Summary
A versatile method for atroposelective synthesis of chiral biaryl diamines and amino alcohols has been developed via para-amination of anilines and phenols with azodicarboxylates enabled by chiral phosphoric acid catalysis. Meanwhile, highly efficient kinetic resolution of the racemic biaryl anilines has also been realized through these reactions, giving selectivity factor up to 246. The gram-scale reaction and facile derivatizations of the chiral products well demonstrate the potential of these reactions in the development of novel chiral ligands and catalysts.
Subject Areas: Organic Chemistry, Organic Synthesis, Stereochemistry
Graphical Abstract

Highlights
-
•
Versatile methods for asymmetric synthesis of biaryl diamines and amino alcohols
-
•
Atroposelective para-aminations of biaryl anilines and phenols
-
•
Kinetic resolution of racemic biaryl anilines
-
•
Facile transformations of chiral products
Organic Chemistry; Organic Synthesis; Stereochemistry
Introduction
Biaryl compounds possessing axial chirality are ubiquitous among biologically active natural products and pharmaceuticals (Bringmann et al., 2011, Kozlowski et al., 2009) and have been extensively exploited as chiral ligands/catalysts in asymmetric catalysis (Brunel, 2005, Brunel, 2007, Chen et al., 2003, Kočovský et al., 2003, McCarthy and Guiry, 2001). To this end, their highly efficient and asymmetric catalytic synthesis has drawn increasing research interests, and various elegant methods have been developed in the last two decades (Bencivenni, 2015, Bonne and Rodriguez, 2018, Liao et al., 2019, Ma and Sibi, 2015, Nguyen, 2019, Renzi, 2017, Wang and Tan, 2018, Wencel-Delord et al., 2015, Zilate et al., 2018). However, in contrast to the numerous well-developed methods for asymmetric synthesis of BINOL-type biaryl diols (Chen et al., 2015, Egami et al., 2010, Guo et al., 2007, Jarvo et al., 2001, Jolliffe et al., 2017, Li et al., 2003, Luo et al., 2002, Ma et al., 2014, Moliterno et al., 2016, Moustafa et al., 2018, Narute et al., 2016, Wang et al., 2016, Xu et al., 2017), methods for enantioselective synthesis of other functionalized chiral biaryls are relatively limited. 1,1′-Binaphthyl-2,2′-diamine (BINAM), a representative chiral biaryl diamine, has been widely exploited in the development of chiral ligands and organocatalysts (Galzerano et al., 2009, Tan et al., 2011, Telfer and Kuroda, 2003, Uraguchi et al., 2009, Wang et al., 2005). However, only limited asymmetric catalytic methods (Brown et al., 1985, Chang et al., 2019) have been developed for its enantioselective synthesis, including asymmetric [3,3]-sigmatropic rearrangement (De et al., 2013, Li et al., 2013) and kinetic resolution (Cheng et al., 2014). 2-Amino-2′-hydroxy-1,1′-binaphthyl (NOBIN) (Smrcina et al., 1992, Smrcina et al., 1993), which is considered as the hybrid analogue of BINOL and BINAM, represents one type of privileged biaryl amino alcohol scaffold for constructing chiral ligands (Ding et al., 2005a, Ding et al., 2005b, Kočovský et al., 2003). However, methods for their asymmetric catalytic synthesis was also limited to kinetic resolutions (Lu et al., 2014, Shirakawa et al., 2013) and enantioselective direct arylation of 2-naphthylamines (Chen et al., 2017). Although the aforementioned elegant methods have provided access to enantioenriched biaryl diamines and amino alcohols, respectively, versatile methods for their asymmetric synthesis remain elusive. Recently, Tan and co-workers reported the asymmetric synthesis of BINAM- and NOBIN-type biaryls via enantioselective additions of 2-naphthols and 2-naphthylamines with 2-azonaphthalenes, which represented the first versatile protocol for their asymmetric synthesis, although two different catalytic systems were required (Qi et al., 2019).
Asymmetric Friedel-Crafts aminations of naphthols and naphthylamines with azodicarboxylates have been well employed in asymmetric synthesis of N-containing chiral scaffolds. For instance, Jørgensen group (Brandes et al., 2006a, Brandes et al., 2006b) and Zhang group (Bai et al., 2019) developed asymmetric construction of C-N axial chirality by chiral amine and phosphoric acid-catalyzed ortho-amination of 2-naphthols and 2-naphthyl amines, respectively (Scheme 1A). You group (Wang et al., 2015, Xia et al., 2019) and Luan group (Nan et al., 2015) reported asymmetric dearomatization of naphthols via direct aminations of naphthols with azodicarboxylates enabled by chiral Brønsted/Lewis acid catalysis, constructing N-containing chiral quaternary centers (Scheme 1B). Nevertheless, most of these methods are still limited to ortho-aminations of naphthols and naphthylamines; asymmetric reactions involving para-aminations of common anilines and phenols (Leblanc and Boudreault, 1995, Tang et al., 2017, Yadav et al., 2002, Zaltsgendler et al., 1993) are still elusive. Herein, we report a versatile protocol for atroposelective synthesis of biaryl diamines and amino alcohols via para-aminations of anilines and phenols (Diener et al., 2015, Gustafson et al., 2010, Miyaji et al., 2015, Miyaji et al., 2017, Mori et al., 2013a, Mori et al., 2013b) with azodicarboxylates via chiral phosphoric acid catalysis (Akiyama, 2007, Akiyama et al., 2004, Akiyama and Mori, 2015, Li and Song, 2018, Parmar et al., 2014, Terada, 2010, Uraguchi and Terada, 2004) (Scheme 1C).
Scheme 1.
Asymmetric Friedel-Crafts Amination with Azodicarboxylates
(A) construction of C-N axial chirality, (B) construction of N-containing chiral quaternary centers, and (C) atroposelective synthesis of biaryl diamines and amino alcohols.
Results and Discussion
Optimization of Reaction Conditions
Our study commenced with using biaryl aniline 1a as substrate and dibenzyl azodicarboxylate 2 as amination reagent under the catalysis of CPA catalysts (Table 1). Interestingly, in the presence of CPA catalyst A1 (10 mol %), the amination reaction between 1a and azodicarboxylate 2 (1.1 equiv.) in toluene (with 5 Å molecular sieves) proceeded smoothly at ambient temperature to afford the triazane 4a (Egger et al., 1983, Tang et al., 2017) as the major product (60% yield), whereas the desired para-amination product 3a was obtained only in 13% yield with 47% enantiomeric excess (ee) (entry 1). Next, a variety of BINOL-derived chiral phosphoric acid catalysts were examined (entries 2–7), and encouragingly the TCYP catalyst (cat A7) provided the desired product 3a in 80% yield with 98% ee, with the undesired N-amination product 4a and diamination product 5a isolated in <10% yield (entry 7). Next, a range of solvents were also investigated (entries 8–10), and CHCl3 turned out to be the optimal one, in which the desired product 3a was produced in 91% yield with 98% ee (entry 9). The role of the molecular sieves was also demonstrated; in the absence of 5-Å molecular sieves, the axially chiral biaryl 3a was obtained only in 66% yield (entry 11). The reduction of catalyst loading was also studied; however, decreasing the catalyst loading to 5 mol % at room temperature led to a diminished yield (entry 12). Interestingly, conducting this reaction with 5 mol % catalyst at 40°C gave product 3a in 87% yield with the same ee (entry 13). The axially chiral biaryl product 3a has high configurational stability, whose ee was retained after storing on bench for more than 2 months at ambient temperature and heating at 100°C in toluene for 36 h.
Table 1.
Optimization of the Reaction Conditions
![]() | ||||||
|---|---|---|---|---|---|---|
| Entrya | Cat. | Sol. | 4a Yieldb (%) | 5a Yieldb (%) | 3a Yieldb (%) | 3a eec (%) |
| 1 | A1 | Tol | 60 | 16 | 13 | 47 |
| 2 | A2 | Tol | 35 | 22 | 19 | 81 |
| 3 | A3 | Tol | 58 | 22 | 5 | 46 |
| 4 | A4 | Tol | 20 | 22 | 28 | 94 |
| 5 | A5 | Tol | 48 | 17 | 5 | 85 |
| 6 | A6 | Tol | 7 | 17 | 54 | 98 |
| 7 | A7 | Tol | 9 | 6 | 80 | 98 |
| 8 | A7 | DCM | 14 | 22 | 56 | 98 |
| 9 | A7 | CHCl3 | – | – | 91 | 98 |
| 10 | A7 | Et2O | 44 | 22 | 16 | 96 |
| 11d | A7 | CHCl3 | – | 17 | 66 | 97 |
| 12e | A7 | CHCl3 | 7 | 18 | 74 | 98 |
| 13e,f | A7 | CHCl3 | – | – | 87 | 98 |
Unless otherwise noted, reactions were performed with 1a (0.1 mmol), 2 (0.11 mmol), CPA catalyst (0.01 mmol), and 5 Å MS (30 mg) in solvents (0.5 mL) for 16 h at ambient temperature.
Yield was isolated yield.
Enantiomeric excess (ee) was determined by HPLC analysis on a chiral stationary phase.
Reaction was performed without 5 Å MS.
Reaction was performed with 5 mol % catalyst.
Reaction was performed at 40°C.
Substrate Scope
With the optimal conditions in hand, we next sought to explore the compatibility of substrate scope of this reaction (Scheme 2). A range of substituted 2-naphthylamine moieties could be well tolerated in the biaryl aniline substrates, affording the axially chiral amination products with high enantioselectivities (3b-3d). A series of substitutions at the ortho- and meta-positions of the aniline moieties in substrates was also compatible with the optimal conditions (3f-3h). It is worth mentioning that the direct amination of substrates 1g and 1h afforded products as inseparable diastereomer mixtures due to the presence of extra C-N axial chirality; therefore, these products were directly converted into –NH2-containing product 3g and 3h by catalytic hydrogenations. The absolute configurations of the axially chiral products 3 were assigned as (S) by analogy to product 3g, whose structure was unambiguously confirmed by X-ray crystallography (see Supplemental Information). The 2-naphthylamine scaffold in the substrates could also be switched to 3-substituted anilines (3i-3k), which also produced the biaryl amination products with high enantioselectivities under the standard conditions. Switching the N-protecting group from -Boc to –Cbz was also well tolerated with the optimal conditions, which produced product 3l with excellent enantioselectivity. However, using the N-protecting group-free biaryl aniline as substrates provided the triazane product as the major product.
Scheme 2.
Substrate Scope for Asymmetric Synthesis of Biaryl Diamines via Para-aminations of Anilines
Reactions were performed with 1 (0.1 mmol), 2 (0.11 mmol), (R)-cat A7 (0.005 mmol), and 5 Å MS (30 mg) in CHCl3 (0.5 mL) at 40°C overnight. Yield was isolated yield. Ee was determined by HPLC analysis on a chiral stationary phase.
aThe amination products were subjected into catalytic hydrogenation (1 atm) with Pd/C (10 mol %) as catalyst to afford 3g and 3h.
With the excellent performance of constructing chiral biaryl diamines via asymmetric para-amination reactions, we envisioned that these reactions could also be adopted in the kinetic resolution of racemic biaryl anilines. Thus, a variety of 2-substituted biaryl anilines possessing axial chirality were synthesized and their kinetic resolution via para-amination reactions with azodicarboxylate 2 (0.6 equiv.) was investigated (Scheme 3). Under the catalysis of (R)-TRIP catalyst (cat A6, 10 mol %) in DCM at room temperature, the kinetic resolutions of these substrates proceeded with high efficiencies to afford both recovered aniline substrates and para-amination products with high enantioselectivities (with s factor up to 246, 3m-3p). The absolute configurations of the axially chiral products and recovered starting materials were assigned by analogy to recovered 1m, whose structure was unambiguously confirmed by X-ray crystallography (see Supplemental Information).
Scheme 3.
Substrate Scope for Kinetic Resolution of Biaryl Anilines by Para-aminations
Reactions were performed with 1 (0.1 mmol), 2 (0.06 mmol), (R)-cat A6 (0.01 mmol), and 5 Å MS (30 mg) in CH2Cl2 (0.5 mL) at room temperature overnight. Yield was isolated yield. Ee was determined by HPLC analysis on a chiral stationary phase.
To achieve enantioselective synthesis of biaryl amino alcohols, the 2-naphthylamine moieties in the substrates were switched to 2-naphthol moieties. However, the amination reactions of the corresponding biaryl anilines 6′ provided only N-amination triazane products 7’ (Scheme 4). Interestingly, the desired biaryl amino alcohols were obtained while phenols 6 were employed as electron-rich arenes instead of anilines. Under the catalysis of (R)-C8-TRIP catalyst (cat A8) in CHCl3 at ambient temperature, the para-amination of biaryl phenol 6a with azodicarboxylate 2 afforded the biaryl amino alcohol 7a in 56% yield with 86% ee (for further details, see Table S1 in the Supplemental Information). Switching the protecting group from O-Me to O-MOM was compatible with the optimal conditions, providing the axially chiral amination product with comparable stereoselectivity (7b). The substrate scope for the 2-naphthol moieties in the substrates were explored under the standard conditions, which showed that a range of substituted 2-naphthol scaffolds (with various substitutions at the 4-, 6-, and 7-positions) could be accommodated, affording biaryl amino alcohols in good yields and high enantioselectivities (7c–7g). The absolute configurations of these biaryl amino alcohols 7 were assigned by analogy to product 7c, whose structure was also confirmed by X-ray crystallography (Supplemental Information). It is worth mentioning that biaryl phenol substrate without O-Me protecting group afforded the para-amination product with diminished enantioselectivity under the optimal conditions.
Scheme 4.
Substrate Scope for Asymmetric Synthesis of Biaryl Amino Alcohols via Para-amination of Phenols
Reactions were performed with 6 (0.1 mmol), 2 (0.3 mmol), (R)-cat A8 (0.01 mmol), and 5 Å MS (30 mg) in CHCl3 (0.5 mL) at ambient temperature for 36 h. Yield was isolated yield. Ees was determined by HPLC analysis on a chiral stationary phase.
aReaction was performed with 2 (0.15 mmol).
Mechanistic Discussion
To gain more insight into the reaction mechanism, several control experiments were performed (Scheme 5). The amination reaction of N-Me biaryl aniline substrate 1q proceeded smoothly under the standard conditions to give the desired para-amination product 3q in 56% yield with 92% ee. However, applying the same conditions on N,N-dimethyl aniline substrate 1r provided only the para-amination product 3r in 25% yield with 5% ee (with the N-amination product as the major by-product), which indicated that the potential hydrogen bonding between CPA catalyst and the aniline N-H group played a key role in controlling both chemoselectivity and stereoselectivity in this reaction (Scheme 5A). Interestingly, subjection of the N-amination triazane product 4a into the optimal conditions without adding azodicarboxylate 2 also gave the para-amination product 3a in 58% yield with 98% ee after 16 h, with the aniline substrate 1a isolated in 28% yield, which suggested the reversible nature of the triazane formation step (Scheme 5B). Based on the above-mentioned experimental study and previous work (Bai et al., 2019, Drouet et al., 2011, Dumoulin et al., 2015), a plausible reaction mechanism is proposed, in which bifunctional activation (Parmar et al., 2014, Simón and Goodman, 2008, Yamanaka et al., 2007) of both the aniline substrate and azodicarboxylate via dual hydrogen-bonding interaction with the CPA catalyst is postulated (Scheme 5C). Under the catalysis of CPA catalyst, there are two alternative reaction pathways between aniline substrates and azodicarboxylates: (1) direct nucleophilic addition of the –NH2 group to the azodicarboxylate facilitated the generation of the triazane products (path a), which is also reversible under these conditions; and (2) the para-selective amination of aniline substrates would give the dearomatized addition product INT A, possessing a chiral center (path b). On subsequent aromatization, INT A underwent the central-to-axial chirality transfer (Qi et al., 2017, Raut et al., 2017) to provide the axial biaryl diamine products.
Scheme 5.
Preliminary Mechanistic Study and Proposed Mechanism
Transformations of Products
To evaluate the practicability of these reactions, a gram-scale amination reaction of 1a was performed, which provided the axially chiral biaryl 3a in 70% yield with 99% ee, with reduced catalyst loading (2 mol %, Scheme 6A). The derivatizations of the chiral products were also studied to prove the value of these reactions. By means of the Sandmeyer reaction, the directing –NH2 group was transformed into an iodide group via diazotization of 3a with NaNO2 followed by treatment with NaI to afford 8a, which could be further employed in Suzuki coupling with phenylboronic acid to give product 9a in 81% yield (Scheme 6B). Notably, the ee of chiral biaryl product was retained through these steps of transformations, including the Suzuki coupling step (105°C, overnight), again demonstrating the high configurational stability of these atropisomeric products. The catalytic hydrogenation of 7a using Pd/C as catalyst facilely reduced the substituted hydrazine moiety to give the biaryl product 10a in 90% yield (Scheme 6C). A two-step procedure of catalytic hydrogenation followed by deprotection of the N-Boc group converted chiral product 3n into biaryl diamine 11n in 82% yield, without erosion of the enantioselectivity (Scheme 6D). Finally, a primary-amine/thiourea bifunctional catalyst 13a was straightforwardly synthesized from the chiral product 3a within 4 steps, with complete retention of the enantiomeric purity (Scheme 6E). The application of this bifunctional catalyst was preliminarily demonstrated in an asymmetric Michael reaction of 3-methyl oxindole 14 with cinnamaldehyde 15 (Galzerano et al., 2009), which readily provided the product 16 (after reduction) in 53% yield with 6:1 d.r. and 73% ee without optimization (Scheme 6F).
Scheme 6.
Gram-Scale Synthesis of 3a and Derivatizations of the Chiral Products
Conclusion
We have disclosed a versatile method for asymmetric synthesis of biaryl diamines and amino alcohols, which was realized through chiral phosphoric acid catalyzed enantioselective para-aminations of biaryl anilines and phenols with azodicarboxylates. These reactions are also well employed in the highly efficient kinetic resolution of racemic biaryl anilines, which give s factor up to 246. Preliminary mechanistic studies were performed to elucidate the reaction mechanism, in which a dual hydrogen-bonding activation mode was proposed in the key chirality induction step. The facile transformations of chiral products into atropisomeric biaryl diamine and amino alcohol derivatives with novel and diversified scaffolds well demonstrate the value of these reactions, especially in the field of developments of novel chiral catalysts and ligands.
Limitations of the Study
The synthesis of the substrates usually needs multiple steps. Different directing groups are required in the synthesis of biaryl diamines and biaryl amino alcohols.
There are also some limitations of the substrate scope (Scheme 7): (1) electron-donating groups were required at the 2-position of the naphthyl moiety; substrates with alkyl groups at this position barely provided the para-amination products (S1a and S1b); (2) kinetic resolution of racemic biaryl phenol substrate did not provide good kinetic resolution performance (S1c); (3) substitutions at the 2-position of the phenol moiety and 3-position of the naphthol moiety were not compatible in the asymmetric para-amination reactions of biaryl phenols (S1d and S1e).
Scheme 7.
Incompatible Substrates
Methods
All methods can be found in the accompanying Transparent Methods supplemental file.
Acknowledgments
We gratefully acknowledge NSFC (Grant No. 21702138), “Thousand Talents Plan” Youth Program, and ShanghaiTech University start-up funding for financial support. We thank the Analytical Instrumentation Centre of ShanghaiTech for facilities and services of characterization of compounds.
Author Contributions
D.W., W.L., and M.T. performed the experiments. N.Y. performed the crystallographic studies. X.Y. conceived the concept, directed the project, and wrote the paper.
Declaration of Interests
The authors declare no competing interests.
Published: December 20, 2019
Footnotes
Supplemental Information can be found online at https://doi.org/10.1016/j.isci.2019.11.024.
Data and Code Availability
The X-ray crystallographic coordinates for structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre (CCDC) under accession number CCDC: 1923360 (3g), 1938295 (7c), and 1923362 ((S)-1m). These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
Supplemental Information
References
- Akiyama T. Stronger Brønsted acids. Chem. Rev. 2007;107:5744–5758. doi: 10.1021/cr068374j. [DOI] [PubMed] [Google Scholar]
- Akiyama T., Itoh J., Yokota K., Fuchibe K. Enantioselective Mannich-type reaction catalyzed by a chiral Brønsted acid. Angew. Chem. Int. Ed. 2004;43:1566–1568. doi: 10.1002/anie.200353240. [DOI] [PubMed] [Google Scholar]
- Akiyama T., Mori K. Stronger Brønsted acids: recent progress. Chem. Rev. 2015;115:9277–9306. doi: 10.1021/acs.chemrev.5b00041. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1038/s41467-019-10858-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bencivenni G. Organocatalytic strategies for the synthesis of axially chiral compounds. Synlett. 2015;26:1915–1922. [Google Scholar]
- Bonne D., Rodriguez J. A bird's eye view of atropisomers featuring a five-membered ring. Eur. J. Org. Chem. 2018;2018:2417–2431. [Google Scholar]
- Brandes S., Bella M., Kjærsgaard A., Jørgensen K.A. Chirally aminated 2-naphthols—organocatalytic synthesis of non-biaryl atropisomers by asymmetric Friedel–Crafts amination. Angew. Chem. Int. Ed. 2006;45:1147–1151. doi: 10.1002/anie.200503042. [DOI] [PubMed] [Google Scholar]
- Brandes S., Niess B., Bella M., Prieto A., Overgaard J., Jørgensen K.A. Non-biaryl atropisomers in organocatalysis. Chem. Eur. J. 2006;12:6039–6052. doi: 10.1002/chem.200600495. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/cr100155e. [DOI] [PubMed] [Google Scholar]
- Brown K.J., Berry M.S., Murdoch J.R. Synthesis of optically active 2,2'-dihalo-1,1'-binaphthyls via stable diazonium salts. J. Org. Chem. 1985;50:4345–4349. [Google Scholar]
- Brunel J.M. BINOL: A versatile chiral reagent. Chem. Rev. 2005;105:857–898. doi: 10.1021/cr040079g. [DOI] [PubMed] [Google Scholar]
- Brunel J.M. Update 1 of: BINOL: A versatile chiral reagent. Chem. Rev. 2007;107:PR1–PR45. doi: 10.1021/cr040079g. [DOI] [PubMed] [Google Scholar]
- Chang X., Zhang Q., Guo C. Switchable smiles rearrangement for enantioselective O-aryl amination. Org. Lett. 2019;21:4915–4918. doi: 10.1021/acs.orglett.9b01848. [DOI] [PubMed] [Google Scholar]
- Chen Y.-H., Cheng D.-J., Zhang J., Wang Y., Liu X.-Y., Tan B. Atroposelective synthesis of axially chiral biaryldiols via organocatalytic arylation of 2-naphthols. J. Am. Chem. Soc. 2015;137:15062–15065. doi: 10.1021/jacs.5b10152. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1002/anie.201710537. [DOI] [PubMed] [Google Scholar]
- Chen Y., Yekta S., Yudin A.K. Modified BINOL ligands in asymmetric catalysis. Chem. Rev. 2003;103:3155–3212. doi: 10.1021/cr020025b. [DOI] [PubMed] [Google Scholar]
- Cheng D.-J., Yan L., Tian S.-K., Wu M.-Y., Wang L.-X., Fan Z.-L., Zheng S.-C., Liu X.-Y., Tan B. Highly enantioselective kinetic resolution of axially chiral BINAM derivatives catalyzed by a Brønsted acid. Angew. Chem. Int. Ed. 2014;53:3684–3687. doi: 10.1002/anie.201310562. [DOI] [PubMed] [Google Scholar]
- De C.K., Pesciaioli F., List B. Catalytic asymmetric benzidine rearrangement. Angew. Chem. Int. Ed. 2013;52:9293–9295. doi: 10.1002/anie.201304039. [DOI] [PubMed] [Google Scholar]
- Diener M.E., Metrano A.J., Kusano S., Miller S.J. Enantioselective synthesis of 3-arylquinazolin-4(3H)-ones via peptide-catalyzed atroposelective bromination. J. Am. Chem. Soc. 2015;137:12369–12377. doi: 10.1021/jacs.5b07726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ding K., Guo H., Li X., Yuan Y., Wang Y. Synthesis of NOBIN derivatives for asymmetric catalysis. Top. Catal. 2005;35:105–116. [Google Scholar]
- Ding K., Li X., Ji B., Guo H., Kitamura M. Ten years of research on NOBIN chemistry. Curr. Org. Syn. 2005;2:499–545. [Google Scholar]
- Drouet F., Lalli C., Liu H., Masson G., Zhu J. Chiral calcium organophosphate-catalyzed enantioselective electrophilic amination of enamides. Org. Lett. 2011;13:94–97. doi: 10.1021/ol102625s. [DOI] [PubMed] [Google Scholar]
- Dumoulin A., Lalli C., Retailleau P., Masson G. Catalytic, highly enantioselective, direct amination of enecarbamates. Chem. Commun. (Camb.) 2015;51:5383–5386. doi: 10.1039/c4cc08052a. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/ja105442m. [DOI] [PubMed] [Google Scholar]
- Egger N., Hoesch L., Dreiding A.S. Azimine. VII. Herstellung durch Oxydation von Triazanen. Helv. Chim. Acta. 1983;66:1599–1607. [Google Scholar]
- Galzerano P., Bencivenni G., Pesciaioli F., Mazzanti A., Giannichi B., Sambri L., Bartoli G., Melchiorre P. Asymmetric iminium ion catalysis with a novel bifunctional primary amine thiourea: controlling adjacent quaternary and tertiary stereocenters. Chem. Eur. J. 2009;15:7846–7849. doi: 10.1002/chem.200802466. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/ja074322f. [DOI] [PubMed] [Google Scholar]
- Gustafson J.L., Lim D., Miller S.J. Dynamic kinetic resolution of biaryl atropisomers via peptide-catalyzed asymmetric bromination. Science. 2010;328:1251. doi: 10.1126/science.1188403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jarvo E.R., Evans C.A., Copeland G.T., Miller S.J. Fluorescence-based screening of asymmetric acylation catalysts through parallel enantiomer analysis. Identification of a catalyst for tertiary alcohol resolution. J. Org. Chem. 2001;66:5522–5527. doi: 10.1021/jo015803z. [DOI] [PubMed] [Google Scholar]
- Jolliffe J.D., Armstrong R.J., Smith M.D. Catalytic enantioselective synthesis of atropisomeric biaryls by a cation-directed O-alkylation. Nat. Chem. 2017;9:558. doi: 10.1038/nchem.2710. [DOI] [PubMed] [Google Scholar]
- Kočovský P., Vyskočil Š., Smrčina M. Non-symmetrically substituted 1,1‘-binaphthyls in enantioselective catalysis. Chem. Rev. 2003;103:3213–3246. doi: 10.1021/cr9900230. [DOI] [PubMed] [Google Scholar]
- Kozlowski M.C., Morgan B.J., Linton E.C. Total synthesis of chiral biaryl natural products by asymmetric biaryl coupling. Chem. Soc. Rev. 2009;38:3193–3207. doi: 10.1039/b821092f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leblanc Y., Boudreault N. Para-directed amination of electron-rich arenes with bis(2,2,2-trichloroethyl) azodicarboxylate. J. Org. Chem. 1995;60:4268–4271. [Google Scholar]
- 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. doi: 10.1021/ja401709k. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/jo0340206. [DOI] [PubMed] [Google Scholar]
- Li X., Song Q. Recent advances in asymmetric reactions catalyzed by chiral phosphoric acids. Chin. Chem. Lett. 2018;29:1181–1192. [Google Scholar]
- 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. (Camb.) 2019;55:8514–8523. doi: 10.1039/c9cc03967h. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1002/anie.201406192. [DOI] [PubMed] [Google Scholar]
- 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. (Camb.) 2002:914–915. doi: 10.1039/b201351g. [DOI] [PubMed] [Google Scholar]
- Ma G., Deng J., Sibi M.P. Fluxionally chiral DMAP catalysts: kinetic resolution of axially chiral biaryl compounds. Angew. Chem. Int. Ed. 2014;53:11818–11821. doi: 10.1002/anie.201406684. [DOI] [PubMed] [Google Scholar]
- Ma G., Sibi M.P. Catalytic kinetic resolution of biaryl compounds. Chemistry. 2015;21:11644–11657. doi: 10.1002/chem.201500869. [DOI] [PubMed] [Google Scholar]
- McCarthy M., Guiry P.J. Axially chiral bidentate ligands in asymmetric catalysis. Tetrahedron. 2001;57:3809–3844. [Google Scholar]
- Miyaji R., Asano K., Matsubara S. Bifunctional organocatalysts for the enantioselective synthesis of axially chiral isoquinoline N-oxides. J. Am. Chem. Soc. 2015;137:6766–6769. doi: 10.1021/jacs.5b04151. [DOI] [PubMed] [Google Scholar]
- Miyaji R., Asano K., Matsubara S. Induction of axial chirality in 8-arylquinolines through halogenation reactions using bifunctional organocatalysts. Chem. Eur. J. 2017;23:9996–10000. doi: 10.1002/chem.201701707. [DOI] [PubMed] [Google Scholar]
- Moliterno M., Cari R., Puglisi A., Antenucci A., Sperandio C., Moretti E., Di Sabato A., Salvio R., Bella M. Quinine-catalyzed asymmetric synthesis of 2,2′-binaphthol-type biaryls under mild reaction conditions. Angew. Chem. Int. Ed. 2016;55:6525–6529. doi: 10.1002/anie.201601660. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/ja311902f. [DOI] [PubMed] [Google Scholar]
- Mori K., Ichikawa Y., Kobayashi M., Shibata Y., Yamanaka M., Akiyama T. Prediction of suitable catalyst by 1H NMR: asymmetric synthesis of multisubstituted biaryls by chiral phosphoric acid catalyzed asymmetric bromination. Chem. Sci. 2013;4:4235–4239. [Google Scholar]
- 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. doi: 10.1002/anie.201804161. [DOI] [PubMed] [Google Scholar]
- Nan J., Liu J., Zheng H., Zuo Z., Hou L., Hu H., Wang Y., Luan X. Direct asymmetric dearomatization of 2-naphthols by scandium-catalyzed electrophilic amination. Angew. Chem. Int. Ed. 2015;54:2356–2360. doi: 10.1002/anie.201409565. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1021/jacs.6b11198. [DOI] [PubMed] [Google Scholar]
- Nguyen T.T. Traceless point-to-axial chirality exchange in the atropselective synthesis of biaryls/heterobiaryls. Org. Bio. Chem. 2019;17:6952–6963. doi: 10.1039/c9ob01304k. [DOI] [PubMed] [Google Scholar]
- Parmar D., Sugiono E., Raja S., Rueping M. Complete field guide to asymmetric BINOL-phosphate derived Brønsted acid and metal catalysis: history and classification by mode of activation; Brønsted acidity, hydrogen bonding, ion pairing, and Metal Phosphates. Chem. Rev. 2014;114:9047–9153. doi: 10.1021/cr5001496. [DOI] [PubMed] [Google Scholar]
- Qi L.-W., Li S., Xiang S.-H., Wang J., Tan B. Asymmetric construction of atropisomeric biaryls via a redox neutral cross-coupling strategy. Nat. Catal. 2019;2:314–323. [Google Scholar]
- Qi L.-W., Mao J.-H., Zhang J., Tan B. Organocatalytic asymmetric arylation of indoles enabled by azo groups. Nat.Chem. 2017;10:58. doi: 10.1038/nchem.2866. [DOI] [PubMed] [Google Scholar]
- Raut V.S., Jean M., Vanthuyne N., Roussel C., Constantieux T., Bressy C., Bugaut X., Bonne D., Rodriguez J. Enantioselective syntheses of furan atropisomers by an oxidative central-to-axial chirality conversion strategy. J. Am. Chem. Soc. 2017;139:2140–2143. doi: 10.1021/jacs.6b11079. [DOI] [PubMed] [Google Scholar]
- Renzi P. Organocatalytic synthesis of axially chiral atropisomers. Org. Bio. Chem. 2017;15:4506–4516. doi: 10.1039/c7ob00908a. [DOI] [PubMed] [Google Scholar]
- Shirakawa S., Wu X., Maruoka K. Kinetic resolution of axially chiral 2-amino-1,1′-biaryls by phase-transfer-catalyzed N-allylation. Angew. Chem. Int. Ed. 2013;52:14200–14203. doi: 10.1002/anie.201308237. [DOI] [PubMed] [Google Scholar]
- Simón L., Goodman J.M. Theoretical study of the mechanism of Hantzsch ester hydrogenation of imines catalyzed by chiral BINOL-phosphoric acids. J. Am. Chem. Soc. 2008;130:8741–8747. doi: 10.1021/ja800793t. [DOI] [PubMed] [Google Scholar]
- Smrcina M., Lorenc M., Hanus V., Sedmera P., Kocovsky P. Synthesis of enantiomerically pure 2,2'-dihydroxy-1,1'-binaphthyl, 2,2'-diamino-1,1'-binaphthyl, and 2-amino-2'-hydroxy-1,1'-binaphthyl. Comparison of processes operating as diastereoselective crystallization and as second order asymmetric transformation. J. Org. Chem. 1992;57:1917–1920. [Google Scholar]
- Smrcina M., Polakova J., Vyskocil S., Kocovsky P. Synthesis of enantiomerically pure binaphthyl derivatives. Mechanism of the enantioselective, oxidative coupling of naphthols and designing a catalytic cycle. J. Org. Chem. 1993;58:4534–4538. [Google Scholar]
- Tan B., Candeias N.R., Barbas Iii C.F. Construction of bispirooxindoles containing three quaternary stereocentres in a cascade using a single multifunctional organocatalyst. Nat. Chem. 2011;3:473. doi: 10.1038/nchem.1039. [DOI] [PubMed] [Google Scholar]
- Tang R.-J., Milcent T., Crousse B. Hexafluoro-2-propanol promotes para-selective C–H amination of free anilines with azodicarboxylates. Eur. J. Org. Chem. 2017;2017:4753–4757. [Google Scholar]
- Telfer S.G., Kuroda R. 1,1′-Binaphthyl-2,2′-diol and 2,2′-diamino-1,1′-binaphthyl: versatile frameworks for chiral ligands in coordination and metallosupramolecular chemistry. Coord. Chem. Rev. 2003;242:33–46. [Google Scholar]
- Terada M. Chiral phosphoric acids as versatile catalysts for enantioselective carbon-carbon bond forming reactions. Synthesis. 2010;2010:1929–1982. [Google Scholar]
- Uraguchi D., Nakashima D., Ooi T. Chiral arylaminophosphonium barfates as a new class of charged Brønsted acid for the enantioselective activation of nonionic Lewis bases. J. Am. Chem. Soc. 2009;131:7242–7243. doi: 10.1021/ja903271t. [DOI] [PubMed] [Google Scholar]
- Uraguchi D., Terada M. Chiral BRØNSTED acid-catalyzed direct Mannich reactions via electrophilic activation. J. Am. Chem. Soc. 2004;126:5356–5357. doi: 10.1021/ja0491533. [DOI] [PubMed] [Google Scholar]
- Wang J.-Z., Zhou J., Xu C., Sun H., Kürti L., Xu Q.-L. Symmetry in cascade chirality-transfer processes: a catalytic atroposelective direct arylation approach to BINOL derivatives. J. Am. Chem. Soc. 2016;138:5202–5205. doi: 10.1021/jacs.6b01458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J., Li H., Duan W., Zu L., Wang W. Organocatalytic asymmetric Michael addition of 2,4-pentandione to nitroolefins. Org. Lett. 2005;7:4713–4716. doi: 10.1021/ol0519137. [DOI] [PubMed] [Google Scholar]
- Wang S.G., Yin Q., Zhuo C.X., You S.L. Asymmetric dearomatization of β-naphthols through an amination reaction catalyzed by a chiral phosphoric acid. Angew. Chem. Int. Ed. 2015;54:647–650. doi: 10.1002/anie.201409756. [DOI] [PubMed] [Google Scholar]
- Wang Y.-B., Tan B. Construction of axially chiral compounds via asymmetric organocatalysis. Acc. Chem. Res. 2018;51:534–547. doi: 10.1021/acs.accounts.7b00602. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1039/c5cs00012b. [DOI] [PubMed] [Google Scholar]
- Xia Z.-L., Zheng C., Xu R.-Q., You S.-L. Chiral phosphoric acid catalyzed aminative dearomatization of α-naphthols/Michael addition sequence. Nat. Commun. 2019;10:3150. doi: 10.1038/s41467-019-11109-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu C., Zheng H., Hu B., Liu X., Lin L., Feng X. Chiral cobalt(ii) complex catalyzed Friedel–Crafts aromatization for the synthesis of axially chiral biaryldiols. Chem. Commun. (Camb.) 2017;53:9741–9744. doi: 10.1039/c7cc05266a. [DOI] [PubMed] [Google Scholar]
- Yadav J.S., Reddy B.V.S., Veerendhar G., Rao R.S., Nagaiah K. Sc(OTf)3 catalyzed electrophilic amination of arenes: an expeditious synthesis of aryl hydrazides. Chem. Lett. 2002;31:318–319. [Google Scholar]
- Yamanaka M., Itoh J., Fuchibe K., Akiyama T. Chiral Brønsted acid catalyzed enantioselective Mannich-type reaction. J. Am. Chem. Soc. 2007;129:6756–6764. doi: 10.1021/ja0684803. [DOI] [PubMed] [Google Scholar]
- Zaltsgendler I., Leblanc Y., Bernstein M.A. Synthesis of aromatic amines from electron-rich arenes and bis(2,2,2-trichloroethyl) azodicarboxylate. Tetrahedron Lett. 1993;34:2441–2444. [Google Scholar]
- Zilate B., Castrogiovanni A., Sparr C. Catalyst-controlled stereoselective synthesis of atropisomers. ACS Catal. 2018;8:2981–2988. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The X-ray crystallographic coordinates for structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre (CCDC) under accession number CCDC: 1923360 (3g), 1938295 (7c), and 1923362 ((S)-1m). These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.








