomycin, an amphoteric glycopeptide antibiotic produced by soil bacterium (Figure 2).
Due to their relevant biological properties, the asymmetric synthesis of atropisomers has attracted the interest of the scientific community. Furthermore, atropisomers have been important chiral ligands since the 1980s, when BINAP was developed for enantioselective reactions catalyzed by transition metals [3].
Although conventional chiral resolution of racemates and chiral auxiliary reactions are generally used for the construction of axially chiral compounds in enantiomerically pure form, asymmetric catalysis meets the demand for high efficiency and economic value [5]. The use of atropisomeric compounds as ligands in metal-mediated catalysis has revolutionized the organometallic chemistry and asymmetric synthesis fields. Due to the high demand and importance of these chiral biaryl scaffolds, several synthetic procedures [6], reviews [7,8,9], and concepts [10] have been published in the literature.
Axially chiral biaryl auxiliaries and catalysts (such as BINOL or BINAP) exhibit excellent chirality transfer properties [11]. Due to the importance of axially chiral biaryl compounds, several interesting methods for their directed atroposelective construction have been developed. Table 1 shows some catalysts used for the transfer of axial chirality, bearing in mind that the substrates used do not have any type of chirality, thus strengthening the excellent transfer of this property.
Herein, we present an overview of the most efficient methods for asymmetric synthesis of chiral BINOLs reported in the last two decades, highlighting the use of metal complexes and organocatalysts, as well as kinetic resolution and further derivatizations.

2. Synthesis of BINOLs Skeleton

2.1. Metal-Mediated Oxidative Enantioselective Coupling

Among the available methods for the synthesis of optically active BINOLs, one of the most explored is the oxidative dimerization of 2-naphthols mediated by complexes of Cu [39,40,41,42,43,44,45,46] Fe [47,48,49], V [50,51,52,53,54,55], Ru [56], and chiral ligands (very often amines), normally generated in situ. In this regard, excellent reviews discussing these methods have been reported by Brunel [57], Wang [58], Bryliakov et al. [59], and Liao et al. [60] (Scheme 1 and Scheme 2).
Oxidative coupling may occur through three different mechanisms: (1) radical–radical coupling, (2) heterolytic coupling of cationic species with 2-naphthol, or (3) radical–anion coupling, the latter generally being the most accepted to support this type of transformation [55,61,62,63,64]. An important step is the one where the catalytic species is complexed with directing groups or coordination assistants at the C3 position of 2-naphthols—notably ester groups [41,45]—which, in many cases, is a “sine qua no” condition for the success of the synthetic protocols (Scheme 3).
With due recognition of the particularities of each case, the radial–anion mechanism [55,61,62,63,64,65] (Scheme 3) usually proceeds via generation of the radical species B resulting from an oxidation of 2-naphthol A by a metal catalyst (Mn+). B is then added to another neutral 2-naphthol molecule to form a new C-C bond and generate the C-radical, which is further oxidized by O2 to restore aromaticity.
The most recent methods for obtaining enantiomerically pure BINOLs are still based on the catalytic dyad metal–chiral ligands. In this sense, Chen and colleagues [66] (Scheme 4) have developed a new chiral 1,5-N,N-bidentate ligand based on a spirocyclic skeleton of pyrrolidine oxazoline and CuBr to couple 2-naphtols 3b. The efficient catalytic species formed in situ allows for (S)-BINOL derivatives (1) with high enantioselectivity (up to 99% ee) and good yields (up to 87%) to be obtained. Based on experimental results and the literature, the authors proposed that this coupling proceeds via radical–anion coupling, where the complex generated in situ coordinates to form species D in the presence of air, which couples with radical E (generated through an electron transfer from the outer sphere with another Cu(II) complex) to form the intermediate F (Scheme 5). The coupling product is obtained after tautomerization of H. The authors found experimental evidence that, during the coupling process, the attack from E to F by the Si face was favored, probably because of the greater steric impediment to attack by the Re face.
Che and co-workers introduced a chiral aminopyridine-like ligand—bisquinolyldiamine [(1R,2R)-N1,N2-di(quinolin-8-yl)cyclohexane-1,2-diamine (BQCN)]— and applied it to the iron-catalyzed asymmetric cis-dihydroxylation of alkenes [67]. Inspired by this work, Liu’s group [68] (Scheme 6) established a methodology for the asymmetric oxidative homo-coupling of 2-naphthols (3c), leading to the synthesis of (S)-BINOL derivatives (1) mediated by a Fe complex and generated in situ from Fe(ClO4)2 and the BQCN ligand. Excellent yields (up to 99%) and enantiomeric excesses (up to 81%) have been reported.
From the same perspective, Uchida’s group [69] (Scheme 7) developed remarkable enantioselective aerobic coupling between 2-naphthols 3d in the presence of the (aqua)ruthenium complex (salen). The protocol provided (R)-BINOLs (1) with yields between 55 and 85% and enantiomeric excesses up to 94%. Through mechanistic studies, these researchers concluded that, in this case, cross-coupling selectivity is dominated by steric rather than electronic effects, which can be controlled by chemoselective oxidation via single electron transfer (SET) and oxidative carbon–carbon bond formation, a process for which ruthenium(salen) catalyst proved to be suitable [62]. Therefore, the authors have proposed that this transformation proceeds via oxidation of one of the coupling partners to the electrophilic intermediate radical I, which is converted to the desired BINOL after chemoselective coupling [62].
Recently, Subramanian et al. [70] (Scheme 8) developed a Cu(II)-2+4-μ4-oxo tetranuclear open frame macrocyclic/BINAN complex and employed it in the asymmetric oxidative coupling of 2-naphthols 3e, obtaining (R)-BINOL derivatives (1) with good to excellent yields (70–96%) and enantiomeric excesses between 68 and 74%.
Ishihara and co-workers [71] (Scheme 9) developed a method for enantioselective oxidative coupling of 2-naphthol derivatives 3d in the presence of a chiral Fe(II)-diphosphine oxide complex. The products of interest were obtained with yields up to 98 % and enantiomeric excesses between 60 and 85%.
A copper catalyst prepared in situ from a ligand synthesized by the fusion of chelating picolinic acid/substituted BINOLs and CuI was employed in the asymmetric oxidative coupling of 2-naphthols (3e). In this work, published by Zhang et al. [72] (Scheme 10), 6,6’-disubstituted (R)-BINOLs (1) were obtained with yields of up to 89% and excellent enantioselectivities (up to 96% ee). The reaction was accompanied by Mass Spectroscopy, and identification of a peak corresponding to the complex J allowed the authors to propose a mechanism pathway through the transition state K.
Continuing work involving multifunctional chiral catalysis via double activation, Takizawa’s group [73] developed complexes A-C (Scheme 11)—from VOSO4 and Schiff base ligands generated via condensation of (S)-tert-leucine and 3,3 ‘-formyl-(R)-BINOL—which have been successfully applied in the synthesis of (R)- and (S)-BINOL (1) with yields between 46 and 76%, in addition to enantiomeric excesses of up to 91%.

2.2. Electrochemical Synthesis

Despite the inherent advantages of electrochemical synthesis, notably in terms of sustainability [74], few examples of enantioselective coupling for the construction of chiral BINOLs have been reported so far. In 1994, which appears to be the first record of this type of synthesis, Bobbitt et al. [75] (Scheme 12) established a method for enantioselective coupling of 2-naphthols (3f) on a TEMPO-modified graphite electrode in the presence of (-)-sparteine in acetonitrile to afford (S)-BINOL (1) with excellent yields and enantiomeric excesses.
Recently, Mei’s group [74] (Scheme 13) demonstrated the first example of a Ni-catalyzed enantioselective electrochemical reductive coupling of 2-naphtols (4) in an undivided cell for the construction of axially chiral BINOL derivatives (1) with good yields (up 91%) and enantiomeric excess of up to 98%.

2.3. Organocatalyzed Synthesis/Kinetic Resolution of BINOLs/BINAPs

Among the methods for the synthesis/kinetic resolution (KR) of axially chiral biaryl and binaphthyl compounds, the organocatalyzed approach has been more explored in the last decade. Some examples were described in the review by Cheng et al. [76]. In this topic, the kinetic synthesis/resolution of BINOL skeletons will be addressed using organocatalysts that do not have axial chirality, thus providing induction to the products of interest.
In 2005, Tsuji and co-workers [77] reported a kinetic resolution of 2,2’-dihydroxy-1,1’-biaryls using a palladium-catalyzed atroposelective alcoholysis of racemic vinyl ethers (5). The method uses the organocatalyst (R,R)-1,2-cyclohexanediamine (6) and a mixture of methanol and dichloromethane as solvent at 20 °C. Five examples were obtained and it was observed that the volume of the acyl group directly influences selectivity, as shown in Scheme 14 where the larger substituent (1-adamantyl) provided high selectivity.
In 2012, Dan and co-workers [78] evaluated the Ferrier-type rearrangement using chiral bicyclic guanidine as a catalyst; however, the reaction was sluggish, affording an optically active product with low yield and enantioselectivity (Scheme 15A). Upon these results, the authors used another strategy, based on studies by Masatoshi and Reiko [79] for synthesis of the optically active biaryl through optical resolution of the corresponding racemate using chiral diamine (12) (Scheme 15B). Initially, there was deprotection of the methyl ether, and in the second step the racemic binaphthol derivative was recrystallized from toluene in the presence of (S,S)-1,2-diphenyl-1,2-ethanediamine (12), leading to compound (R)-11 with 95% ee and 22% yield.
In 2014, Sibi and co-workers [80] proposed a new kinetic resolution that employs a chiral 4-(dimethylamino)pyridine (DMAP) derivative 14 as a catalyst via O-acylation (Scheme 16). The method proved to be efficient, and both secondary alcohols and axially chiral biaryl compounds were obtained with selectivity factors of up to 37 and 51, respectively. Increased conversion was also observed for binaphthyl substrates with an electron-rich group in the ortho position.
In 2014, Zhao and co-workers [81] reported an atroposelective kinetic resolution using N-heterocyclic carbene (NHC) 18 as a catalyst to resolve the enantiomers of BINOL (Scheme 17). The 1,1’-biaryl-2,2’-diol derivatives 16 were explored, where the products obtained, both acylated 19 and the recovered BINOL 16, had high selectivity.
Sparr and Link [82] described a highly enantioselective synthesis of binaphthyl 20 by intramolecular aldol condensation using (S)-pyrrolidinyl-tetrazole 21 as a catalyst (Scheme 18). The authors described that the high selectivity of the process stems from the efficient transfer of stereochemical information from the catalyst into the axis of chirality of biaryl products. The examples obtained showed good yields and high enantiomeric ratios.
In 2017, Shirakawa and co-workers [83] reported a highly enantioselective organocatalytic method for the synthesis of atropisomeric biaryls through cation-directed O-alkylation. Reaction of racemic 1-aryl-2-tetralones (23) with the ammonium salt 24 (obtained from chiral quinidine) under basic conditions and using an alkylation agent leads to highly enantioselective O-alkylation (up to 98:2 er). According to the proposed mechanism, the basic medium initially promotes deprotonation and makes it possible to generate two enolate enantiomers associated with the chiral salt that form diastereomeric ion pairs; however, the chiral ammonium counterion is capable of rapidly differentiating balancer atropisomeric enolates, leading to highly atropselective O-alkylation. The in situ oxidation step with 2,3-Dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) then takes place to obtain the BINOL derivatives (25) without loss of the enantiomeric ratio. The authors also noted that enantioselectivity can be controlled by the catalyst structure and the type of base and solvent used. Under optimized conditions, they obtained a broad scope of 20 examples with moderate to excellent yields (Scheme 19).
In 2019, Sparr and co-workers [84] described a non-canonical polyketide cyclization, affording atropisomeric tetra-ortho-substituted binaphthalenes. Using hexacarbonyl substrates 26 and aminoethanol-derived proline-based catalyst 27 via the cascade of two arene ring formation reactions (Scheme 20), it was possible to obtain enantioenriched tetra-ortho-substituted binaphthalenes 28 through atroposelective aldol condensation. The explanation of the mechanism was based on NMR studies, where it was observed that two sequential aldol additions take place before a double dehydration, hence alleviating the acute nonbonding interaction during C−C bond formation. With product 28 (without substituents), it was possible to perform a derivatization through a triflation and arene coupling, obtaining (S)-29 with 84% yield (over two steps) and >99:1 er.
Based on a previous work, where high atroposelectivity was acquired starting from racemic 2-tetralones, Jones et al. [85] extended the strategy for the use of BINOLs (Scheme 21). The reaction proceeds under basic conditions using cinchona derived catalyst 32 in a mixture of toluene and ethyl ether at room temperature for 48 h, leading to the formation of compound (R)-30 with 47% yield and 96% ee and compound (S)-31 with 49% yield and 80% ee.
The reversible deprotonation of compound 30 leads to the formation of a diastereoisomeric BINOlate ammonium salt, which reacts at different rates in the alkylation step. The proposed transition state involves a hydrogen bonding of the ammonium salt with the secondary alcohol of the BINOlate anion, and an additional hydrogen bonding of the benzyl electrophile to the methyl ether.

2.4. Enzymatic Kinetic Resolution of BINOL

For decades, enzymatic kinetic resolution was considered the most reliable strategy for obtaining optically enriched compounds. In 1989, the first efficient method for enzymatic resolution of rac-BINOL (33) was described by Kazlaukas [86], which was based on enantiospecific hydrolysis catalyzed by cholesterol esterase (Scheme 22). The procedure uses low-cost bovine pancreatic acetone powder (PAP). The enantiomer (S)-1 was obtained with 66% yield and 99% ee and recrystallized with toluene, with compound (R)-33 formed after hydrolysis; enantiomer (R)-1 was obtained with 63% yield and 99% ee after filtration and being washed with cold toluene.
The potential of enantioselective kinetic resolution to prepare atropisomeric compounds was initially proven by designing an enzymatic kinetic resolution, as illustrated by the elegant work of Aoyagi et al. [87]. Lipase (Candida antarctica) was used to catalyze the hydrolysis of BINOL monoester-derivatives 34, affording (R)-1 and (S)-34 with excellent yields and good enantioselective excesses (Scheme 23).
In 2018, Moustafa and co-workers [88] described the lipase-catalyzed KR process, which uses immobilized Pseudomonas sp. lipoprotein lipase (Toyobo LIP301). Lipase selectively catalyzes readily available racemic substrates and provides stability to acylated products against racemization, promoting the formation of product (R)-35 with 24% yield and 99% enantiomeric excess (Scheme 24).

2.5. Chemical Derivatizations on the BINOL Skeleton

As previously mentioned, most chiral catalysts with the BINOL backbone are now commercially available. In this section, we will discuss the synthesis of new chiral catalysts.
Since 1999, Maruoka’s group [89,90] (Scheme 25) has synthesized a series of chiral phase transfer catalysts based on quaternary ammonium bromides salts prepared from 1-(S)-BINOL, which have been successfully applied in the synthesis of natural and unnatural amino acids. In the same context, the group reported, in 2013 [12], the preparation of catalyst 40 (via insertion of piperazine 37) from the brominated strategic derivative 36, with 49% yield. In this case, the new chiral auxiliary was employed in the asymmetric phase transfer functionalization of 1-alkylalene-1,3-dicarboxylates with N-arylsulfonyl imines and allylic/benzyl bromides for the preparation of tetrasubstituted allenes.
Recently, Schaus’ group [91] designed and prepared the (S)-3,3’,6,6’-tetrakis(trifluoromethyl)-BINOL 42 from (S)-1 by the subsequent protection, bromination, and insertion of the CF3 groups into the desired position (Scheme 26). This chiral catalyst was employed in the asymmetric synthesis of 1,3-substituted chiral allenes via boronate addition to sulfonyl hydrazones.
For the catalysts based on BINOL derivatives with phosphoric acid, the most efficient method described in the literature for the insertion of phosphoric acid is phosphorylation with POCl3 (Scheme 27), which can be performed in both enantiomerically pure and racemic BINOL [92,93].

3. Conclusions

There have been significant advances in the aerobic enantioselective coupling methods of 2-naphthols for the synthesis/kinetic resolution of chiral BINOLs via transition metal-mediated, electrochemical, organocatalytic, and enzymatic resolution. However, it was evident that there are still challenges to be overcome in both fields, such as the relatively high reaction time in some cases and the use of toxic reagents and solvents such as dichloromethane and toluene. In addition, as already highlighted by other authors, the mechanistic discussions that govern asymmetric induction in these cases are still preambular, denoting the need for deepened theoretical exploration in this particular area.

Author Contributions

Writing—original draft preparation, review, and editing, E.M.d.S., H.D.A.V. and M.A.P.J.; manuscript review, editing, and supervision and funding acquisition, A.G.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge FAPESP (grants 2013/07600-3 and 2014/50249-8), GlaxoSmithKline, CAPES (Finance Code 001), and CNPq (grants 429748/2018-3 and 302140/2019-0) for funding and fellowships.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Christie, G.H.; Kenner, J. LXXI.—The Molecular Configurations of Polynuclear Aromatic Compounds. Part I. The Resolution of γ-6 : 6′-Dinitro- and 4 : 6 : 4′ : 6′-Tetranitro-Diphenic Acids into Optically Active Components. J. Chem. Soc. Trans. 1922, 121, 614–620. [Google Scholar] [CrossRef]
  2. Kumarasamy, E.; Raghunathan, R.; Sibi, M.P.; Sivaguru, J. Nonbiaryl and Heterobiaryl Atropisomers: Molecular Templates with Promise for Atropselective Chemical Transformations. Chem. Rev. 2015, 115, 11239–11300. [Google Scholar] [CrossRef]
  3. Bringmann, G.; Gulder, T.; Gulder, T.A.M.; Breuning, M. Atroposelective Total Synthesis of Axially Chiral Biaryl Natural Products. Chem. Rev. 2011, 111, 563–639. [Google Scholar] [CrossRef]
  4. Zhang, X.; Zhao, K.; Gu, Z. Transition Metal-Catalyzed Biaryl Atropisomer Synthesis via a Torsional Strain Promoted Ring-Opening Reaction. Acc. Chem. Res. 2022, 55, 1620–1633. [Google Scholar] [CrossRef]
  5. Wang, Y.-B.; Tan, B. Construction of Axially Chiral Compounds via Asymmetric Organocatalysis. Acc. Chem. Res. 2018, 51, 534–547. [Google Scholar] [CrossRef]
  6. Bringmann, G.; Menche, D. Stereoselective Total Synthesis of Axially Chiral Natural Products via Biaryl Lactones. Acc. Chem. Res. 2001, 34, 615–624. [Google Scholar] [CrossRef] [PubMed]
  7. Rosini, C.; Franzini, L.; Raffaelli, A.; Salvadori, P. Synthesis and Applications of Binaphthylic C2-Symmetry Derivatives as Chiral Auxiliaries in Enantioselective Reactions. Synthesis 1992, 1992, 503–517. [Google Scholar] [CrossRef]
  8. Wencel-Delord, J.; Panossian, A.; Leroux, F.R.; Colobert, F. Recent Advances and New Concepts for the Synthesis of Axially Stereoenriched Biaryls. Chem. Soc. Rev. 2015, 44, 3418–3430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Mei, G.-J.; Koay, W.L.; Guan, C.-Y.; Lu, Y. Atropisomers beyond the C–C Axial Chirality: Advances in Catalytic Asymmetric Synthesis. Chem 2022, 8, 1855–1893. [Google Scholar] [CrossRef]
  10. LaPlante, S.R.; Edwards, P.J.; Fader, L.D.; Jakalian, A.; Hucke, O. Cover Picture: Revealing Atropisomer Axial Chirality in Drug Discovery. ChemMedChem 2011, 6, 381. [Google Scholar] [CrossRef]
  11. McCarthy, M.; Guiry, P.J. Axially Chiral Bidentate Ligands in Asymmetric Catalysis. Tetrahedron 2001, 57, 3809–3844. [Google Scholar] [CrossRef]
  12. Hashimoto, T.; Sakata, K.; Tamakuni, F.; Dutton, M.J.; Maruoka, K. Phase-Transfer-Catalysed Asymmetric Synthesis of Tetrasubstituted Allenes. Nat. Chem. 2013, 5, 240–244. [Google Scholar] [CrossRef]
  13. Li, X.; Sun, J. Organocatalytic Enantioselective Synthesis of Chiral Allenes: Remote Asymmetric 1,8-Addition of Indole Imine Methides. Angew. Chem. Int. Ed. 2020, 59, 17049–17054. [Google Scholar] [CrossRef]
  14. Woldegiorgis, A.G.; Han, Z.; Lin, X. Organocatalytic Asymmetric Dearomatization Reaction for the Synthesis of Axial Chiral Allene-Derived Naphthalenones Bearing Quaternary Stereocenters. Org. Lett. 2021, 23, 6606–6611. [Google Scholar] [CrossRef]
  15. Bai, H.-Y.; Tan, F.-X.; Liu, T.-Q.; Zhu, G.-D.; Tian, J.-M.; Ding, T.-M.; Chen, Z.-M.; Zhang, S.-Y. Highly Atroposelective Synthesis of Nonbiaryl Naphthalene-1,2-Diamine N-C Atropisomers through Direct Enantioselective C-H Amination. Nat. Commun. 2019, 10, 3063. [Google Scholar] [CrossRef] [Green Version]
  16. Chen, M.; Qian, D.; Sun, J. Organocatalytic Enantioconvergent Synthesis of Tetrasubstituted Allenes via Asymmetric 1,8-Addition to Aza- Para -Quinone Methides. Org. Lett. 2019, 21, 8127–8131. [Google Scholar] [CrossRef]
  17. Li, F.; Liang, S.; Luan, Y.; Chen, X.; Zhao, H.; Huang, A.; Li, P.; Li, W. Organocatalytic Regio-, Diastereo- and Enantioselective γ-Additions of Isoxazol-5(4 H )-Ones to β,γ-Alkynyl-α-Imino Esters for the Synthesis of Axially Chiral Tetrasubstituted α-Amino Allenoates. Org. Chem. Front. 2021, 8, 1243–1248. [Google Scholar] [CrossRef]
  18. Tap, A.; Blond, A.; Wakchaure, V.N.; List, B. Chiral Allenes via Alkynylogous Mukaiyama Aldol Reaction. Angew. Chem. Int. Ed. 2016, 55, 8962–8965. [Google Scholar] [CrossRef]
  19. Wang, J.; Chen, M.-W.; Ji, Y.; Hu, S.-B.; Zhou, Y.-G. Kinetic Resolution of Axially Chiral 5- or 8-Substituted Quinolines via Asymmetric Transfer Hydrogenation. J. Am. Chem. Soc. 2016, 138, 10413–10416. [Google Scholar] [CrossRef]
  20. Zhu, W.-R.; Su, Q.; Diao, H.-J.; Wang, E.-X.; Wu, F.; Zhao, Y.-L.; Weng, J.; Lu, G. Enantioselective Dehydrative γ-Arylation of α-Indolyl Propargylic Alcohols with Phenols: Access to Chiral Tetrasubstituted Allenes and Naphthopyrans. Org. Lett. 2020, 22, 6873–6878. [Google Scholar] [CrossRef]
  21. Mori, K.; Itakura, T.; Akiyama, T. Enantiodivergent Atroposelective Synthesis of Chiral Biaryls by Asymmetric Transfer Hydrogenation: Chiral Phosphoric Acid Catalyzed Dynamic Kinetic Resolution. Angew. Chem. 2016, 128, 11814–11818. [Google Scholar] [CrossRef]
  22. Wang, Z.; Lin, X.; Chen, X.; Li, P.; Li, W. Organocatalytic Stereoselective 1,6-Addition of Thiolacetic Acids to Alkynyl Indole Imine Methides: Access to Axially Chiral Sulfur-Containing Tetrasubstituted Allenes. Org. Chem. Front. 2021, 8, 3469–3474. [Google Scholar] [CrossRef]
  23. Zhang, P.; Huang, Q.; Cheng, Y.; Li, R.; Li, P.; Li, W. Remote Stereocontrolled Construction of Vicinal Axially Chiral Tetrasubstituted Allenes and Heteroatom-Functionalized Quaternary Carbon Stereocenters. Org. Lett. 2019, 21, 503–507. [Google Scholar] [CrossRef] [PubMed]
  24. Li, G.-Q.; Gao, H.; Keene, C.; Devonas, M.; Ess, D.H.; Kürti, L. Organocatalytic Aryl–Aryl Bond Formation: An Atroposelective [3,3]-Rearrangement Approach to BINAM Derivatives. J. Am. Chem. Soc. 2013, 135, 7414–7417. [Google Scholar] [CrossRef] [PubMed]
  25. Qi, L.-W.; Mao, J.-H.; Zhang, J.; Tan, B. Organocatalytic Asymmetric Arylation of Indoles Enabled by Azo Groups. Nat. Chem. 2018, 10, 58–64. [Google Scholar] [CrossRef] [PubMed]
  26. Ma, C.; Sheng, F.-T.; Wang, H.-Q.; Deng, S.; Zhang, Y.-C.; Jiao, Y.; Tan, W.; Shi, F. Atroposelective Access to Oxindole-Based Axially Chiral Styrenes via the Strategy of Catalytic Kinetic Resolution. J. Am. Chem. Soc. 2020, 142, 15686–15696. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, Y.-H.; Qi, L.-W.; Fang, F.; Tan, B. Organocatalytic Atroposelective Arylation of 2-Naphthylamines as a Practical Approach to Axially Chiral Biaryl Amino Alcohols. Angew. Chem. Int. Ed. 2017, 56, 16308–16312. [Google Scholar] [CrossRef]
  28. Hu, Y.-L.; Wang, Z.; Yang, H.; Chen, J.; Wu, Z.-B.; Lei, Y.; Zhou, L. Conversion of Two Stereocenters to One or Two Chiral Axes: Atroposelective Synthesis of 2,3-Diarylbenzoindoles. Chem. Sci. 2019, 10, 6777–6784. [Google Scholar] [CrossRef] [Green Version]
  29. Man, N.; Lou, Z.; Li, Y.; Yang, H.; Zhao, Y.; Fu, H. Organocatalytic Atroposelective Construction of Axially Chiral N-Aryl Benzimidazoles Involving Carbon–Carbon Bond Cleavage. Org. Lett. 2020, 22, 6382–6387. [Google Scholar] [CrossRef]
  30. Bisag, G.D.; Pecorari, D.; Mazzanti, A.; Bernardi, L.; Fochi, M.; Bencivenni, G.; Bertuzzi, G.; Corti, V. Central-to-Axial Chirality Conversion Approach Designed on Organocatalytic Enantioselective Povarov Cycloadditions: First Access to Configurationally Stable Indole–Quinoline Atropisomers. Chem. A Eur. J. 2019, 25, 15694–15701. [Google Scholar] [CrossRef]
  31. Kwon, Y.; Li, J.; Reid, J.P.; Crawford, J.M.; Jacob, R.; Sigman, M.S.; Toste, F.D.; Miller, S.J. Disparate Catalytic Scaffolds for Atroposelective Cyclodehydration. J. Am. Chem. Soc. 2019, 141, 6698–6705. [Google Scholar] [CrossRef]
  32. Lu, D.-L.; Chen, Y.-H.; Xiang, S.-H.; Yu, P.; Tan, B.; Li, S. Atroposelective Construction of Arylindoles by Chiral Phosphoric Acid-Catalyzed Cross-Coupling of Indoles and Quinones. Org. Lett. 2019, 21, 6000–6004. [Google Scholar] [CrossRef]
  33. Wang, C.; Li, T.; Liu, S.; Zhang, Y.; Deng, S.; Jiao, Y.; Shi, F. Axially Chiral Aryl-Alkene-Indole Framework: A Nascent Member of the Atropisomeric Family and Its Catalytic Asymmetric Construction. Chin. J. Chem. 2020, 38, 543–552. [Google Scholar] [CrossRef]
  34. Zhang, H.; Wang, C.; Li, C.; Mei, G.; Li, Y.; Shi, F. Design and Enantioselective Construction of Axially Chiral Naphthyl-Indole Skeletons. Angew. Chem. 2017, 129, 122–127. [Google Scholar] [CrossRef]
  35. Lu, S.; Ng, S.V.H.; Lovato, K.; Ong, J.-Y.; Poh, S.B.; Ng, X.Q.; Kürti, L.; Zhao, Y. Practical Access to Axially Chiral Sulfonamides and Biaryl Amino Phenols via Organocatalytic Atroposelective N-Alkylation. Nat. Commun. 2019, 10, 3061. [Google Scholar] [CrossRef] [Green Version]
  36. Jiang, F.; Chen, K.; Wu, P.; Zhang, Y.; Jiao, Y.; Shi, F. A Strategy for Synthesizing Axially Chiral Naphthyl-Indoles: Catalytic Asymmetric Addition Reactions of Racemic Substrates. Angew. Chem. Int. Ed. 2019, 58, 15104–15110. [Google Scholar] [CrossRef]
  37. Mori, K.; Ichikawa, Y.; Kobayashi, M.; Shibata, Y.; Yamanaka, M.; Akiyama, T. Enantioselective Synthesis of Multisubstituted Biaryl Skeleton by Chiral Phosphoric Acid Catalyzed Desymmetrization/Kinetic Resolution Sequence. J. Am. Chem. Soc. 2013, 135, 3964–3970. [Google Scholar] [CrossRef]
  38. Vaidya, S.D.; Toenjes, S.T.; Yamamoto, N.; Maddox, S.M.; Gustafson, J.L. Catalytic Atroposelective Synthesis of N-Aryl Quinoid Compounds. J. Am. Chem. Soc. 2020, 142, 2198–2203. [Google Scholar] [CrossRef]
  39. Feringa, B.; Wynberg, H. Biomimetic Asymmetric Oxidative Coupling of Phenols. Bioorg. Chem. 1978, 7, 397–408. [Google Scholar] [CrossRef] [Green Version]
  40. Brussee, J.; Groenendijk, J.L.G.; te Koppele, J.M.; Jansen, A.C.A. On the Mechanism of the Formation of s(−)-(1, 1’-Binaphthalene)-2,2’-Diol via Copper(II)Amine Complexes. Tetrahedron 1985, 41, 3313–3319. [Google Scholar] [CrossRef]
  41. Nakajima, M.; Miyoshi, I.; Kanayama, K.; Hashimoto, S.; Noji, M.; Koga, K. Enantioselective Synthesis of Binaphthol Derivatives by Oxidative Coupling of Naphthol Derivatives Catalyzed by Chiral Diamine Copper Complexes. J. Org. Chem. 1999, 64, 2264–2271. [Google Scholar] [CrossRef]
  42. Caselli, A.; Giovenzana, G.B.; Palmisano, G.; Sisti, M.; Pilati, T. Synthesis of C2-Symmetrical Diamine Based on (1R)-(+)-Camphor and Application to Oxidative Aryl Coupling of Naphthols. Tetrahedron Asymmetry 2003, 14, 1451–1454. [Google Scholar] [CrossRef]
  43. Gao, J.; Reibenspies, J.H.; Martell, A.E. Structurally Defined Catalysts for Enantioselective Oxidative Coupling Reactions. Angew. Chem. Int. Ed. 2003, 42, 6008–6012. [Google Scholar] [CrossRef] [PubMed]
  44. Li, X.; Hewgley, J.B.; Mulrooney, C.A.; Yang, J.; Kozlowski, M.C. Enantioselective Oxidative Biaryl Coupling Reactions Catalyzed by 1,5-Diazadecalin Metal Complexes:  Efficient Formation of Chiral Functionalized BINOL Derivatives. J. Org. Chem. 2003, 68, 5500–5511. [Google Scholar] [CrossRef]
  45. Kim, K.H.; Lee, D.-W.; Lee, Y.-S.; Ko, D.-H.; Ha, D.-C. Enantioselective Oxidative Coupling of Methyl 3-Hydroxy-2-Naphthoate Using Mono-N-Alkylated Octahydrobinaphthyl-2,2′-Diamine Ligand. Tetrahedron 2004, 60, 9037–9042. [Google Scholar] [CrossRef]
  46. Temma, T.; Habaue, S. Highly Selective Oxidative Cross-Coupling of 2-Naphthol Derivatives with Chiral Copper(I)–Bisoxazoline Catalysts. Tetrahedron Lett. 2005, 46, 5655–5657. [Google Scholar] [CrossRef]
  47. Zhang, Q.; Cui, X.; Chen, L.; Liu, H.; Wu, Y. Syntheses of Chiral Ferrocenophanes and Their Application to Asymmetric Catalysis. Eur. J. Org. Chem. 2014, 2014, 7823–7829. [Google Scholar] [CrossRef]
  48. Egami, H.; Katsuki, T. Iron-Catalyzed Asymmetric Aerobic Oxidation: Oxidative Coupling of 2-Naphthols. J. Am. Chem. Soc. 2009, 131, 6082–6083. [Google Scholar] [CrossRef]
  49. Tkachenko, N.V.; Lyakin, O.Y.; Samsonenko, D.G.; Talsi, E.P.; Bryliakov, K.P. Highly Efficient Asymmetric Aerobic Oxidative Coupling of 2-Naphthols in the Presence of Bioinspired Iron Aminopyridine Complexes. Catal. Commun. 2018, 104, 112–117. [Google Scholar] [CrossRef]
  50. Hon, S.-W.; Li, C.-H.; Kuo, J.-H.; Barhate, N.B.; Liu, Y.-H.; Wang, Y.; Chen, C.-T. Catalytic Asymmetric Coupling of 2-Naphthols by Chiral Tridentate Oxovanadium(IV) Complexes. Org. Lett. 2001, 3, 869–872. [Google Scholar] [CrossRef]
  51. Chu, C.-Y.; Hwang, D.-R.; Wang, S.-K.; Uang, B.-J. Chiral Oxovanadium Complex Catalyzed Enantioselective Oxidative Coupling of 2-Naphthols. Chem. Commun. 2001, 11, 980–981. [Google Scholar] [CrossRef] [Green Version]
  52. Barhate, N.B.; Chen, C.-T. Catalytic Asymmetric Oxidative Couplings of 2-Naphthols by Tridentate N-Ketopinidene-Based Vanadyl Dicarboxylates. Org. Lett. 2002, 4, 2529–2532. [Google Scholar] [CrossRef]
  53. Luo, Z.; Liu, Q.; Gong, L.; Cui, X.; Mi, A.; Jiang, Y. The Rational Design of Novel Chiral Oxovanadium(IV) Complexes for Highly Enantioselective Oxidative Coupling of 2-Naphthols. Chem. Commun. 2002, 8, 914–915. [Google Scholar] [CrossRef]
  54. Somei, H.; Asano, Y.; Yoshida, T.; Takizawa, S.; Yamataka, H.; Sasai, H. Dual Activation in a Homolytic Coupling Reaction Promoted by an Enantioselective Dinuclear Vanadium(IV) Catalyst. Tetrahedron Lett. 2004, 45, 1841–1844. [Google Scholar] [CrossRef]
  55. Guo, Q.-X.; Wu, Z.-J.; Luo, Z.-B.; Liu, Q.-Z.; Ye, J.-L.; Luo, S.-W.; Cun, L.-F.; Gong, L.-Z. Highly Enantioselective Oxidative Couplings of 2-Naphthols Catalyzed by Chiral Bimetallic Oxovanadium Complexes with Either Oxygen or Air as Oxidant. J. Am. Chem. Soc. 2007, 129, 13927–13938. [Google Scholar] [CrossRef]
  56. Tanaka, H.; Nishikawa, H.; Uchida, T.; Katsuki, T. Photopromoted Ru-Catalyzed Asymmetric Aerobic Sulfide Oxidation and Epoxidation Using Water as a Proton Transfer Mediator. J. Am. Chem. Soc. 2010, 132, 12034–12041. [Google Scholar] [CrossRef] [PubMed]
  57. Brunel, J.M. BINOL: A Versatile Chiral Reagent. Chem. Rev. 2005, 105, 857–898. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, H. Recent Advances in Asymmetric Oxidative Coupling of 2-Naphthol and Its Derivatives. Chirality 2010, 22, 827–837. [Google Scholar] [CrossRef]
  59. Tkachenko, N.V.; Bryliakov, K.P. Transition Metal Catalyzed Aerobic Asymmetric Coupling of 2-Naphthols. Mini. Rev. Org. Chem. 2019, 16, 392–398. [Google Scholar] [CrossRef]
  60. Liao, G.; Zhou, T.; Yao, Q.-J.; Shi, B.-F. Recent Advances in the Synthesis of Axially Chiral Biaryls via Transition Metal-Catalysed Asymmetric C–H Functionalization. Chem. Commun. 2019, 55, 8514–8523. [Google Scholar] [CrossRef]
  61. Matsushita, M.; Kamata, K.; Yamaguchi, K.; Mizuno, N. Heterogeneously Catalyzed Aerobic Oxidative Biaryl Coupling of 2-Naphthols and Substituted Phenols in Water. J. Am. Chem. Soc. 2005, 127, 6632–6640. [Google Scholar] [CrossRef]
  62. Egami, H.; Matsumoto, K.; Oguma, T.; Kunisu, T.; Katsuki, T. Enantioenriched Synthesis of C1-Symmetric BINOLs: Iron-Catalyzed Cross-Coupling of 2-Naphthols and Some Mechanistic Insight. J. Am. Chem. Soc. 2010, 132, 13633–13635. [Google Scholar] [CrossRef] [PubMed]
  63. Narute, S.; Parnes, R.; Toste, F.D.; Pappo, D. Enantioselective Oxidative Homocoupling and Cross-Coupling of 2-Naphthols Catalyzed by Chiral Iron Phosphate Complexes. J. Am. Chem. Soc. 2016, 138, 16553–16560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Libman, A.; Shalit, H.; Vainer, Y.; Narute, S.; Kozuch, S.; Pappo, D. Synthetic and Predictive Approach to Unsymmetrical Biphenols by Iron-Catalyzed Chelated Radical–Anion Oxidative Coupling. J. Am. Chem. Soc. 2015, 137, 11453–11460. [Google Scholar] [CrossRef] [PubMed]
  65. Shibasaki, M.; Matsunaga, S. Design and Application of Linked-BINOL Chiral Ligands in Bifunctional Asymmetric Catalysis. Chem. Soc. Rev. 2006, 35, 269. [Google Scholar] [CrossRef] [PubMed]
  66. Tian, J.; Wang, A.; Yang, J.; Zhao, X.; Tu, Y.; Zhang, S.; Chen, Z. Copper-Complex-Catalyzed Asymmetric Aerobic Oxidative Cross-Coupling of 2-Naphthols: Enantioselective Synthesis of 3,3′-Substituted C1 -Symmetric BINOLs. Angew. Chem. 2019, 131, 11139–11143. [Google Scholar] [CrossRef]
  67. Zang, C.; Liu, Y.; Xu, Z.-J.; Tse, C.-W.; Guan, X.; Wei, J.; Huang, J.-S.; Che, C.-M. Highly Enantioselective Iron-Catalyzed Cis -Dihydroxylation of Alkenes with Hydrogen Peroxide Oxidant via an Fe III -OOH Reactive Intermediate. Angew. Chem. 2016, 128, 10409–10413. [Google Scholar] [CrossRef]
  68. Wu, L.Y.; Usman, M.; Liu, W.B. Enantioselective Iron/Bisquinolyldiamine Ligand-catalyzed Oxidative Coupling Reaction of 2-naphthols. Molecules 2020, 25, 852. [Google Scholar] [CrossRef] [Green Version]
  69. Hayashi, H.; Ueno, T.; Kim, C.; Uchida, T. Ruthenium-Catalyzed Cross-Selective Asymmetric Oxidative Coupling of Arenols. Org. Lett. 2020, 22, 1469–1474. [Google Scholar] [CrossRef]
  70. Chinnaraja, E.; Arunachalam, R.; Pillai, R.S.; Peuronen, A.; Rissanen, K.; Subramanian, P.S. One-Pot Synthesis of [2+2]-Helicate-like Macrocycle and 2+4-Μ4-Oxo Tetranuclear Open Frame Complexes: Chiroptical Properties and Asymmetric Oxidative Coupling of 2-Naphthols. Appl. Organomet. Chem. 2020, 34, e5666. [Google Scholar] [CrossRef]
  71. Horibe, T.; Nakagawa, K.; Hazeyama, T.; Takeda, K.; Ishihara, K. An Enantioselective Oxidative Coupling Reaction of 2-Naphthol Derivatives Catalyzed by Chiral Diphosphine Oxide-Iron(II) Complexes. Chem. Commun. 2019, 55, 13677–13680. [Google Scholar] [CrossRef]
  72. Wang, P.; Cen, S.; Gao, J.; Shen, A.; Zhang, Z. Novel Axially Chiral Ligand-Enabled Copper-Catalyzed Asymmetric Oxidative Coupling of 2-Naphthols for the Synthesis of 6,6′-Disubstituted BINOLs. Org. Lett. 2022, 24, 2321–2326. [Google Scholar] [CrossRef] [PubMed]
  73. Kumar, A.; Sasai, H.; Takizawa, S. Atroposelective Synthesis of C-C Axially Chiral Compounds via Mono- and Dinuclear Vanadium Catalysis. Acc. Chem. Res. 2022, 55, 2949–2965. [Google Scholar] [CrossRef]
  74. Qiu, H.; Shuai, B.; Wang, Y.Z.; Liu, D.; Chen, Y.G.; Gao, P.-S.; Ma, H.X.; Ma, H.X.; Chen, S.; Mei, T.S. Enantioselective Ni-Catalyzed Electrochemical Synthesis of Biaryl Atropisomers. J. Am. Chem. Soc. 2020, 142, 9872–9878. [Google Scholar] [CrossRef] [PubMed]
  75. Osa, T.; Kashiwagi, Y.; Yanagisawa, Y.; Bobbitt, J.M. Enantioselective, Electrocatalytic Oxidative Coupling of Naphthol, Naphthyl Ether and Phenanthrol on a TEMPO-Modified Graphite Felt Electrode in the Presence of (-)-Sparteine (TEMPO = 2,2,6,6-Tetramethylpiperidin-1-Yloxyl). J. Chem. Soc. Chem. Commun. 1994, 1, 2535–2537. [Google Scholar] [CrossRef]
  76. Cheng, J.K.; Xiang, S.-H.; Tan, B. Organocatalytic Enantioselective Synthesis of Axially Chiral Molecules: Development of Strategies and Skeletons. Acc. Chem. Res. 2022, 55, 2920–2937. [Google Scholar] [CrossRef]
  77. Aoyama, H.; Tokunaga, M.; Kiyosu, J.; Iwasawa, T.; Obora, Y.; Tsuji, Y. Kinetic Resolution of Axially Chiral 2,2‘-Dihydroxy-1,1‘-biaryls by Palladium-Catalyzed Alcoholysis. J. Am. Chem. Soc. 2005, 127, 10474–10475. [Google Scholar] [CrossRef]
  78. Terada, M.; Dan, K. Synthesis of unsymmetrically substituted 2,2′ -dihydroxy-1,1′-biaryl derivatives using organic-base-catalyzed Ferrier-type rearrangement as the key step. Chem. Commun. 2012, 48, 5781–5783. [Google Scholar] [CrossRef]
  79. Kawashima, M.; Hirata, R. Epimerization-Crystallization Method in Optical Resolution of 2,2′-dihydroxy-1,1′-binaphthyl, and Kinetic Study. Bull. Chem. Soc. Jpn. 1993, 66, 2002–2005. [Google Scholar] [CrossRef]
  80. Ma, G.; Deng, J.; Sibi, M.P. Fluxionally Chiral DMAP Catalysts: Kinetic Resolution of Axially Compounds. Angew. Chem. Int. Ed. 2014, 53, 11818–11821. [Google Scholar] [CrossRef]
  81. Lu, S.; Poh, S.B.; Zhao, Y. Kinetic Resolution of 1,1′-Biaryl-2,2′-Diols and Amino Alcohols through NHC-Catalyzed Atroposelective Acylation. Angew. Chem. Int. Ed. 2014, 53, 11041–11045. [Google Scholar] [CrossRef] [PubMed]
  82. Link, A.; Sparr, C. Organocatalytic Atroposelective Aldol Condensation: Synthesis of Axially Chiral Biaryls by Arene Formation. Angew. Chem. Int. Ed. 2014, 53, 5458–5461. [Google Scholar] [CrossRef]
  83. Shirakawa, S.; Maruoka, K. A New Strategy for Organocatalyzed Asymmetric Synthesis of BINOL Derivatives. Chem 2017, 2, 329–331. [Google Scholar] [CrossRef] [Green Version]
  84. Witzig, R.M.; Fäseke, V.C.; Häussinger, D.; Sparr, C. Atroposelective Synthesis of Tetra-Ortho-Substituted Biaryls by Catalyst-Controlled Non-Canonical Polyketide Cyclizations. Nat. Catal. 2019, 2, 925–930. [Google Scholar] [CrossRef]
  85. Jones, B.A.; Balan, T.; Jolliffe, J.D.; Campbell, C.D.; Smith, M.D. Practical and Scalable Kinetic Resolution of BINOLs Mediated by a Chiral Counterion. Angew. Chem. Int. Ed. 2019, 58, 4596–4600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Kazlauskas, R.J. Resolution of Binaphthols and Spirobiindanols Using Cholesterol Esterase. J. Am. Chem. Soc. 1989, 111, 4953–4959. [Google Scholar] [CrossRef]
  87. Aoyagi, N.; Ogawa, N.; Izumi, T. Effects of Reaction Temperature and Acyl Group for Lipase-Catalyzed Chiral Binaphthol Synthesis. Tetrahedron Lett. 2006, 47, 4797–4801. [Google Scholar] [CrossRef]
  88. Moustafa, G.A.I.; Oki, Y.; Akai, S. Lipase-Catalyzed Dynamic Kinetic Resolution of C1- and C2-Symmetric Racemic Axially Chiral 2,2′-Dihydroxy-1,1′-Biaryls. Angew. Chem. Int. Ed. 2018, 57, 10278–10282. [Google Scholar] [CrossRef]
  89. Ooi, T.; Kameda, M.; Maruoka, K. Molecular Design of a C2-Symmetric Chiral Phase-Transfer Catalyst for Practical Asymmetric Synthesis of α-Amino Acids. J. Am. Chem. Soc. 1999, 121, 6519–6520. [Google Scholar] [CrossRef]
  90. Ooi, T.; Kameda, M.; Maruoka, K. Design of N-Spiro C2-Symmetric Chiral Quaternary Ammonium Bromides as Novel Chiral Phase-Transfer Catalysts: Synthesis and Application to Practical Asymmetric Synthesis of r -Amino Acids. J. Am. Chem. Soc. 2003, 125, 5139–5151. [Google Scholar] [CrossRef]
  91. Jiang, Y.; Diagne, A.B.; Thomson, R.J.; Schaus, S.E. Enantioselective Synthesis of Allenes by Catalytic Traceless Petasis Reactions. J. Am. Chem. Soc. 2017, 139, 1998–2005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Nakashima, D.; Yamamoto, H. Design of Chiral N-Triflyl Phosphoramide as a Strong Chiral Brønsted Acid and Its Application to Asymmetric Diels-Alder Reaction. J. Am. Chem. Soc. 2006, 128, 9626–9627. [Google Scholar] [CrossRef] [PubMed]
  93. Rueping, M.; Nachtsheim, B.J.; Koenigs, R.M.; Ieawsuwan, W. Synthesis and Structural Aspects of N-Triflylphosphoramides and Their Calcium Salts’highly Acidic and Effective Brønsted Acids. Chem. A Eur. J. 2010, 16, 13116–13126. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Enantiomers of the biaryl 6,6’-dinitro-2,2’-diphenic acid.
Figure 1. Enantiomers of the biaryl 6,6’-dinitro-2,2’-diphenic acid.
Molecules 28 00012 g001
Figure 2. Examples of biologically active compounds with axial chirality.
Figure 2. Examples of biologically active compounds with axial chirality.
Molecules 28 00012 g002
Scheme 1. Chiral Cu-amine catalytic systems employed in the construction of BINOL derivatives with axial chirality [39,40,41,42,43,44,45,46,47].
Scheme 1. Chiral Cu-amine catalytic systems employed in the construction of BINOL derivatives with axial chirality [39,40,41,42,43,44,45,46,47].
Molecules 28 00012 sch001
Scheme 2. Chiral V, Fe, and Ru catalytic systems employed in the construction of optically pure C2-symmetric BINOL derivatives [50,51,52,53,54,55,56,57,58,59].
Scheme 2. Chiral V, Fe, and Ru catalytic systems employed in the construction of optically pure C2-symmetric BINOL derivatives [50,51,52,53,54,55,56,57,58,59].
Molecules 28 00012 sch002
Scheme 3. General aspects of the mechanism for aerobic radical–anion coupling of 2-napthols in the presence of metallic chiral complex catalysts (Mn+). DG: coordination aids. Adapted from Brunel et al [57].
Scheme 3. General aspects of the mechanism for aerobic radical–anion coupling of 2-napthols in the presence of metallic chiral complex catalysts (Mn+). DG: coordination aids. Adapted from Brunel et al [57].
Molecules 28 00012 sch003
Scheme 4. Recent synthetic protocols for the construction of axially chiral BINOL derivatives.
Scheme 4. Recent synthetic protocols for the construction of axially chiral BINOL derivatives.
Molecules 28 00012 sch004
Scheme 5. Mechanistic proposal for the enantioselective aerobic coupling of 2-naphthols based on a ligand catalytic system containing a spirocyclic skeleton of pyrrolidine oxazoline/CuBr.
Scheme 5. Mechanistic proposal for the enantioselective aerobic coupling of 2-naphthols based on a ligand catalytic system containing a spirocyclic skeleton of pyrrolidine oxazoline/CuBr.
Molecules 28 00012 sch005
Scheme 6. Enantioselective coupling between 2-naphthols (3c) mediated by an iron/bisquinolyldiamine ligand complex.
Scheme 6. Enantioselective coupling between 2-naphthols (3c) mediated by an iron/bisquinolyldiamine ligand complex.
Molecules 28 00012 sch006
Scheme 7. (Aqua)ruthenium (salen) catalyzed enantioselective aerobic coupling between 2-naphthols for access to C1-symmetric BINOL derivatives.
Scheme 7. (Aqua)ruthenium (salen) catalyzed enantioselective aerobic coupling between 2-naphthols for access to C1-symmetric BINOL derivatives.
Molecules 28 00012 sch007
Scheme 8. Asymmetric oxidative coupling of 2-naphthols mediated by a macrocyclic Cu(II) complex.
Scheme 8. Asymmetric oxidative coupling of 2-naphthols mediated by a macrocyclic Cu(II) complex.
Molecules 28 00012 sch008
Scheme 9. Chiral diphosphine oxide-iron(II) complex catalyzed enantioselective aerobic coupling between 2-naphthols to access C1-symmetric BINOL derivatives.
Scheme 9. Chiral diphosphine oxide-iron(II) complex catalyzed enantioselective aerobic coupling between 2-naphthols to access C1-symmetric BINOL derivatives.
Molecules 28 00012 sch009
Scheme 10. Copper-catalyzed asymmetric oxidative coupling of 2-naphthols for the synthesis of 6,6′- disubstituted BINOLs.
Scheme 10. Copper-catalyzed asymmetric oxidative coupling of 2-naphthols for the synthesis of 6,6′- disubstituted BINOLs.
Molecules 28 00012 sch010
Scheme 11. Atroposelective synthesis of (R)- and (S)-BINOLs (1) via mono- and binuclear vanadium catalysts.
Scheme 11. Atroposelective synthesis of (R)- and (S)-BINOLs (1) via mono- and binuclear vanadium catalysts.
Molecules 28 00012 sch011
Scheme 12. Electrochemical synthesis of (S)-BINOL (1) using a TEMPO-modified graphite electrode.
Scheme 12. Electrochemical synthesis of (S)-BINOL (1) using a TEMPO-modified graphite electrode.
Molecules 28 00012 sch012
Scheme 13. Enantioselective Ni-promoted electrochemical synthesis of (R)-BINOL derivatives (1).
Scheme 13. Enantioselective Ni-promoted electrochemical synthesis of (R)-BINOL derivatives (1).
Molecules 28 00012 sch013
Scheme 14. Palladium-catalyzed kinetic resolution of 2,2’-dihydroxy-1,1’-biaryls using 6.
Scheme 14. Palladium-catalyzed kinetic resolution of 2,2’-dihydroxy-1,1’-biaryls using 6.
Molecules 28 00012 sch014
Scheme 15. (A): Enantioselective intramolecular multi-step transformation catalyzed by chiral guanidine 9. (B): synthesis of (R)-11 through optical resolution using chiral diamine (S,S)-12.
Scheme 15. (A): Enantioselective intramolecular multi-step transformation catalyzed by chiral guanidine 9. (B): synthesis of (R)-11 through optical resolution using chiral diamine (S,S)-12.
Molecules 28 00012 sch015
Scheme 16. Kinetic resolution of a BINOL derivative promoted by a DMAP-type (14) catalyst through O-acylation.
Scheme 16. Kinetic resolution of a BINOL derivative promoted by a DMAP-type (14) catalyst through O-acylation.
Molecules 28 00012 sch016
Scheme 17. Kinetic resolution of a BINOL derivative promoted by a NHC catalyst 18.
Scheme 17. Kinetic resolution of a BINOL derivative promoted by a NHC catalyst 18.
Molecules 28 00012 sch017
Scheme 18. Synthesis of binaphthyl derivatives through an intramolecular aldol condensation using catalyst 21.
Scheme 18. Synthesis of binaphthyl derivatives through an intramolecular aldol condensation using catalyst 21.
Molecules 28 00012 sch018
Scheme 19. Asymmetric synthesis of BINOL derivatives 25 via O-alkylation using catalyst 24.
Scheme 19. Asymmetric synthesis of BINOL derivatives 25 via O-alkylation using catalyst 24.
Molecules 28 00012 sch019
Scheme 20. Synthesis of (S)-29 via atroposelective aldol condensation using catalyst 27.
Scheme 20. Synthesis of (S)-29 via atroposelective aldol condensation using catalyst 27.
Molecules 28 00012 sch020
Scheme 21. Kinetic resolution of BINOL mediated by catalyst 32.
Scheme 21. Kinetic resolution of BINOL mediated by catalyst 32.
Molecules 28 00012 sch021
Scheme 22. Enantiospecific hydrolysis catalyzed by cholesterol esterase.
Scheme 22. Enantiospecific hydrolysis catalyzed by cholesterol esterase.
Molecules 28 00012 sch022
Scheme 23. Kinetic resolution by lipase (Candida antarctica).
Scheme 23. Kinetic resolution by lipase (Candida antarctica).
Molecules 28 00012 sch023
Scheme 24. Kinetic resolution with immobilized Pseudomonas sp. lipoprotein lipase.
Scheme 24. Kinetic resolution with immobilized Pseudomonas sp. lipoprotein lipase.
Molecules 28 00012 sch024
Scheme 25. Preparation of chiral quaternary ammonium bromide as a phase transfer catalyst from (S)-BINOL (1).
Scheme 25. Preparation of chiral quaternary ammonium bromide as a phase transfer catalyst from (S)-BINOL (1).
Molecules 28 00012 sch025
Scheme 26. Synthesis of (S)-3,3’,6,6’-tetrakis(trifluoromethyl)-BINOL (42) from (S)-BINOL (1).
Scheme 26. Synthesis of (S)-3,3’,6,6’-tetrakis(trifluoromethyl)-BINOL (42) from (S)-BINOL (1).
Molecules 28 00012 sch026
Scheme 27. Synthesis of BINOL derivatives containing phosphoric acid.
Scheme 27. Synthesis of BINOL derivatives containing phosphoric acid.
Molecules 28 00012 sch027
Table 1. Most used catalysts for axial chirality transfer.
Table 1. Most used catalysts for axial chirality transfer.
Molecules 28 00012 i001
EntryR1/R4R2/R3Ref.EntryR1/R4R2/R3Ref.
(R)-13,5-(3,5-(CF3)2C6H3)2C6H3Molecules 28 00012 i002[12](R)-19-anthrylMolecules 28 00012 i003[13]
(R)-13,5-(3,5-(CF3)2C6H3)2C6H3Molecules 28 00012 i004[12](R)-19-phenanthrylMolecules 28 00012 i005[14,15]
(R)-11-pyrenylMolecules 28 00012 i005[16](R)-12,4,6-(i-Pr)3C6H2Molecules 28 00012 i005[17]
(R)-19-anthrylMolecules 28 00012 i005[16](S)-1Molecules 28 00012 i006Molecules 28 00012 i007[18]
(S)-12,5-Me2-4-t-BuC6H2Molecules 28 00012 i005[13](S)-12,4,6-(CH3)3C6H2Molecules 28 00012 i005[19]
(S)-11-naphthylMolecules 28 00012 i005[20](S)-1Si(3-F-C6H4)3Molecules 28 00012 i005[21]
(S)-11-pyrenylMolecules 28 00012 i005[20,22,23](S)-13,5-(CF3)2C6H3Molecules 28 00012 i005[24]
(S)-1C6F5Molecules 28 00012 i005[25](S)-12-naphthylMolecules 28 00012 i005[26]
(S)-12,4,6-(i-Pr)3C6H2Molecules 28 00012 i005[27,28,29,30](S)-19-anthrylMolecules 28 00012 i005[23,31,32]
(R)-22-naphtylMolecules 28 00012 i005[21,33](S)-22,4,6-Me2C6H2Molecules 28 00012 i005[34]
(R)-22,4,6-(i-Pr)3C6H2Molecules 28 00012 i005[35](S)-22-naphtylMolecules 28 00012 i005[36]
(S)-29-anthrylMolecules 28 00012 i005[37](S)-21-naphthylMolecules 28 00012 i005[38]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

da Silva, E.M.; Vidal, H.D.A.; Januário, M.A.P.; Corrêa, A.G. Advances in the Asymmetric Synthesis of BINOL Derivatives. Molecules 2023, 28, 12. https://doi.org/10.3390/molecules28010012

AMA Style

da Silva EM, Vidal HDA, Januário MAP, Corrêa AG. Advances in the Asymmetric Synthesis of BINOL Derivatives. Molecules. 2023; 28(1):12. https://doi.org/10.3390/molecules28010012

Chicago/Turabian Style

da Silva, Everton Machado, Hérika Danielle Almeida Vidal, Marcelo Augusto Pereira Januário, and Arlene Gonçalves Corrêa. 2023. "Advances in the Asymmetric Synthesis of BINOL Derivatives" Molecules 28, no. 1: 12. https://doi.org/10.3390/molecules28010012

Article Metrics

Back to TopTop