You are on page 1of 10

Articles

https://doi.org/10.1038/s41929-019-0247-1

Asymmetric construction of atropisomeric biaryls


via a redox neutral cross-coupling strategy
Liang-Wen Qi1,3, Shaoyu Li   1,2,3, Shao-Hua Xiang   1,2, Jun (Joelle) Wang   1* and Bin Tan   1*

Atropisomerically enriched biaryl frameworks are ubiquitous in many fields of chemistry. Enantioselective aryl–aryl cross-
coupling provides the most straightforward entry to atropisomeric biaryls, with remarkable application potential in the field
of chemical science. However, their development is hindered due to the lack of convenient and pragmatic protocols. Here, we
report a method for the asymmetric synthesis of a myriad of 2-amino-2′-hydroxy-1,1′-binaphthyl (NOBIN) and 1,1’-binaph-
thyl-2,2’-diamine (BINAM) derivatives in excellent yields and enantioselectivities via a redox-neutral cross-coupling protocol.
Two complementary systems were devised employing a chiral phosphoric acid–salt complex or Ni(OTf)2/chiral bis(oxazoline)
ligand catalytic system for accessing atropisomeric NOBIN and BINAM derivatives, respectively. This work provides an alterna-
tive avenue to enantioenriched biaryls, and provides the capability to explore the synthetic and catalytic potentials of NOBIN-
and BINAM-based frameworks.

A
tropisomerically enriched biaryl frameworks are abun- To transcend these synthetic constraints, inverting the electronics
dantly featured in a wide range of natural products1, phar- of an intrinsically nucleophilic aryl component to electrophilic for
maceuticals2 and bioactive molecules3. They also constitute chiral aryl C–H/C–H cross-coupling represents one ideal solution.
the enantiopure building blocks for axially chiral hosts in the reso- The interaction of a chiral electrophilic catalyst with a conjugated
lution of racemic amino acids via host–guest chemistry4 as well as functional group tethered to an arene was envisioned to further
stereogenic fluorescent sensors for highly enantioselective and sen- enhance the electrophilicity of the aromatic ring and, at the same
sitive recognition of chiral amino alcohols, α-hydroxycarboxylic time, impose steric hindrance to restrain the plausible 1,2-addition
acids and amino acid derivatives5. Furthermore, material scien- (Fig. 1d). However, the realization of this redox-neutral aryl–aryl
tists exploit axially chiral frameworks as stationary phases for cross-coupling poses a formidable challenge, given that control of
enantioselective separation6, dopants in liquid-crystalline materi- the axial chirality often requires mild reaction conditions. In this
als7, chiroptical molecular switches8, microporous soluble poly- regard, our recent research endeavour showed that the azo group is
mers9 and interlocked nanotubes10. They are, nevertheless, most an adequate activating and directing group for aryl C–H functional-
prominently prized as chiral ligands or catalysts in asymmetric ization29. Moreover, the azo functionality could deliver the nitrogen
synthesis11–18 (Fig. 1a). functionality for the high-value biaryls 2-amino-2′-hydroxy-1,1′-
Their distinct structural features, together with the significance binaphthyl (NOBIN) and 1,1’-binaphthyl-2,2’-diamine (BINAM).
of these biaryl core scaffolds, have galvanized the development of Aside from the abovementioned oxidative aryl C–H/C–H cross-
highly efficient preparation methods. To this end, atroposelective coupling via metal-induced radical formation, the synthesis of
aryl–aryl cross-coupling is the most efficient and straightforward enantioenriched NOBIN from BINOL still requires a lengthy syn-
route, as is evident from documentation of the diverse range of thetic route30. On the other hand, with the exception of the works
transition-metal-catalysed protocols (Fig. 1b). These include con- by List31 and Kürti32 that afford enantiopure BINAMs via exquisite
ventional transition-metal-catalysed cross-coupling reactions and organocatalytic [3,3]-diaza Cope rearrangement from mostly sym-
aryl C–H activation19–26 and oxidative C–H/C–H cross-couplings metrical 2,2′-hydrazonaphthalene (Fig. 1c), a versatile synthetic
enabled by a metal complex via a radical course27,28. Although a protocol with wider substrate range remains to be demonstrated.
direct C–H functionalization strategy negates the need to prefunc- In line with our research goal to devise concise and pragmatic
tionalize both coupling partners, it was rarely fruitful for enanti- synthetic methods for high-value axial chiral frameworks, we there-
oselective construction of hindered biaryls before Colobert’s work26 fore embarked on the exploration of azo-tethered aryl substrates
due to the discord between the temperature tolerance of the rota- for the asymmetric construction of NOBIN and BINAM deriva-
tional axis and the high temperature required for C–H activation. tives in a redox-neutral cross-coupling. Driven by recent prog-
On the other hand, the oxidative aryl C–H/C–H cross-coupling ress33–38 and our understanding of asymmetric biaryl synthesis39,
that proceeds via metal-induced radical formation suffers from we here detail a highly efficient chiral phosphoric acid (CPA)–salt
poor chemoselectivity and limited substrate scope for a stringent complex or Ni(OTf)2/chiral bis(oxazoline) (BOX) ligand-catalysed
redox potential difference between two coupling partners27,28. The redox-neutral cross-coupling of 2-naphthols or 2-naphthylamines
pursuit of a universal and robust catalytic approach to obtain opti- with azonaphthalene derivatives, providing a wide range of opti-
cally active biaryls by designing and exploiting coupling technology cally active NOBIN and BINAM derivatives with excellent yields
is therefore of great synthetic significance and would open the gate and enantiocontrol. Moreover, the valuable NOBINs and BINAMs
for the development of enantioselective catalytic transformations. were afforded via facile cleavage of the N–N bond with high

1
Department of Chemistry and Shenzhen Grubbs Institute, Southern University of Science and Technology, Shenzhen, China. 2Academy for Advanced
Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen, China. 3These authors contributed equally: Liang-Wen Qi and Shaoyu Li.
*e-mail: wang.j@sustc.edu.cn; tanb@sustc.edu.cn

Nature Catalysis | www.nature.com/natcatal


Articles NATurE CATAlysIs

a OR
Cl HN NH2·HCl O CO2H
O O

Cl NH
HO OH O
i
Pr O O O
H H H O O
OH O N N N
N H H
CHO OH H N N N N CO2Bn O O
HN H H O
HO O O O O
OH O O O
OH CHO HO2C
HO NH2 O CO2H
OH
i
Pr (–)-Gossypol OH Vancomycin

Natural product Pharmaceutical Bioactive molecule Host–guest chemistry

Ph Ph

NH OH O CH3 OMe
O N
O O N OR
OH N O
OH O O N O
RO N
H O
OH Polymer N
N CH3 OMe
R = C8H17, C10H21, C12H25, C14H29
Ph Ph
Chiral stationary phase
Fluorescent sensor Liquid-crystalline material Chiroptical switch

O O O O
O O OH PPh2 NH2 NH2
N
N H H N OH PPh2 OH NH2
N O O
O O O
O
n BINOL BINAP NOBIN BINAM

Microporous soluble polymer Interlocked nanotube Chiral ligands and catalysts


N

b c

Y R1 R1
[M] R1 NH Chiral phosphoric acid catalyst NH2
R2
+
R1 Chiral ligand R2 NH NH2
[3,3]-Sigmatropic rearrangement
X R2 R2

d
Key point:

2 Nucleophilic
4 H2N
NHR A 1 OH
NH2 B OH H
H 3 H Aryl umpolung strategy
st

H
taly

NH2 NH2
Ca

Redox-neutral Redox-neutral R N
N
cross-coupling cross-coupling Electrophilic
BINAM NOBIN

Fig. 1 | Axially chiral biaryl molecules and reaction development. a, Representative applications of optically active biaryls. b, Metal-catalysed asymmetric
cross-coupling for the construction of optically active biaryls. c, Asymmetric organocatalytic benzidine rearrangement by List and Kürti. d, Our designed
umpolung strategy for asymmetric cross-coupling of aryls.

enantioretention. This work is expected to provide an avenue for desired product 3a was formed in less than 20% yield for most exam-
rapid construction of NOBIN- and BINAM-based frameworks with ined CPAs when the reactions were conducted with 2-naphthol 2a
synthetic and catalytic potentials. in CH2Cl2 at room temperature. Furthermore, the enantioselectivity
was no higher than 47% e.e. with CPA1, suggesting CPA alone was
Results not sufficiently competent for this transformation (Table 1, entry 1).
Reaction development. First reported by the groups of Akiyama Inclusion of Lewis acids led to reduced enantioselectivity due to the
and Terada in 200440,41, uncovering the utility of CPAs has long been strong background reactions. Interestingly, reversed stereoselectiv-
part of our ongoing research efforts. Considering the adequacy of ity was observed when ScCl3, YbCl3 and particularly ZnCl2 were
these Brønsted acid catalysts in azo group activation, the proposed used, hinting at a CPA–Lewis acid interaction (Table 1, entries 2 and
organocatalytic atroposelective synthesis of NOBIN commenced 3). At this stage, we recognized the following issues: poor solubility
using azonaphthalene 1a as the electrophile. Intriguingly, by-prod- of ZnCl2 in organic solvents could prevent an effective Lewis acid–
ucts (3a-1 and 3a-2, Supplementary Table 1) predominated and the CPA interaction, and both ZnCl2 and CPA could compete to facilitate

Nature Catalysis | www.nature.com/natcatal


NATurE CATAlysIs Articles
Table 1 | Optimization of the reaction conditions for CPA–salt-catalysed asymmetric cross-coupling

H
OH N
Catalyst, Lewis acid N CO2Me
+ H
N OH
N CO2Me CH2Cl2, temp.

1a, 0.11 mmol 2a, 0.1 mmol 3a

(S)-C1: Ar = 4-PhC6H4 Ar
Ar (S)-C2: Ar = 4-CF3C6H4
Ar
(S)-C3: Ar = 3,5-(CF3)2C6H3 O O
O O (S)-C4: Ar = 1-naphthyl P Ca
P O O O O
O OH P Ca (S)-C5: Ar = 2-naphthyl
O O (S)-C6: Ar = 9-phenanthryl
Ar (S)-C7: Ar = 9-anthryl Ar
Ar 2
(S)-C8: Ar = 4-(2-naphthyl)C 6H4 (R)-C9
2 Ar = 2-naphthyl
CPA1: Ar = 4-PhC6H4

Entry Lewis acid Catalyst Temperature (°C) Yield (%) e.e. (%)
1 – CPA1 rt <20 47
2 YbCl3 CPA1 rt <20 −5
3 ZnCl2 CPA1 rt <20 −10
4 ZnCl2 C1 rt 45 −25
5 ZnCl2 + AgBArF C1 rt 75 −33
6a Zn(BArF)2(CH3CN)6 C1 rt 86 −55
7a Zn(BArF)2(CH3CN)6 C1 0 82 −28
8 a
Zn(BArF)2(CH3CN)6 C1 60 87 −72
9a Zn(BArF)2(CH3CN)6 C1 80 87 −70
10b Zn(BArF)2(CH3CN)6 C1 rt 87 −75
11c Zn(BArF)2 C1 rt 88 −77
12 c
Zn(BArF)2 C2 rt 87 −62
13c Zn(BArF)2 C3 rt 75 −35
14c Zn(BArF)2 C4 rt 72 −20
15c Zn(BArF)2 C5 rt 91 −89
16 c
Zn(BArF)2 C6 rt 81 −40
17c Zn(BArF)2 C7 rt 85 −58
18c Zn(BArF)2 C8 rt 97 −86
19c Zn(BArF)2 C9 rt 55 −44
20 c,d
Zn(BArF)2 C5 0 93 −92
Conditions: Lewis acid (5 mol%), catalyst (6 mol%) and CH2Cl2 (6 ml) were added to a Schlenk tube under argon. After stirring at room temperature (rt) for 1 h, 1a and 2a were added and the reaction was
carried out at rt for 20 h. Isolated yield was provided and e.e. was determined by chiral HPLC analysis. aC1 (6 mol%), Zn(BArF)2(CH3CN)6 (5 mol%) and CH2Cl2 (6 ml) were added to a Schlenk tube under
argon. After stirring at 50 °C for 2 h, 1a and 2a were added. bPre-treatment of catalyst: Zn(BArF)2(CH3CN)6 (5 mol%) and C1 (6 mol%) were mixed in DCE (10 ml) for 1 h at 60 °C. Solvent was then removed
at 60 °C and this process was repeated twice to give the freshly prepared catalyst. cPre-treated catalyst: Zn(BArF)2(CH3CN)6 (5 mol%) and C1 (6.5 mol%) were used. dConducted in 8 ml CH2Cl2 at 0 °C.

the reaction, leading to difficult stereocontrol. Correspondingly, model reaction with this pre-generated catalyst gave 3a in 75% e.e.
AgBArF (10 mol%) was reacted with ZnCl2 (5 mol%) to generate at room temperature (Table 1, entry 10). Subsequent studies with
soluble Zn(BArF)2 in situ and CPA–Ca (6 mol%), imposing neg- CPA–Ca loadings found that a combination of 6.5 mol% CPA–Ca
ligible background reaction at room temperature, was enlisted as and 5 mol% Zn(BArF)2(CH3CN)6 was able to deliver 3a in 77% e.e.
the chiral catalyst. Although improved chemical yield and enanti- and 88% yield (Table 1, entry 11). Following these trials, we set out
oselectivity were indeed observed, the instability of AgBArF hin- to refine the reaction conditions by screening various CPAs, and
dered its practicality (Table 1, entries 4 and 5). We resorted to using (S)-C5, with a methyl-spiro backbone, stood out as the best catalyst
the pre-prepared Zn(BArF)2(CH3CN)6 as the Lewis acid instead. in the context of stereoselectivity (Table 1, entries 12–19). A lower
Gratifyingly, 3a was formed with 86% yield and 55% e.e. at room reaction concentration at 0 °C further improved the chemical yield
temperature (Table 1, entry 6). Notably, the reaction at 0 °C gener- to 93% with 92% e.e. (Table 1, entry 20).
ated 3a with 28% e.e., while 72% and 70% e.e. values were obtained
from those at 60 °C and 80 °C, respectively; suggesting the exchange Substrate generality. With the optimal reaction conditions in hand,
of CPA–Ca and Zn(BArF)2(CH3CN)6 had to occur at higher tem- we probed the substrate generality using a broad range of azonaph-
perature (Table 1, entries 7–9). To substantiate this hypothesis and thalene and 2-naphthol derivatives. Overall, all tested substrates
eliminate any unwanted background reaction promoted by excess gave the desired non-C2-symmetrical biaryls in synthetically use-
CPA–Ca at temperatures above 40 °C, this exchange procedure ful yields and enantioselectivities, which were hardly attainable via
was pre-conducted at 60 °C in 1,2-dichloroethane (DCE) and the conventional oxidative cross-coupling (Fig. 2). Swapping the methyl

Nature Catalysis | www.nature.com/natcatal


Articles NATurE CATAlysIs

Ar
Zn(BArF)2(CH3CN)6 (5 mol%) R2 H
OH N
(S)-C5 (6.5 mol%) N CO2 R1 O O
R2 + R3 H P Ca
N OH O O
N CO2R1 DCM (8 ml), 0 °C
3–20 h R3
Ar
2
1 2 3 (S)-C5
Ar = 2-naphthyl

a Substrate scope of azonaphthalenes


Me MeO
H H H H H
N N N N N
N CO2Et N CO2iPr N CO2Bn N CO2Me N CO2Me
H H H H H
OH OH OH OH OH

3b, 88% yield, 91% e.e. 3c, 84% yield, 92% e.e. 3d, 93% yield, 89% e.e. 3e, 97% yield, 92% e.e. 3f, 80% yield, 87% e.e.

Ph Br NC Cl
H H H H H
N N N N N
MeO N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
OH OH OH OH OH

3g, 85% yield, 84% e.e. 3h, 99% yield, 90% e.e. 3i, 86% yield, 90% e.e. 3j, 84% yield, 86% e.e. 3k, 99% yield, 81% e.e.

b Substrate scope of 2-naphthols

H H H H H
N N N N N
N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
Me OH OH OH MeO OH OH

Me Me MeO
3l, 84% yield, 91% e.e. 3m, 99% yield, 91% e.e. 3n, 99% yield, 91% e.e. 3o, 82% yield, 90% e.e. 3p, 82% yield, 84% e.e.

H H H H H
N N N N N
N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
OH Ph OH OH Br OH OH

Ph Br
3q, 70% yield, 90% e.e. 3r, 78% yield, 87% e.e. 3s, 70% yield, 83% e.e. 3t, 72% yield, 90% e.e. 3u, 80% yield, 92% e.e.

H
H H H H N
N N N N N CO2Me
N CO2Me N CO2Me N CO2Me N CO2Me H
H H H H OH
AcO OH OH OH OH

NC OHC MeO2C
CO2Me
3v, 85% yield, 90% e.e. 3w, 73% yield, 95% e.e. 3x, 70% yield, 93% e.e. 3y, 85% yield, 90% e.e. 3z, 85% yield, 87% e.e.

c Phenol as the nucleophile and modifications of both coupling partners

H Me Br Me MeO
N H H H H
N CO2Me N N N N
H N CO2Me N CO2Me N CO2Me N CO2Me
MeO OH H H H H
OH OH OH OH

Me Br Me Me
OMe
3aa, 99% yield, 75% e.e. 3ab, 90% yield, 90% e.e. 3ac, 79% yield, 92% e.e. 3ad, 99% yield, 93% e.e. 3ae, 98% yield, 92% e.e.

Br Ph Me Br
H H H H H
N N N N N
MeO N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
OH OH OH OH OH

Me Me Me Br Me
3af, 95% yield, 91% e.e. 3ag, 99% yield, 93% e.e. 3ah, 99% yield, 94% e.e. 3ai, 78% yield, 90% e.e. 3aj, 87% yield, 89% e.e.

Fig. 2 | Substrate generality for CPA-salt catalysed asymmetric cross-coupling. Reaction conditions: 1 (0.11 mmol), 2 (0.10 mmol) and the freshly
prepared catalyst (5 mol%) in CH2Cl2 (8 ml) at 0 °C. a, Substrate scope of azonaphthalenes. b, Substrate scope of 2-naphthols. c, Modifications of both
coupling partners.

Nature Catalysis | www.nature.com/natcatal


NATurE CATAlysIs Articles
SA SA
Ni(OTf)2 (10 mol%) R1 H
NHR3 N O O
L10 (15 mol%) N CO2Me
R1 + H
N R2 N N
N CO2Me DCM, 0 °C, 4–72 h NHR3
R2 Cy L10 Cy
1 4 5 SA = 3,5-(MeO)2C6H3

a Substrate scope of azonaphthalenes


Me Ph Br
H H H H H
N N N N N
N CO2Me N CO2Me MeO N CO2Me N CO2Me N CO2Me
H H H H H
NHBn NHBn NHBn NHBn NHBn

5a, 99% yield, 94% e.e. 5b, 98% yield, 91% e.e. 5c, 94% yield, 89% e.e. 5d, 96% yield, 91% e.e. 5e, 99% yield, 93% e.e.
Gram-scale : 99% yield, 94% e.e.

b Substrate scope of 2-naphthylamines

H H H H H
N N N N N
N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
NHBn NHBn NHBn NHBn NHBn

Me Ph Br MeO2C Ac
5f, 99% yield, 92% e.e. 5g, 99% yield, 92% e.e. 5h, 96% yield, 91% e.e. 5i, 98% yield, 93% e.e. 5j, 70% yield, 91% e.e.

H H H H
N N N N
N CO2Me N CO2Me N CO2Me N CO2Me
H H H H
Me NHBn MeO NHBn Br NHBn NHBn

5k, 99% yield, 94% e.e. 5l, 99% yield, 95% e.e. 5m, 99% yield, 94% e.e. 5n, 99% yield, 91% e.e.

c Variation of amine protecting group and modifications of both coupling partners


Me
5o, R3 = 4-ClC6H4CH2, 99% yield, 94% e.e. H H
H N
N 5p, R3 = 4-MeOC6H4CH2, 99% yield, 94% e.e. N
N CO2Me N CO2Me N CO2Me
H H H
5q, R3 = 4-PhC6H4CH2, 99% yield, 93% e.e. H
NHR3 N NHBn
5r, R3 = allyl, 66% yield, 90% e.e.
5s, R3 = propargyl, 63% yield, 90% e.e. Me
5t, 93% yield, 93% e.e. 5u, 98% yield, 91% e.e.

Br Br Me Ph
H H H H H
N N N N N
N CO2Me MeO N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H
NHBn MeO NHBn NHBn NHBn Br NHBn

Br Me MeO2C
5v, 96% yield, 91% e.e. 5w, 94% yield, 87% e.e. 5x, 98% yield, 92% e.e. 5y, 98% yield, 91% e.e. 5z, 98% yield, 90% e.e.

Fig. 3 | Substrate scope of BINAM derivatives via Ni(OTf)2/SaBOX-catalysed asymmetric cross-coupling. Reaction conditions: 1 (0.10 mmol),
4 (0.12 mmol), Ni(OTf)2 (10 mol%) and L10 (15 mol%) in CH2Cl2 (4 ml) at 0 °C. a, Substrate scope of azonaphthalenes. b, Substrate scope of
2-naphthylamines. c, Modifications of both coupling partners.

moiety on the ester with an ethyl or isopropyl group provided the and e.e. values were observed (3e, 3h and 3i). The introduction
adduct 3b or 3c in marginally lower yields, while the e.e. value of of a methoxy group at the C6 or C7 position, on the other hand,
3d was slightly reduced with benzyl ester. The substituents on the gave products 3f and 3g with e.e. values of 87% and 84%, respec-
naphthalene ring of azonaphthalene exerted negligible influence on tively. Of note, a C3 substitution resulted in an obvious decline of
the reaction outcome (3e–3k). When methyl, phenyl or bromide the enantioselectivity (3k). Further assessments revealed that all
was appended at the C6 position of the substrate, excellent yields tested 2-naphthols with varied functionalities and substitution

Nature Catalysis | www.nature.com/natcatal


Articles NATurE CATAlysIs

SA SA

Ni(OTf)2 (10 mol%) R2 H H O O H


OH N
L12 (15 mol%) N CO2R1
R2 + R3 H N N
N OH
N CO2R1 NaHCO3, CHCl3, rt 6–16 h
H H
R3 L12
1 2 3 SA = f-tBuC 6H4

a Verifying the method's generality

Me Ph
H H H H H H
N N N N N N
N CO2iPr N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H H
OH OH OH MeO OH Br OH AcO OH

3c, 99% yield, 93% e.e. 3e, 98% yield, 91% e.e. 3h, 96% yield, 91% e.e. 3o, 99% yield, 92% e.e. 3t, 94% yield, 92% e.e. 3v, 98% yield, 91% e.e.

b Improvement from previous results

H H H H H H
N N N N N N
N CO2Bn MeO N CO2Me N CO2Me N CO2Me N CO2Me N CO2Me
H H H H H H
OH OH OH Ph OH OH MeO OH

MeO
OMe
3d, 98% yield, 93% e.e. 3g, 98% yield, 91% e.e. 3p, 99% yield, 92% e.e. 3r, 98% yield, 91% e.e. 3s, 96% yield, 86% e.e.a 3aa, 96% yield, 94% e.e.b

c New substrates explored

Ph Me Br
H H H H H
N N N N N
N CO2Me MeO N CO2Me N CO2Me N CO2Me MeO N CO2Me
H H H H H
OH MeO OH Me OH Br OH Br OH

Ph
3ak, 91% yield, 90% e.e. 3al, 98% yield, 91% e.e. 3am, 94% yield, 91% e.e.c 3an, 96% yield, 90% e.e.c 3ao, 98% yield, 90% e.e.

Fig. 4 | Substrate scope of NOBIN derivatives via Ni(OTf)2/SaBOX-catalysed asymmetric cross-coupling. Reaction conditions: 1 (0.10 mmol),
2 (0.12 mmol), NaHCO3 (0.05 mmol), Ni(OTf)2 (10 mol%) and L12 (15 mol%) in CHCl3 (4 ml) at room temperature. aConducted in 20 ml CHCl3.
b
3 equiv. of 2 was used. cConducted at 15 °C. a, Verifying the generality of the method. b, Improvement from previous results. c, New substrates explored.

patterns could undergo effective transformations to furnish atro- ligand for a model atroposelective transformation of 1a and ben-
pisomers 3l–3z. Interestingly, a C3 methyl-substituted substrate zyl-protected 2-naphthylamine 4a (for details see Supplementary
exhibited commendable compatibility, as evident by the excellent Table 6). A meticulous examination of various reaction parameters
yield and selectivity offered in 3n, while most other products were identified that a combination of 10 mol% of Ni(OTf)2 and 15 mol%
furnished in lower yields, ranging from 70% to 85% (3l, 3o–3z). of sidearm-modified bis(oxazoline) (SaBOX) ligand L10 exhibited
2-Naphthols bearing electron-withdrawing functionalities such as the best catalytic activity with respect to the overall reaction out-
cyano, aldehyde and ester groups underwent the cross-coupling come. This 10 h reaction in CH2Cl2 could form the anticipated prod-
reaction to afford compounds 3×−3z with excellent stereocontrol. uct 5a in quantitative yield with up to 94% e.e. at 0 °C.
Remarkably, a substituted phenol in place of 2-naphthol was also a The optimal reaction conditions were then applied to synthesize
suitable nucleophilic partner, giving 3aa in quantitative yield, albeit a variety of BINAM derivatives. As depicted in Fig. 3, the influence
in only 75% e.e. The synthesis of axially chiral adducts 3ab–3aj was of the substitution on the azonaphthalene 1 was initially investigated
performed using substituted azonaphthalene and 2-naphthol part- and it was apparent that the position and electronics of the substitu-
ners. Experimental data from 3ad–3ah further attested to our pre- ents exerted very limited effect, affording 5b–5e in excellent yields
vious observation that a C3 methyl-substituted naphthol substrate and enantioselectivities. Similar results were obtained with the
performed well under the reaction conditions. introduction of substituents on 2-naphthylamines, giving BINAM
Subsequently, the generality and limitations of our protocol were derivatives 5f–5n in near-quantitative yields and e.e. values exceed-
further evaluated by adopting 2-naphthylamine as the nucleophile ing 90%. This chemistry was also well accommodated by other
to assemble enantioenriched BINAM derivatives. Disappointingly, benzyl-type protecting groups to generate 5o–5q and 5t with excel-
extensive efforts in a massive parallel screening of conditions were lent results, while the employment of allyl- and propargyl-protected
unsuccessful in obtaining good stereocontrol. The best result (74% 2-naphthylamines gave rise to 5r and 5s correspondingly in reduced
yield and 53% e.e.) was obtained for the current CPA–salt complex- chemical yields but similar levels of stereoinduction. The flexibility
based catalytic mode when the reaction was performed in CH2Cl2 in the substituents on both coupling partners increased the product
at −30 °C with 10 mol% of CPA with a methyl-spiro backbone range to also include 5u–5z with satisfactory enantioselectivities.
(Ar = 3,5-Ph2C6H3). Considering the capability of Lewis acids to Notably, 5u–5w were amenable precursors for C2-symmetrical sub-
activate the azo group and the innate difference between 2-naph- stituted BINAMs through facile deprotection and N–N bond cleav-
thylamine and 2-naphthol, we turned to the Lewis acid and chiral age steps. The near-quantitative chemical yields obtained for most

Nature Catalysis | www.nature.com/natcatal


NATurE CATAlysIs Articles
a Synthetic transformations

R1 R1 H
N
NH2 Raney-Ni, H2 (1 atm) N CO2Me Raney-Ni, H2 (30 bar) NH2
H
OH MeOH/KOH aq., 60 °C OH MeOH, rt, 8 h OH
R2 R2
86% yield, 93% e.e.
6 3 7, R1 = H, R2 = H
6a, R1 = R2 = H, 92% yield, 93% e.e.
H Raney-Ni, H2 (1 atm)
N
6g, R1 = MeO, R2 = H, 92% yield, 91% e.e. N CO2Me Ultrasound NH2
H
NHBn MeOH, 50 °C, 18 h NH2
6l, R1 = H, R2 = Me, 94% yield, 92% e.e.

6ak, R1 = R2 = Ph, 88% yield, 90% e.e.


5a, 94% e.e. 8, 82% yield, 94% e.e.

b Mechanistic studies
31P NMR ESI–HRMS of catalyst

O O Zn
O PO
645.1333
646.1329
2

645.6342 647.1330

646.6336

O O Ca 645.4722
O PO 645.7182 646.4698

645.0 645.5 646.0 646.5 647.0


2 m /z
2+ CPA–Ca + Zn(BArF)2

Zn 633.1492
CPA–Ca + Zn(BArF)2 O O O O (BArF)2
O PO P
O O
Ca

633.6514
634.1532
2+
634.5711
635.0729
Ca
CPA–Ca + Ca(BArF)2 O O O O 633.0 633.5 634.0 634.5 635.0 635.5
O PO P
O O
(BArF)2
m /z
Ca
CPA–Ca + Ca(BArF)2
659.1135

2+ 657.1151 658.1157

CPA–Zn + Zn(BArF)2 Zn
O O O O
(BArF)2
OPO P
O O 658.6173
Zn
659.2179

657.6197

657.2658 657.9369

657.0 657.5 658.0 658.5 659.0


m /z
7 –8 –9 –10 –11 –12 –13 –14 –15 –16 -
CPA–Zn + Zn(BArF)2
f1 (ppm)

Fig. 5 | Synthetic transformations and mechanistic studies. a, Synthetic transformations of NOBIN and BINAM derivatives. (b, Mechanistic investigations
for CPA–salt complex-based catalytic system.

biaryl products demonstrate the practicality of this method. Also, a temperature. For the gram-scale reactions, no deterioration of
gram-scale reaction delivered 5a in analogous yield and enantiose- reaction outcome was observed under standard conditions, and
lectivity, implying that this protocol should be well suited for large- a slightly lower e.e. (91%) was only observed with 2.5 mol% cata-
scale chemical production. The absolute configuration of 5o was lyst loading (Supplementary Fig. 2). Selected substrates with
determined by X-ray crystallographic analysis after recrystallization diverse ester groups as well as substitutions on azonaphthalene 1
and those of other products in Fig. 3 were assigned by analogy. and 2-naphthol 2 were examined, all of which accomplished the
We next investigated the compatibility and superiority of reaction in stereoselectivities similar to those with the previous
this Ni/chiral BOX ligand system to construct NOBIN deriva- CPA–salt complex but with higher yields (3c, 3e, 3h, 3o, 3t and
tives with respect to the CPA–salt complex-based catalytic sys- 3v). This method also complemented the former approach by
tem. Although these conditions were first found inapplicable for enhancing the stereochemical outcome for substrates formed in
the 2-naphthol nucleophile, exciting results were obtained after poorer enantioselectivities (3d, 3g, 3p, 3r, 3s and 3aa). Finally,
switching the chiral ligand and reaction solvent and including five additional highly enantioenriched NOBIN derivatives were
NaHCO3 as the base. The best results (99% yield and 93% e.e.) prepared with excellent stereocontrol, highlighting the expedi-
were secured with the reaction conducted in CHCl3 at room ency and breadth of this protocol (Fig. 4).

Nature Catalysis | www.nature.com/natcatal


Articles NATurE CATAlysIs

Table 2 | Control experiments

H
OH N
Catalyst N CO2Me
+ H
N OH
N CO2Me CH2Cl2 (8 ml), 0 °C, 2 days

1a, 0.11 mmol 2a, 0.10 mmol 3a

Entry Catalyst system Yield (%) e.e. (%)


1 Ca(CPA)2 (10 mol%) – –
2 Zn(CPA)2 (10 mol%) 15 38
3 Ca(CPA)2 (3 mol%) Zn(CPA)2 (10 mol%) 10 70
4 Zn(BArF)2 (5 mol%) Zn(CPA)2 (6.5 mol%) 95 70
5 Ca(BArF)2 (5 mol%) Ca(CPA)2 (6.5 mol%) 18 36
6 Zn(BArF)2 (5 mol%) Ca(CPA)2 (6.5 mol%) 93 92
Conditions: The catalyst (entries 1–2) or its solution (entries 3–6) was added to the solution of 1a (0.11 mmol) and 2a (0.10 mmol) in CH2Cl2 at 0 °C. The reaction was then carried out at 0 °C for 2 days.
Pre-treatment of catalyst (entries 3–6): the two catalysts were mixed in DCE (10 ml) for 1 h at 60 °C, the solvent was then removed at 60 °C. This process was repeated twice to give the freshly prepared
catalyst, which was dissolved in 2 ml of CH2Cl2 to form its solution.

Ni(OTf)2 + SaBOX

1
NOBIN (3) [NiL*]2+ 2 –OTf

Chirality transfer

2+
2+ O
H OMe
N
N O
H H O R N
O Ni N
Ni
L* N O
N R
C
OMe A

2 + Base

Base O +

O O [Base-H]+ –OTf
[Base-H]+ –OTf R N
Ni N
N O
N R
OMe
B

Fig. 6 | Plausible mechanism for Ni-catalysed asymmetric cross-coupling. The catalytic cycle involes stereocontrol nucleophilic aromatic substitution and
a rearomatization process along with central-to-axial chirality conversion.

Synthetic transformations. The transfer of axial chirality com- Mechanistic studies. To acquire detailed mechanistic insights, a
pleted with high fidelity when several NOBIN derivatives (3a, 3g, 3l series of control experiments were performed. First, a nonlinear
and 3ak) were converted to corresponding NOBINs 6 in good yields relationship between the enantiopurity of the catalyst and product
via N–N bond cleavage in basic conditions with Raney nickel cata- was observed, implying the generation of an active species contain-
lyst under a 1 atm hydrogen atmosphere. Notably, NOBIN 6a could ing two chiral phosphate anions and precluding single metal (Zn or
be further hydrogenated to an analogue 7 in one pot with syntheti- Ca) phosphate complex catalysis with a BArF counterion (for details
cally useful yield and enantioselectivity when the reaction was car- see Supplementary Fig. 3)31,42,43. Furthermore, the lack of product
ried out under higher pressure under neutral conditions (Fig. 5a). formation with the CPA–Ca catalyst at 0 °C was consistent with the
Similarly, enantiopure BINAM 8 was obtained in 82% yield with negligible background reaction below room temperature. The pres-
high enantioretention (94%) from 5a. The absolute configuration of ence of CPA–Zn (10 mol%) alone furnished 3a in 15% yield with
3a was assigned as (S) by comparing the chiral HPLC spectrum with 38% e.e. At the same time, the increased stereoselectivity of 70% e.e.
that of commercially available (S)-NOBIN. Those of other products obtained with 3 mol% CPA–Ca and 10 mol% CPA–Zn implied for-
3 were assigned analogously. mation of the Ca–CPA–Zn complex, vital for chirality induction.

Nature Catalysis | www.nature.com/natcatal


NATurE CATAlysIs Articles
The low yield was attributed to the poor solubility of the salts in Synthesis of (S)-3 by Ni(OTf)2/SaBOX. A mixture of Ni(OTf)2 (3.6 mg,
CH2Cl2 when the BArF counterion was absent. Additionally, reac- 0.01 mmol), bisoxazoline L12 (9.3 mg, 0.015 mmol) and NaHCO3 (4.2 mg,
0.05 mmol) in CHCl3 (4 ml) was stirred at room temperature overnight under a
tion with Zn(CPA)2 and Zn(BArF)2 afforded the target product nitrogen atmosphere. Compound 1 (0.1 mmol) and 2-naphthol 2 (0.12 mmol) were
in 95% yield with 70% e.e., but only 18% yield and 36% e.e. were then added to the reaction. The mixture was stirred at room temperature until TLC
achieved with Ca(CPA)2 and Ca(BArF)2 as co-catalyst. The most indicated that compound 1 had disappeared. The reaction mixture was directly
optimal results (92% e.e. and 93% yield) were obtained when using a purified by flash chromatography on silica gel to afford the corresponding NOBIN
derivatives (S)-3.
mixture of 6.5 mol% Ca(CPA)2 and 5 mol% Zn(BArF)2 as co-catalyst
(catalyst ratio of the optimized conditions) (Table 2). Pronounced Synthesis of (R)-5 by Ni(OTf)2/SaBOX. A mixture of Ni(OTf)2 (3.6 mg,
variation of the chemical shifts in 31P NMR (ZnCa(CPA)22+, 0.01 mmol) and BOX L10 (9.3 mg, 0.015 mmol) in CH2Cl2 (4 ml) was stirred at
−11.88 ppm; Ca2(CPA)22+, −11.46 ppm; Zn2(CPA)22+, −8.13 ppm) as room temperature overnight under a nitrogen atmosphere. The reaction was
well as ESI-HRMS analysis of the freshly prepared catalyst showing cooled to 0 °C for 15 min and compound 1 (0.1 mmol) and 2-naphthylamine 4 were
added to the reaction. The mixture was stirred at 0 °C until TLC indicated that
a peak of m/z 645.1333 bearing two positive charges further point to compound 1 had disappeared. The reaction mixture was directly purified
ZnCPA2Ca2+ as the coordination mode of the active catalyst (Fig. 5b). by flash chromatography on silica gel to afford the corresponding BINAM
Based on the experimental results from the Ni(OTf)2/chiral derivatives (R)-5.
BOX ligand catalytic system and reference to preceding mechanistic
investigations of Ni-catalysed asymmetric reactions44–48, a plausible Data availability
catalytic cycle was outlined in Fig. 6. This reaction is thought to The X-ray crystallographic coordinates for structures of 3a-2 and 5o reported in
this Article have been deposited at the Cambridge Crystallographic Data Centre
commence with coordination of compound 1 with the nickel(ii) and (CCDC) under deposition numbers CCDC 1846849 (3a-2) and 1846841 (5o).
SaBOX ligand complex in a bidentate fashion, giving rise to species These data can be obtained free of charge from http://www.ccdc.cam.ac.uk/data_
A, with enhanced electrophilicity on the aromatic ring. Subsequently, request/cif. Experimental procedures and characterization of the new compounds
the deprotonated 2-naphthol 2 approaches A from the apical posi- are available in the Supplementary Information. All other data are available from
tion, which points away from the indane group of the BOX ligand the authors upon reasonable request.
for minimal steric repulsion and sits parallel to the naphthalene ring Received: 5 September 2018; Accepted: 5 February 2019;
of 1a due to π–π stacking (species B). Compound 2 then performs Published: xx xx xxxx
nucleophilic attack on activated azonaphthalene from the Si face to
generate C bearing two adjacent central chiralities. An efficient cen- References
tral-to-axial chirality conversion featuring a rearomatization process 1. Bringmann, G., Gulder, T., Gulder, T. A. M. & Breuning, M. Atroposelective
via C–H bond cleavage and N–H/O–H bond formation gives the total synthesis of axially chiral biaryl natural products. Chem. Rev. 111,
desired thermodynamically stable compound 3. 563–639 (2011).
2. McCormick, M. H., Stark, W. M., Pittenger, G. E., Pittenger, R. C. &
McGuire, J. M. Vancomycin, a new antibiotic. I. Chemical and biologic
Conclusions properties. Antibiot. Annu. 56, 606–611 (1955).
In summary, this work signifies a major departure from transition- 3. Bremner, J. B. et al. Synthesis and antibacterial studies of binaphthyl-based
metal based cross-coupling reactions to generate a myriad of unex- tripeptoids. Bioorg. Med. Chem. 18, 2611–2620 (2010).
plored NOBIN and BINAM derivatives in excellent yields and 4. Cram, D. J. & Cram, J. M. Host–guest chemistry. Science 183,
803–809 (1974).
enantiopurities through a rationally designed redox-neutral cross- 5. Pu, L. Enantioselective fluorescent sensors: a tale of BINOL. Acc. Chem. Res.
coupling strategy. First, a CPA–salt complex-based system catalysed 45, 150–163 (2012).
the asymmetric redox-neutral cross-coupling of azonaphthalenes 6. Sogah, G. D. Y. & Cram, D. J. Total chromatographic optical resolution of
and 2-naphthols to yield NOBIN derivatives in which mechanistic α-amino acid and ester salts through chiral recognition by a host covalently
bound to polystyrene resin. J. Am. Chem. Soc. 98, 3038–3041 (1976).
studies credited a ZnCPA2Ca2+ complex as the actual active catalyst.
7. Li, Q., Green, L., Venkataraman, N., Shiyanovskaya, I. & Khan, A. Reversible
Complementarily, optically pure BINAM derivatives were prepared photoswitchable axially chiral dopants with high helical twisting power.
from 2-naphthylamines with a Ni(OTf)2 and chiral BOX ligand J. Am. Chem. Soc. 129, 12908–12909 (2007).
catalytic system. This Lewis acid catalysis approach was applicable 8. Takaishi, K., Yasui, M. & Ema, T. Binaphthyl-bipyridyl cyclic dyads as a
to 2-naphthol substrates by marginal modification of the reaction chiroptical switch. J. Am. Chem. Soc. 140, 5334–5338 (2018).
9. Ritter, N., Senkovska, I., Kaskel, S. & Weber, J. Towards chiral microporous
conditions, allowing access to even more NOBIN derivatives with soluble polymers-binaphthalene-based polyimides. Macromol. Rapid
improved chemical yields and enantioselectivities, thereby comple- Commun. 32, 438–443 (2011).
menting the former. Adept enantioretentive chemical transforma- 10. Cui, Y., Lee, S. J. & Lin, W. Interlocked chiral nanotubes assembled from
tions on the obtained products to NOBINs and BINAMs through quintuple helices. J. Am. Chem. Soc. 125, 6014–6015 (2003).
Raney nickel-catalysed hydrogenation showed the synthetic utility of 11. Noyori, R. & Takaya, H. BINAP: an efficient chiral element for asymmetric
catalysis. Acc. Chem. Res. 23, 345–350 (1990).
our protocols. This work is expected to give new impetus to the explo- 12. Pu, L. 1,1′-Binaphthyl dimers, oligomers, and polymers: molecular
ration of these axially chiral molecules in terms of both synthesis and recognition, asymmetric catalysis, and new materials. Chem. Rev. 98,
application. Furthermore, our forays into the direct redox neutral 2405–2494 (1998).
cross-coupling strategy would open up new vistas for the asymmetric 13. Kočovský, P., Vyskočil, Š. & Smrčina, M. Non-symmetrically substituted
construction of other biaryls with unmined application potentials. 1,1′-binaphthyls in enantioselective catalysis. Chem. Rev. 103,
3213–3245 (2003).
14. Chen, Y., Yekta, S. & Yudin, A. K. Modified BINOL ligands in asymmetric
Methods catalysis. Chem. Rev. 103, 3155–3211 (2003).
Materials. Chemicals were purchased from commercial suppliers and used without 15. Ding, K., Guo, H., Li, X., Yuan, Y. & Wang, Y. Synthesis of NOBIN
further purification unless otherwise stated. derivatives for asymmetric catalysis. Top. Catal. 35, 105–116 (2005).
16. Ding, K., Li, X., Ji, B., Guo, H. & Kitamura, M. Ten years of research on
Synthesis of (S)-3 by CPA–Ca/Zn(BArF)2. Pre-treatment of catalyst: NOBIN chemistry. Curr. Org. Syn. 2, 499–545 (2005).
Zn(BArF)2(CH3CN)6 (51 mg, 0.025 mmol) and CPA–Ca (39 mg, 0.0325 mmol) were 17. Berthod, M., Mignani, G., Woodward, G. & Lemaire, M. Modified BINAP:
mixed in DCE (30 ml) for 1 h at 60 °C. The solvent was then removed in vacuo at the how and the why. Chem. Rev. 105, 1801–1836 (2005).
60 °C and this process was repeated twice to give the freshly prepared catalyst, which 18. Brunel, J. M. BINOL: a versatile chiral reagent. Chem. Rev. 107, PR1–PR45 (2007).
was dissolved in 10 ml of DCM to form 2.5 mM l−1 catalyst. Standard synthetic 19. Kozlowski, M. C., Morgan, B. J. & Linton, E. C. Total synthesis of chiral
procedure: 2-naphthol 2 (0.1 mmol) was added to a solution of freshly prepared biaryl natural products by asymmetric biaryl coupling. Chem. Soc. Rev. 38,
catalyst (5 mol%) and compound 1 (0.11 mmol) in CH2Cl2 (8 ml) at 0 °C. The mixture 3193–3207 (2009).
was then stirred at 0 °C for 3–20 h. After completion of the reaction (monitored 20. Wencel-Delord, J., Panossian, A., Leroux, F. R. & Colobert, F. Recent
by thin-layer chromatography, TLC), the solution was directly purified by flash advances and new concepts for the synthesis of axially stereoenriched biaryls.
chromatography on silica gel to afford the corresponding NOBIN derivatives (S)-3. Chem. Soc. Rev. 44, 3418–3430 (2015).

Nature Catalysis | www.nature.com/natcatal


Articles NATurE CATAlysIs
21. Zhang, D. & Wang, Q. Palladium catalyzed asymmetric Suzuki–Miyaura 39. Wang, Y.-B. & Tan, B. Construction of axially chiral compounds via
coupling reactions to axially chiral biaryl compounds: chiral ligands and asymmetric organocatalysis. Acc. Chem. Res. 51, 534–547 (2018).
recent advances. Coord. Chem. Rev. 286, 1–16 (2015). 40. Akiyama, T., Itoh, J., Yokota, K. & Fuchibe, K. Enantioselective Mannich-type
22. Loxq, P., Manoury, E., Poli, R., Deydier, E. & Labande, A. Synthesis of axially reaction catalyzed by a chiral Brønsted acid. Angew. Chem. Int. Ed. 43,
chiral biaryl compounds by asymmetric catalytic reactions with transition 1566–1568 (2004).
metals. Coord. Chem. Rev. 308, 131–190 (2016). 41. Uraguchi, D. & Terada, M. Chiral Brønsted acid-catalyzed direct
23. Zilate, B., Castrogiovanni, A. & Sparr, C. Catalyst-controlled stereoselective Mannich reactions via electrophilic activation. J. Am. Chem. Soc. 126,
synthesis of atropisomers. ACS Catal. 8, 2981–2988 (2018). 5356–5357 (2004).
24. Yamaguchi, K., Yamaguchi, J., Studerb, A. & Itami, K. Hindered biaryls by 42. Satyanarayana, T., Abraham, S. & Kagan, H. B. Nonlinear effects in
C–H coupling: bisoxazoline-Pd catalysis leading to enantioselective C–H asymmetric catalysis. Angew. Chem. Int. Ed. 48, 456–494 (2009).
coupling. Chem. Sci. 3, 2165–2169 (2012). 43. Rauniyar, V. et al. Asymmetric electrophilic fluorination using an anionic
25. Yamaguchi, K., Kondo, H., Yamaguchi, J. & Itami, K. Aromatic C–H coupling chiral phase-transfer catalyst. Science 334, 1681–1683 (2011).
with hindered arylboronic acids by Pd/Fe dual catalysts. Chem. Sci. 4, 44. Pellissier, H. Recent developments in enantioselective nickel(ii)-catalyzed
3753–3757 (2013). conjugate additions. Adv. Synth. Catal. 357, 2745–2780 (2015).
26. Dherbassy, Q., Djukic, J.-P., Wencel-Delord, J. & Colobert, F. Two 45. Liao, S., Sun, X.-L. & Tang, Y. Side arm strategy for catalyst design: modifying
stereoinduction events in one C–H activation step: a route towards bisoxazolines for remote control of enantioselection and related. Acc. Chem.
terphenyl ligands with two atropisomeric axes. Angew. Chem. Int. Ed. 57, Res. 47, 2260–2272 (2014).
4668–4672 (2018). 46. Desimoni, G., Faita, G. & Jørgensen, K. A. C2-Symmetric chiral bis(oxazoline)
27. Smrčina, M., Lorenc, M., Hanuš, V. & Kočovský, P. A facile synthesis of ligands in asymmetric catalysis. Chem. Rev. 106, 3561–3651 (2006).
2-amino-2′-hydroxy-1,1′-binaphthyl and 2,2′-diamino-1,1′-binaphthyl by 47. Evans, D. A. & Thomson, R. J. Ni(ii) Tol-BINAP-catalyzed enantioselective
oxidative coupling using copper (ii) chloride. Synlett 1991, 231–232 (1991). orthoester alkylations of N-acylthiazolidinethiones. J. Am. Chem. Soc. 127,
28. Smrčina, M. et al. Selective cross-coupling of 2-naphthol and 2-naphthamine 10506–10507 (2005).
derivatives. A facile synthesis of 2,2′,3-trisubstituted and 48. Evans, D. A., Downey, C. W. & Hubbs, J. L. Ni(ii) bis(oxazoline)-catalyzed
2,2′,3,3′-tetrasubstituted 1,1′-binaphthyls. J. Org. Chem. 59, enantioselective syn aldol reactions of N-propionylthiazolidinethiones in the
2156–2163 (1994). presence of silyl triflates. J. Am. Chem. Soc. 125, 8706–8707 (2003).
29. Qi, L.-W., Mao, J.-H. & Tan, B. Organocatalytic asymmetric arylation of
indoles enabled by azo groups. Nat. Chem. 10, 58–64 (2018). Acknowledgements
30. Singer, R. A. & Buchwald, S. L. Preparation of 2-amino-2′-hydroxy-1,1′- The authors acknowledge financial support from the National Natural Science
binaphthyl and N-arylated 2-amino-1,1′-binaphthyl derivatives via Foundation of China (grants nos. 21772081, 21825105, 21702092 and 21801117),
palladium-catalyzed amination. Tetrahedron Lett. 40, 1095–1098 (1999). Shenzhen Nobel Prize Scientists Laboratory Project (C17213101) and Shenzhen Special
31. De, C. K., Pesciaioli, F. & List, B. Catalytic asymmetric benzidine Funds for the Development of Biomedicine, Internet, New Energy and New Material
rearrangement. Angew. Chem. Int. Ed. 52, 9293–9295 (2013). Industries (JCYJ20170412151701379 and KQJSCX20170328153203). We dedicate this
32. Li, G. et al. Organocatalytic aryl–aryl bond formation: an atroposelective paper to the memory of L.-R. Jin.
[3,3]-rearrangement approach to BINAM dericatives. J. Am. Chem. Soc. 135,
7414–7417 (2013).
33. Kumarasamy, E., Raghunathan, R., Sibi, M. P. & Sivaguru, J. Nonbiaryl and Author contributions
heterobiaryl atropisomers: molecular templates with promise for B.T. and J.W. conceived and directed the project. L.-W.Q. and S.L. designed and
atropselective chemical transformations. Chem. Rev. 115, performed experiments and prepared the Supplementary Information. S.-H.X. helped
11239–11300 (2015). collect some new compounds and in analysing the data. B.T., J.W. and S.-H.X. wrote the
34. Gustafson, J. L., Lim, D. & Miller, S. J. Dynamic kinetic resolution of biaryl paper. All authors discussed the results and commented on the manuscript.
atropisomers via peptide-catalysed asymmetric bromination. Science 328,
1251–1255 (2010). Competing interests
35. Mori, K. et al. Enantioselective synthesis of multisubstituted biaryl skeleton The authors declare no competing interests.
by chiral phosphoric acid catalysed desymmetrization/kinetic resolution
sequence. J. Am. Chem. Soc. 135, 3964–3970 (2013).
36. Barrett, K. T., Metrano, A. J., Rablen, P. R. & Miller, S. J. Spontaneous transfer Additional information
of chirality in an atropisomerically enriched two-axis system. Nature 509, Supplementary information is available for this paper at https://doi.org/10.1038/
71–75 (2014). s41929-019-0247-1.
37. Wang, J.-Z. et al. Symmetric in cascade chirality-transfer processes: a catalytic Reprints and permissions information is available at www.nature.com/reprints.
atroposelective direct arylation approach to BINOL derivatives. J. Am. Chem. Correspondence and requests for materials should be addressed to J.(.W. or B.T.
Soc. 138, 5202–5205 (2016).
38. Jolliffe, J. D., Armstrong, R. J. & Smith, M. D. Catalytic enantioselective Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
synthesis of atropisomeric biaryls by a cation-directed O-alkylation. published maps and institutional affiliations.
Nat. Chem. 9, 558–562 (2017). © The Author(s), under exclusive licence to Springer Nature Limited 2019

Nature Catalysis | www.nature.com/natcatal

You might also like