Next Article in Journal
Durability of Flame-Retarded, Co-Extruded Profiles Based on High-Density Polyethylene and Wheat Straw Residues
Next Article in Special Issue
Diaryl-Pyrano-Chromenes Atropisomers: Stereodynamics and Conformational Studies
Previous Article in Journal
Advanced Oxidation Processes and Biotechnological Alternatives for the Treatment of Tannery Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copper-Catalyzed Synthesis of Axially Chiral Biaryls with Diaryliodonium Salts as Arylation Reagents

1
Shenzhen Key Laboratory of Small Molecule Drug Discovery and Synthesis, Department of Chemistry, Southern University of Science and Technology, Shenzhen 518055, China
2
Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(11), 3223; https://doi.org/10.3390/molecules26113223
Submission received: 9 May 2021 / Revised: 17 May 2021 / Accepted: 20 May 2021 / Published: 27 May 2021

Abstract

:
NOBIN and BINAM derivatives harboring biaryl frameworks are recognized as a class of important atropisomers with versatile applications. Here, we present an efficient synthetic route to access such compounds through copper-catalyzed domino arylation of N-arylhydroxylamines or N-arylhydrazines with diaryliodonium salts and [3,3]-sigmatropic rearrangement. This reaction features mild conditions, good substrate compatibility, and excellent efficiency. The practicality of this protocol was further extended by the synthesis of biaryl amino alcohols.

1. Introduction

Axially chiral biaryl frameworks constitute the core structure of a wide range of natural products and biologically active molecules. They are also found widespread applications as chiral catalysts and ligands. In the asymmetric catalysis field, 2-amino-2′-hydroxy-1,1′-binaphthyl (NOBIN) and 1,1′-binaphthyl-2,2′-diamine (BINAM) are among the most frequently utilized structures [1,2,3,4]. Now, NOBIN and BINAM derivatives have been involved in metal catalysis [5,6,7], organocatalysis [8,9], photocatalysis [10], and even heterogeneous catalysis [11] for effective chirality induction. Meanwhile, the significance of such backbones is further illustrated by their prevalence in functional materials [12,13] (Figure 1). Accordingly, the construction of NOBIN and BINAM scaffolds has attracted extensive attentions from the synthetic research community.
For NOBIN and its derivatives, they could be accessed from other binaphthyl compounds such as BINOL or BINAM [14,15,16]. However, in these synthetic processes, excess noble metal reagents, harsh conditions, or expensive reagents were commonly required to achieve satisfactory efficiency [17]. The oxidative cross-coupling of 2-naphthol and 2-naphthylamine catalyzed by transition metal represents the most effective and direct method to establish the aryl–aryl axis. Kočovský and coworkers pioneered the strategy of oxidative cross-coupling using copper amine complexes as oxidants [18,19,20,21]. Subsequently, Ding [22] and Carreira [23] provided a series of improved approaches to enhance the synthetic efficiency and applicability in large-scale preparations and inhibit inseparable homo-coupling by-products. Recently, Tu and coworkers successfully constructed enantioenriched 3,3′-disubstituted NOBINs by aerobic oxidative cross-coupling utilizing a novel Cu/SPDO catalytic system [24]. The redox potential difference between two coupling partners ensured good chemoselectivity and chemical yield during the coupling process. Our group developed an efficient coupling approach for the synthesis of NOBINs via a palladium-catalyzed highly site-selective C-H arylation reaction of N-Boc-2-naphthylamines with diazoquinones under mild conditions [25].
As an effective Ar-Ar bond formation method, [3,3]-sigmatropic rearrangement reaction was emerged as an attractive alternative [26,27,28,29,30,31,32]. In this context, Gao [33] and our group [34] independently developed a transition metal-free approach to generate NOBIN derivatives following a domino arylation of naphthylhydroxylamines with diaryliodonium salts and [3,3]-sigmatropic rearrangement. It should be mentioned that moderate yields were normally obtained for Gao’s conditions, while the mixed solvent of dichloromethane and trifluoroethanol was required to improve reaction results.
The progress in synthesis of biaryls employing diaryliodonium salts as aryl cation equivalents has made this class of bench-stable, nontoxic, and readily available reagents attract attention [35,36,37,38,39]. Moreover, copper catalyst can be oxidized in the presence of diaryliodonium salts to form a highly electrophilic aryl-Cu(III) intermediate and a range of latent nucleophiles undergo arylation reactions to form synthetically versatile products [39,40]. In view of the advantages and reliability with diaryliodonium salts, Cu-catalyzed arylation of N-arylhydroxylamine or N-arylhydrazine can effectively generate transient diaryl groups linked by heteroatoms, which can easily undergo rearrangement reactions. Motivated by our continuous research interests in constructing biaryl frameworks [41,42,43,44], we turned our attention to construct the NOBIN and BINAM derivatives via copper-catalyzed N-/O-arylation with diaryliodonium salts and subsequent [3,3]-sigmatropic rearrangement under mild conditions (Figure 2). The properties of biaryls are affected by the steric hindrance and electronic effect of substituents, which will bring new opportunities in application and expansion. The arylation-rearrangement sequence that allows new library synthesis is still desired.

2. Results and Discussion

2.1. Optimization of Reaction Conditions

Upon exploring some reaction conditions through variation of the copper catalysts, solvents and bases (see Supplementary Materials for details), the following protocol was identified to be optimal: reaction of 1a and 2a with the molar ratio of 1.0:1.2 by using Cu(TFA)2 (10 mol%) as catalyst in dichloromethane (DCM) at room temperature, 3a was obtained almost quantitatively (Table 1, entry 1). When evaluating different solvents, DCM outcompeted others to form a desired product (Table 1, entries 2–5). As far as the catalyst is concerned, other screened Cu(II) or Cu(I) bearing different anions also gave 3a in high yield under mild conditions (Table 1, entries 6–9). Finally, other carbonate salts, NaOH, tBuOK, and amine are inferior to Na2CO3 in facilitating the arylation process (Table 1, entries 10–14).

2.2. Substrate Scope

With the optimized conditions in hand, the generality of this transformation was then explored with respect to N-naphthylhydroxylamines 1 and diaryliodonium salts 2. As shown in Figure 3, all the investigated substrates were completely transformed and furnished the respective product in generally high efficiency with a yield of up to 98%. In detail, the replacement of the Cbz-protecting group with a methyl formate gave the desired product in 91% yield. Different substituents on the aromatic ring including halides, methyl ester, phenyl, and methoxy were all compatible for this set of reaction conditions, and meanwhile, the substitution patterns and electronic properties of substituents exerted a limited influence on the reaction outcome. Further evaluations revealed that all the tested N-naphthylhydroxylamines 1 and diaryliodonium salts 2 with varied substitutions could undergo effective combination to give multi-substituted NOBINs 3mx in 81–92% yield. In addition, diaryliodonium salt with an extended fused ring system proved to be an applicable arylation reagent and produced the corresponding NOBINs in about 90% yield (products 3vx). It should be mentioned that the Br atom, which could act as an effective handle for further transformation, survived during this process.
Subsequently, N-phenylhydroxyamines or diphenyliodonium salts were evaluated for this reaction to synthesize the biaryl amino alcohols. A series of structurally diverse compounds (Figure 4a, 3yac) were obtained in good yields under the standard conditions. Cyclic diaryliodonium salt was verified to be a suitable arylation reagent, and the expected diaxial product was obtained in 72% yield as a pair of diastereomers with a ratio of 1.2:1 (Figure 4a, 3ad). The successful establishment of a highly efficient domino approach to construct NOBINs inspired us to explore the feasibility in constructing BINAMs, which is another type of privileged biaryl atropisomers, to further extend the applicability and flexibility of the developed method. Pleasingly, when N-naphthylhydrazines 4 were utilized, the reactions with diaryliodonium salts 2 underwent smoothly to give BINAMs in moderate yield (Figure 4b, 5ad).

2.3. Control Experiments and Plausible Mechanism

Under transition metal-free conditions, the substrates could be completely converted, and the desired product 3a was obtained in 70% yield (Figure 5a), along with several by-products. The use of Cu(TFA)2 not only improved the yield significantly but also shortened the reaction time, indicating that copper salt had an obvious catalytic effect on this type of reactions. Moreover, other examined Lewis acids such as the triflate of aluminum, magnesium, zinc, or nickel brought about a negligible effect on the reaction outcome (Figure 5b). In addition, a stoichiometric base was necessary for this reaction. When sodium carbonate was removed from the standard conditions, the target product 3a could only be obtained in 33% yield (Figure 5c).
According to the results of the control experiments and previous reports on copper-catalyzed arylation reactions with diaryliodonium salts [36,39,45,46,47,48], a plausible reaction pathway involving 2-naphthyl-Cu(III) species was proposed, as shown in Figure 6. At first, the Cu(I) salt initially formed by Cu(II) disproportionation [49,50,51] undergoes oxidative addition into the Ar-I(III) bond to form the highly electrophilic aryl-Cu(III) intermediate A. Then, the complexation or nucleophilic substitution of aryl-Cu(III) species A with N-arylhydroxylamine 1 produces intermediate B under basic conditions. Upon reductive elimination, N,O-dinaphthylhydroxylamine C is generated, and active Cu(I) catalyst is released to continue the catalytic cycle. Next, the [3,3]-sigmatropic rearrangement step and subsequent rearomatization proceed rapidly to afford the product NOBIN 3. As a transient precursor, C is quite difficult to be isolated from the reaction system, indicating a strong driving force for the following rearrangement.

3. Materials and Methods

Reagents were purchased at the highest commercial quality and used without further purification, unless otherwise stated. Cu(TFA)2 was purchased from Energy Chemical (Shanghai, China); Na2CO3 was purchased from Aladdin (Shanghai, China); Dichloromethane was purchased from TiTan (Shanghai, China). Analytical thin layer chromatography (TLC) was performed on precoated silica gel 60 F254 plates (Qingdao, China). Flash column chromatography was performed using Tsingdao silica gel (60, particle size 0.040–0.063 mm; Qingdao, China). Visualization on TLC was achieved by use of UV light (254 nm). NMR spectra were recorded on a Bruker DPX 400 spectrometer (Bruker BioSpin GmbH, Rheinstetten, Germany) at 400 MHz for 1H-NMR, 100 MHz for 13C-NMR and 376 MHz for 19F-NMR in CDCl3 or Acetone-d6 with tetramethylsilane (TMS) as internal standard. Chemical shifts are reported in ppm and coupling constants are given in Hz. Data for 1H-NMR are recorded as follows: chemical shift (ppm), multiplicity (s, singlet; d, doublet; t, triplet; m, multiplet), coupling constant (Hz), integration. Data for 13C-NMR are reported in terms of chemical shift (δ, ppm). High resolution mass spectra (HRMS) were recorded on a LC-TOF spectrometer (Thermo Fisher Scientific, Waltham, MA, USA).
General procedures for synthesis of NOBIN and BINAM derivatives: 1 or 4 (0.20 mmol), 2 (0.24 mmol), Na2CO3 (27.6 mg, 0.26 mmol), and Cu(TFA)2 (5.8 mg, 10 mol%) were added to a bottle with a magnetic stirring bar. DCM (4.0 mL) was added, and the reaction mixture was stirred at room temperature until 1 or 4 was completely consumed (monitored by TLC). After the solvent evaporated, the residue was purified by flash chromatography eluted with DCM to afford the corresponding product 3 or 5.
Benzyl-(2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3a) White solid. Yield: 98%. 1H-NMR (400 MHz, CDCl3) δ 8.56 (d, J = 9.6 Hz, 1H), 8.08 (d, J = 9.2 Hz, 1H), 8.00 (d, J = 9.2 Hz, 1H), 7.94 (t, J = 7.2 Hz, 2H), 7.47–7.27 (m, 10H), 7.15 (d, J = 8.8 Hz, 1H), 7.05 (d, J = 8.8 Hz, 1H), 6.54 (s, 1H), 5.21 (s, 1H), 5.07 (s, 2H). 13C-NMR (100 MHz, CDCl3) δ 153.6, 152.0, 135.9, 135.7, 133.2, 132.9, 131.3, 130.8, 130.4, 129.4, 128.6, 128.5, 128.4, 128.3, 128.3, 127.5, 127.4, 125.3, 125.1, 124.1, 124.0, 119.7, 117.9, 116.7, 112.7, 67.1. HRMS (ESI) calcd for [M + H] C28H22NO3, m/z: 420.1594, found: 420.1595.
Methyl-(2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3b) Yellowish solid. Yield: 91%. 1H-NMR (400 MHz, CDCl3) δ 8.37 (d, J = 9.2 Hz, 1H), 7.89 (d, J = 8.8 Hz, 1H), 7.84 (d, J = 9.2 Hz, 1H), 7.77 (d, J = 8.0 Hz, 2H), 7.30–7.22 (m, 3H), 7.16–7.12 (m, 2H), 7.00 (d, J = 8.4 Hz, 1H), 6.89 (d, J = 8.0 Hz, 1H), 6.34 (s, 1H), 5.30 (s, 1H), 3.39 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 154.0, 152.1, 135.9, 133.2, 133.0, 131.2, 130.7, 130.3, 129.4, 128.4, 128.3, 127.5, 127.3, 125.1, 125.1, 124.1, 124.0, 119.3, 118.0, 116.5, 112.7, 52.3. HRMS (ESI) calcd for [M + H] C22H18NO3, m/z: 344.1281, found: 344.1281.
Benzyl-(2′-hydroxy-6-methoxy-[1,1′-binaphthalen]-2-yl)carbamate (3c) White solid. Yield: 94%. 1H-NMR (400 MHz, CDCl3) δ 8.45 (d, J = 8.8 Hz, 1H), 7.98 (d, J = 8.8 Hz, 1H), 7.96 (d, J = 9.6 Hz, 1H), 7.92 (dd, J = 8.4, 1.2 Hz, 1H), 7.42–7.26 (m, 8H), 7.24 (d, J = 2.4 Hz, 1H), 7.08 (d, J = 9.2 Hz, 1H), 7.06 (d, J = 8.4 Hz, 1H), 6.99 (dd, J = 9.2, 2.4 Hz, 1H), 6.48 (s, 1H), 5.45 (s, 1H), 5.07 (d, J = 12.4 Hz, 1H), 5.03 (d, J = 12.4 Hz, 1H), 3.92 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 157.3, 153.8, 152.0, 135.8, 133.8, 133.3, 132.1, 131.2, 129.4, 129.0, 128.6, 128.4, 128.3, 128.3, 128.3, 127.4, 126.8, 124.2, 123.9, 120.8, 119.9, 118.0, 117.9, 113.0, 106.5, 67.1, 55.4. HRMS (ESI) calcd for [M + H] C29H24NO4, m/z: 450.1700, found: 450.1697.
Benzyl-(2′-hydroxy-6-phenyl-[1,1′-binaphthalen]-2-yl)carbamate (3d) Yellowish solid. Yield: 93%. 1H-NMR (400 MHz, CDCl3) δ 8.48 (d, J = 9.2 Hz, 1H), 8.05 (d, J = 2.0 Hz, 1H), 8.02 (d, J = 9.2 Hz, 1H), 7.92 (d, J = 8.8 Hz, 1H), 7.86 (d, J = 7.6 Hz, 1H), 7.61 (dd, J = 7.6, 2.4 Hz, 2H), 7.49 (dd, J = 8.8, 2.0 Hz, 1H), 7.41 (t, J = 7.6 Hz, 2H), 7.35–7.30 (m, 3H), 7.27–7.21 (m, 4H), 7.19–7.14 (m, 3H), 7.02 (d, J = 8.4 Hz, 1H), 6.52 (s, 1H), 5.46 (s, 1H), 4.98 (d, J = 12.4 Hz, 1H), 4.93 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.7, 152.2, 140.7, 138.0, 135.9, 135.7, 133.3, 132.2, 131.3, 131.1, 130.6, 129.5, 129.0, 128.6, 128.5, 128.4, 128.3, 127.5, 127.5, 127.3, 127.0, 126.1, 125.8, 124.2, 124.1, 120.2, 118.1, 117.1, 112.7, 67.2. HRMS (ESI) calcd for [M + H] C34H26NO3, m/z: 496.1907, found: 496.1909.
Benzyl-(6-bromo-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3e) Yellowish solid. Yield: 96%. 1H-NMR (400 MHz, CDCl3) δ 8.54 (d, J = 9.2 Hz, 1H), 8.09 (d, J = 2.0 Hz, 1H), 7.99 (d, J = 9.2 Hz, 1H), 7.94 (d, J = 8.8 Hz, 1H), 7.94 (d, J = 8.0 Hz, 1H), 7.43–7.25 (m, 9H), 7.02 (d, J = 9.2 Hz, 1H), 7.01 (d, J = 8.4 Hz, 1H), 6.57 (s, 1H), 5.60 (s, 1H), 5.05 (d, J = 12.0 Hz, 1H), 5.01 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.6, 152.1, 136.1, 135.6, 133.2, 131.9, 131.6, 131.5, 130.6, 130.2, 129.4, 129.2, 128.6, 128.5, 128.4, 128.3, 127.6, 127.1, 124.1, 124.0, 120.9, 119.3, 118.1, 117.6, 112.2, 67.3. HRMS (ESI) calcd for [M + H] C28H21BrNO3, m/z: 498.0700, found: 498.0700.
Benzyl-(6-fluoro-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3f) White solid. Yield: 92%. 1H-NMR (400 MHz, CDCl3) δ 8.44 (d, J = 9.2 Hz, 1H), 7.91 (dd, J = 9.2, 2.8 Hz, 2H), 7.85 (d, J = 8.0 Hz, 1H), 7.47 (dd, J = 9.2, 2.8 Hz, 1H), 7.35–7.18 (m, 8H), 7.07 (dd, J = 9.3, 5.6 Hz, 1H), 7.00 (td, J = 8.8, 2.8 Hz, 1H), 6.94 (d, J = 8.4 Hz, 1H), 6.45 (s, 1H), 5.46 (s, 1H), 4.99 (d, J = 12.4 Hz, 1H), 4.94 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 160.3 (d, J = 245.0 Hz), 153.8, 152.1, 135.7, 135.1 (d, J = 2.0 Hz), 133.2, 131.6 (d, J = 9.0 Hz), 131.4, 130.0, 129.4, 129.4, 129.4, 128.6, 128.5, 128.4, 128.3, 127.8 (d, J = 9.0 Hz), 127.6, 124.0 (d, J = 6.0 Hz), 121.2, 118.1, 117.9, 117.5 (d, J = 25.0 Hz) 112.5, 111.4 (d, J = 21.0 Hz), 67.2. 19F-NMR (376 MHz, CDCl3) δ −115.88. HRMS (ESI) calcd for [M + H] C28H21FNO3, m/z: 438.1500, found: 438.1500
Methyl-2-(((benzyloxy)carbonyl)amino)-2′-hydroxy-[1,1′-binaphthalene]-6-carboxylate (3g) Yellowish solid. Yield: 93%. 1H-NMR (400 MHz, CDCl3) δ 8.62 (d, J = 9.2 Hz, 1H), 8.41 (d, J = 2.0 Hz, 1H), 8.03 (d, J = 9.2 Hz, 1H), 8.01 (d, J = 8.8 Hz, 1H), 7.93 (d, J = 7.6 Hz, 1H), 7.75 (dd, J = 9.2, 2.0 Hz, 1H), 7.44 (d, J = 9.2 Hz, 1H), 7.40 (t, J = 8.0 Hz, 1H), 7.35–7.26 (m, 6H), 7.16 (d, J = 8.8 Hz, 1H), 6.98 (d, J = 8.0 Hz, 1H), 6.70 (s, 1H), 6.27 (s, 1H), 5.07 (d, J = 12.4 Hz, 1H), 5.03 (d, J = 12.4 Hz, 1H), 3.86 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 167.1, 153.4, 152.6, 138.0, 135.6, 135.5, 133.2, 131.5, 131.5, 131.2, 129.5, 129.4, 128.6, 128.5, 128.4, 128.4, 127.5, 126.5, 126.1, 125.4, 124.0, 123.9, 119.9, 118.5, 117.1, 112.0, 67.3, 52.3. HRMS (ESI) calcd for [M + H] C30H24NO5, m/z: 478.1649, found: 478.1649.
Benzyl-(2′-hydroxy-7-methoxy-[1,1′-binaphthalen]-2-yl)carbamate (3h) Yellowish solid. Yield: 97%. 1H-NMR (400 MHz, CDCl3) δ 8.39 (d, J = 8.8 Hz, 1H), 7.99 (d, J = 9.2 Hz, 1H), 7.98 (d, J = 9.2 Hz, 1H), 7.93 (d, J = 8.0 Hz, 1H), 7.83 (d, J = 8.8 Hz, 1H), 7.42–7.39 (m, 2H), 7.37–7.26 (m, 6H), 7.14–7.10 (m, 2H), 6.54 (s, 1H), 6.45 (d, J = 2.8 Hz, 1H), 5.48 (s, 1H), 5.08 (d, J = 12.4 Hz, 1H), 5.04 (d, J = 12.0 Hz, 1H), 3.51 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 158.9, 153.6, 152.1, 136.4, 135.8, 134.4, 133.0, 131.3, 130.0, 129.9, 129.4, 128.6, 128.5, 128.3, 128.3, 127.4, 126.3, 124.2, 124.0, 118.0, 117.5, 117.3, 115.9, 112.8, 103.9, 67.1, 55.1. HRMS (ESI) calcd for [M + H] C29H24NO4, m/z: 450.1700, found: 450.1699.
Benzyl-(2′-hydroxy-7-phenyl-[1,1′-binaphthalen]-2-yl)carbamate (3i) White solid. Yield: 98%. 1H-NMR (400 MHz, CDCl3) δ 8.60 (d, J = 8.8 Hz, 1H), 8.12 (d, J = 9.2 Hz, 1H), 8.04 (d, J = 8.4 Hz, 1H), 8.03 (d, J = 8.8 Hz, 1H), 7.97 (d, J = 8.0 Hz, 1H), 7.76 (d, J = 8.4 Hz, 1H), 7.48–7.30 (m, 14H), 7.19 (dd, J = 8.4, 2.0 Hz, 1H), 6.65 (s, 1H), 5.67 (s, 1H), 5.12 (d, J = 12.0 Hz, 1H), 5.08 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.7, 152.3, 141.0, 140.2, 136.3, 135.8, 133.4, 133.3, 131.4, 130.1, 130.0, 129.5, 128.9, 128.8, 128.6, 128.6, 128.4, 128.3, 127.5, 127.5, 127.5, 125.2, 124.2, 124.1, 123.1, 119.9, 118.1, 117.6, 112.7, 67.2. HRMS (ESI) calcd for [M + H] C34H26NO3, m/z: 496.1907, found: 496.1907.
Benzyl-(7-bromo-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3j) White solid. Yield: 95%. 1H-NMR (400 MHz, CDCl3) δ 8.54 (d, J = 9.2 Hz, 1H), 8.00 (dd, J = 8.8, 3.2 Hz, 2H), 7.94 (d, J = 8.0 Hz, 1H), 7.78 (d, J = 8.4 Hz, 1H), 7.52 (dd, J = 8.4, 2.0 Hz, 1H), 7.44–7.25 (m, 9H), 7.02 (d, J = 8.4 Hz, 1H), 6.54 (s, 1H), 5.47 (s, 1H), 5.06 (d, J = 12.4 Hz, 1H), 5.02 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.5, 152.1, 136.7, 135.6, 134.3, 133.0, 131.6, 130.2, 130.0, 129.5, 129.2, 128.7, 128.6, 128.6, 128.4, 128.3, 127.7, 127.1, 124.1, 123.9, 122.0, 120.0, 118.1, 116.4, 111.9, 67.3. HRMS (ESI) calcd for [M + H] C28H21BrNO3, m/z: 498.0700, found: 498.0699.
Benzyl-(7′-bromo-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3k) Yellowish solid. Yield: 94%. 1H-NMR (400 MHz, CDCl3) δ 8.52 (d, J = 8.8 Hz, 1H), 8.08 (d, J = 9.2 Hz, 1H), 7.95 (d, J = 8.0 Hz, 1H), 7.94 (d, J = 8.8 Hz, 1H), 7.77 (d, J = 8.8 Hz, 1H), 7.49–7.45 (m, 2H), 7.39 (d, J = 8.8 Hz, 1H), 7.36–7.28 (m, 6H), 7.19 (d, J = 2.4 Hz, 1H), 7.11 (d, J = 8.8 Hz, 1H), 6.48 (s, 1H), 5.49 (s, 1H), 5.09 (d, J = 12.4 Hz, 1H), 5.05 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.6, 153.0, 135.9, 135.6, 134.6, 132.8, 131.2, 130.9, 130.7, 130.1, 128.6, 128.4, 128.4, 128.4, 127.8, 127.6, 127.5, 126.1, 125.4, 124.9, 122.1, 119.9, 118.5, 116.2, 112.2, 67.3. HRMS (ESI) calcd for [M + H] C28H21BrNO3, m/z: 498.0700, found: 498.0701.
Methyl-2′-(((benzyloxy)carbonyl)amino)-2-hydroxy-[1,1′-binaphthalene]-6-carboxyl- ate (3l) Yellowish solid. Yield: 90%. 1H-NMR (400 MHz, Acetone-d6) δ 8.96 (s, 1H), 8.62 (d, J = 2.0 Hz, 1H), 8.44 (d, J = 9.2 Hz, 1H), 8.15 (d, J = 8.0 Hz, 1H), 8.06 (d, J = 8.8 Hz, 1H), 7.97 (d, J = 8.4 Hz, 1H), 7.78 (dd, J = 8.8, 2.0 Hz, 1H), 7.50 (d, J = 9.2 Hz, 1H), 7.43–7.39 (m, 2H), 7.29–7.20 (m, 6H), 7.08 (d, J = 8.4 Hz, 1H), 7.04 (d, J = 8.8 Hz, 1H), 5.04 (s, 2H), 3.89 (s, 3H). 13C-NMR (100 MHz, Acetone-d6) δ 171.7, 161.1, 158.9, 141.9, 141.7, 140.8, 138.4, 137.3, 136.4, 136.1, 134.1, 134.1, 133.5, 133.3, 133.3, 133.1, 133.1, 131.7, 131.2, 130.4, 130.1, 129.9, 129.5, 125.8, 124.9, 119.2, 71.3, 56.6. HRMS (ESI) calcd for [M + H] C30H24NO5, m/z: 478.1649, found: 478.1651.
Dimethyl-2-(((benzyloxy)carbonyl)amino)-2′-hydroxy-[1,1′-binaphthalene]-6,6′-dicarboxylate (3m) Yellowish solid. Yield: 81%. 1H-NMR (400 MHz, CDCl3) δ 8.58 (d, J = 1.6 Hz, 1H), 8.56 (d, J = 8.8 Hz, 1H), 8.39 (d, J = 2.0 Hz, 1H), 8.05 (d, J = 8.8 Hz, 1H), 8.01 (d, J = 9.2 Hz, 1H), 7.78 (dd, J = 8.8, 1.6 Hz, 1H), 7.72 (dd, J = 8.8, 1.6 Hz, 1H), 7.43 (d, J = 9.2 Hz, 1H), 7.31–7.26 (m, 3H), 7.24–7.21 (m, 2H), 7.05 (d, J = 8.8 Hz, 1H), 6.94 (d, J = 8.8 Hz, 1H), 6.60 (s, 1H), 6.34 (s, 1H), 5.02 (s, 2H), 3.91 (s, 3H), 3.84 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 167.1, 167.0, 154.6, 153.3, 138.1, 135.7, 135.4, 135.3, 133.0, 131.8, 131.6, 131.2, 129.6, 128.6, 128.5, 128.4, 128.4, 127.0, 126.6, 126.3, 125.6, 125.1, 124.0, 120.0, 119.3, 116.3, 112.3, 67.4, 52.3, 52.2. HRMS (ESI) calcd for [M + H] C32H26NO7, m/z: 536.1704, found: 536.1706.
Benzyl-(6,6′-dibromo-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3n) White solid. Yield: 90%. 1H-NMR (400 MHz, Acetone-d6) δ 8.74 (s, 1H), 8.48 (d, J = 9.2 Hz, 1H), 8.19 (d, J = 2.4 Hz, 1H), 8.12 (d, J = 2.0 Hz, 1H), 8.04 (d, J = 9.2 Hz, 1H), 7.97 (d, J = 8.8 Hz, 1H), 7.44 (d, J = 8.8 Hz, 1H), 7.43 (s, 1H), 7.38 (dd, J = 9.2, 2.4 Hz, 1H), 7.34 (dd, J = 8.8, 2.0 Hz, 1H), 7.32–7.26 (m, 3H), 7.23–7.21 (m, 2H), 6.99 (d, J = 8.8 Hz, 1H), 6.86 (d, J = 8.8 Hz, 1H), 5.06 (d, J = 12.4 Hz, 1H), 5.03 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, Acetone-d6) δ 154.1, 153.6, 136.6, 136.2, 132.5, 132.0, 131.8, 130.3, 130.2, 130.0, 130.0, 129.9, 129.6, 128.3, 128.1, 127.9, 127.9, 127.5, 126.1, 121.7, 120.5, 120.0, 118.0, 116.5, 113.4, 66.2. HRMS (ESI) calcd for [M + H] C28H20Br2NO3, m/z: 575.9805, found: 575.9808.
Benzyl-(7,7′-dibromo-2′-hydroxy-[1,1′-binaphthalen]-2-yl)carbamate (3o) Yellowish solid. Yield: 85%. 1H-NMR (400 MHz, CDCl3) δ 8.53 (dd, J = 9.2, 2.0 Hz, 1H), 8.02 (d, J = 9.2 Hz, 1H), 7.94 (d, J = 8.8 Hz, 1H), 7.79 (d, J = 8.8 Hz, 1H), 7.78 (d, J = 8.4 Hz, 1H), 7.53 (dd, J = 8.8, 2.0 Hz, 1H), 7.48 (dd, J = 8.8, 2.0 Hz, 1H), 7.38–7.32 (m, 4H), 7.29–7.26 (m, 2H), 7.22 (d, J = 2.0 Hz, 1H), 7.13 (d, J = 1.6 Hz, 1H), 6.44 (s, 1H), 5.49 (s, 1H), 5.08 (d, J = 12.4 Hz, 1H), 5.04 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.4, 153.0, 136.8, 135.5, 134.4, 134.1, 131.6, 130.5, 130.3, 130.1, 129.2, 128.9, 128.6, 128.5, 128.4, 127.9, 127.7, 126.8, 125.8, 122.4, 122.2, 120.1, 118.6, 115.4, 111.3, 67.4. HRMS (ESI) calcd for [M + H] C28H20Br2NO3, m/z: 575.9805, found: 575.9804.
Benzyl-(2′-hydroxy-7,7′-diphenyl-[1,1′-binaphthalen]-2-yl)carbamate (3p) White solid. Yield: 83%. 1H-NMR (400 MHz, CDCl3) δ 8.58 (d, J = 8.8 Hz, 1H), 8.12 (d, J = 8.8 Hz, 1H), 8.05–8.01 (m, 3H), 7.74 (dd, J = 8.4, 2.0 Hz, 1H), 7.69 (dd, J = 8.4, 2.0 Hz, 1H), 7.47–7.41 (m, 7H), 7.39–7.27 (m, 11H), 6.63 (s, 1H), 5.38 (s, 1H), 5.12 (d, J = 12.4 Hz, 1H), 5.09 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ153.7, 152.5, 141.0, 141.0, 140.4, 140.3, 136.3, 135.7, 133.5, 133.2, 131.2, 130.2, 130.1, 129.1, 129.0, 128.8, 128.8, 128.7, 128.6, 128.4, 128.3, 127.5, 127.5, 127.4, 127.4, 125.2, 123.9, 122.9, 121.9, 120.0, 118.0, 117.2, 112.9, 67.2. HRMS (ESI) calcd for [M + H] C40H30NO3, m/z: 572.2220, found: 572.2223.
Methyl-2′-((benzyl carbonyl)amino)-2-hydroxy-6′-methoxy-[1,1′-binaphthalene]-6-carboxylate (3q) Yellowish solid. Yield: 88%. 1H-NMR (400 MHz, CDCl3) δ 8.62 (d, J = 2.0 Hz, 1H), 8.37 (d, J = 8.8 Hz, 1H), 8.04 (d, J = 8.8 Hz, 1H), 7.93 (d, J = 9.2 Hz, 1H), 7.81 (dd, J = 9.2, 2.0 Hz, 1H), 7.42 (d, J = 8.8 Hz, 1H), 7.34–7.20 (m, 6H), 7.05 (d, J = 8.8 Hz, 1H), 7.00 (d, J = 9.2 Hz, 1H), 6.95 (dd, J = 9.2, 2.8 Hz, 1H), 6.45 (s, 1H), 5.99 (s, 1H), 5.05 (d, J = 12.0 Hz, 1H), 5.00 (d, J = 12.0 Hz, 1H), 3.94 (s, 3H), 3.88 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 167.2, 157.3, 154.3, 153.9, 135.8, 135.7, 133.7, 132.5, 132.2, 131.5, 129.1, 128.5, 128.3, 128.3, 128.2, 128.1, 126.8, 126.6, 125.4, 124.3, 121.0, 120.0, 119.1, 117.7, 113.4, 106.5, 67.1, 55.3, 52.2. HRMS (ESI) calcd for [M + H] C31H26NO6, m/z: 508.1755, found: 508.1759.
Benzyl-(6′-fluoro-2′-hydroxy-6-methoxy-[1,1′-binaphthalen]-2-yl)carbamate (3r) White solid. Yield: 85%. 1H-NMR (400 MHz, CDCl3) δ 8.40 (d, J = 8.8 Hz, 1H), 7.94 (d, J = 8.8 Hz, 1H), 7.89 (d, J = 8.8 Hz, 1H), 7.54 (dd, J = 9.5, 2.4 Hz, 1H), 7.40 (d, J = 8.8 Hz, 1H), 7.37-7.31 (m, 3H), 7.28–7.23 (m, 3H), 7.10–6.98 (m, 4H), 6.46 (s, 1H), 5.60 (s, 1H), 5.07 (d, J = 12.4 Hz, 1H), 5.01 (d, J = 12.4 Hz, 1H), 3.91 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 159.5 (d, J = 243.0 Hz), 157.3, 153.9, 151.5 (d, J = 2.0 Hz), 135.8, 133.6, 132.2, 130.2 (d, J = 5.0 Hz), 130.2, 129.9 (d, J = 9.0 Hz), 129.1, 128.6, 128.4, 128.3, 128.1, 120.9, 126.7, 126.5 (d, J = 8.0 Hz), 120.0, 119.3, 118.0, 117.4 (d, J = 25.0 Hz), 113.5, 111.6 (d, J = 21.0 Hz), 106.5, 67.2, 55.4. 19F-NMR (376 MHz, CDCl3) δ −118.10. HRMS (ESI) calcd for [M + H] C29H23FNO4, m/z: 468.1606, found: 468.1607.
Benzyl-(6′-fluoro-2′-hydroxy-7-phenyl-[1,1′-binaphthalen]-2-yl)carbamate (3s) Yellowish solid. Yield: 84%. 1H-NMR (400 MHz, CDCl3) δ 8.55 (d, J = 9.2 Hz, 1H), 8.10 (d, J = 9.2 Hz, 1H), 8.02 (d, J = 8.8 Hz, 1H), 7.93 (d, J = 8.8 Hz, 1H), 7.75 (dd, J = 8.4, 2.0 Hz, 1H), 7.57 (dd, J = 9.6, 2.4 Hz, 1H), 7.47–7.44 (m, 3H), 7.41–7.28 (m, 9H), 7.16–7.08 (m, 2H), 6.57 (s, 1H), 5.55 (s, 1H), 5.12 (d, J = 12.0 Hz, 1H), 5.06 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 159.6 (d, J = 242.0 Hz), 153.7, 151.7 (d, J = 2.0 Hz), 140.9, 140.3, 136.2, 135.7, 133.2, 130.5 (d, J = 5.0 Hz), 130.2, 130.1, 130.1, 130.0 (d, J = 10.0 Hz), 129.0, 128.8, 128.6, 128.5, 128.4, 127.5, 127.5, 126.5 (d, J = 8.0 Hz), 125.3, 122.9, 120.0, 119.4, 117.5 (d, J = 25.0 Hz), 117.3, 113.0, 111.8 (d, J = 21.0 Hz), 67.3. 19F-NMR (376 MHz, CDCl3) δ −117.88. HRMS (ESI) calcd for [M + H] C34H25FNO3, m/z: 514.1813, found: 514.1814.
Methyl-2-(((benzyloxy)carbonyl)amino)-7′-bromo-2′-hydroxy-[1,1′-binaphthalene]- 6-carboxylate (3t) Yellowish solid. Yield: 86%. 1H-NMR (400 MHz, CDCl3) δ 8.61 (d, J = 9.2 Hz, 1H), 8.41 (d, J = 2.0 Hz, 1H), 8.04 (d, J = 9.2 Hz, 1H), 7.95 (d, J = 8.8 Hz, 1H), 7.77 (d, J = 8.8 Hz, 1H), 7.76 (dd, J = 8.8, 2.0 Hz, 1H), 7.46 (dd, J = 8.4, 2.0 Hz, 1H), 7.42 (d, J = 8.8 Hz, 1H), 7.37–7.27 (m, 5H), 7.09 (d, J = 8.8 Hz, 1H), 7.09 (d, J = 2.0 Hz, 1H), 6.60 (s, 1H), 6.29 (s, 1H), 5.09 (d, J = 12.0 Hz, 1H), 5.05 (d, J = 12.0 Hz, 1H) 3.87 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 167.1, 153.4, 153.3, 138.0, 135.4, 135.3, 134.5, 131.9, 131.5, 131.3, 130.2, 129.6, 128.6, 128.5, 128.4, 127.8, 127.5, 126.7, 126.3, 125.8, 125.1, 122.2, 119.9, 118.9, 116.1, 111.4, 67.4, 52.3. HRMS (ESI) calcd for [M + H] C30H23BrNO5, m/z: 556.0754, found: 556.0756.
Benzyl-(7-bromo-2′-hydroxy-7′-phenyl-[1,1′-binaphthalen]-2-yl)carbamate (3u) Yellowish solid. Yield: 88%. 1H-NMR (400 MHz, CDCl3) δ 8.50 (d, J = 9.2 Hz, 1H), 7.96 (d, J = 9.2 Hz, 2H), 7.95 (d, J = 8.4 Hz, 1H), 7.73 (d, J = 8.4 Hz, 1H), 7.62 (dd, J = 8.4, 1.6 Hz, 1H), 7.47 (dd, J = 8.8, 2.0 Hz, 1H), 7.38–7.18 (m, 12H), 7.13 (d, J = 2.0 Hz, 1H), 6.52 (s, 1H), 5.35 (s, 1H), 5.02 (d, J = 12.0 Hz, 1H), 4.99 (d, J = 12.4 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.6, 152.5, 140.9, 140.5, 136.8, 135.6, 134.3, 133.3, 131.4, 130.3, 130.0, 129.3, 129.2, 128.8, 128.8, 128.7, 128.6, 128.4, 128.3, 127.5, 127.5, 127.1, 124.0, 122.1, 121.6, 120.1, 118.1, 116.2, 112.2, 67.3. HRMS (ESI) calcd for [M + H] C34H25BrNO3, m/z: 574.1013, found: 574.1016.
Benzyl-(1-(10-hydroxyphenanthren-9-yl)naphthalen-2-yl)carbamate (3v) Yellowish solid. Yield: 91%. 1H-NMR (400 MHz, CDCl3) δ 8.85 (d, J = 8.0 Hz, 1H), 8.79 (d, J = 8.0 Hz, 1H), 8.63 (d, J = 8.8 Hz, 1H), 8.51 (dd, J = 8.0, 1.6 Hz, 1H), 8.13 (d, J = 9.2 Hz, 1H), 7.98 (d, J = 8.0 Hz, 1H), 7.87–7.83 (m, 1H), 7.79–7.75 (m, 1H), 7.59–7.55 (m, 1H), 7.49–7.45 (m, 1H), 7.41–7.37 (m, 1H), 7.33–7.28 (m, 5H), 7.27–7.22 (m, 2H), 7.10 (dd, J = 8.4, 1.6 Hz, 1H), 6.68 (s, 1H), 5.61 (s, 1H), 5.06 (d, J = 12.4 Hz, 1H), 5.02 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.6, 148.4, 136.3, 135.7, 133.0, 132.0, 131.6, 130.9, 130.5, 128.5, 128.3, 128.3, 128.3, 128.0, 127.7, 127.5, 127.0, 126.9, 125.3, 125.2, 125.1, 124.8, 124.8, 123.6, 123.0, 122.8, 119.7, 116.7, 108.9, 67.1. HRMS (ESI) calcd for [M + H] C32H24NO3, m/z: 470.1751, found: 470.1753.
Benzyl-(1-(10-hydroxyphenanthren-9-yl)-7-methoxynaphthalen-2-yl)carbamate (3w) White solid. Yield: 90%. 1H-NMR (400 MHz, CDCl3) δ 8.79 (dd, J = 8.4, 1.2 Hz, 1H), 8.72 (dd, J = 8.4, 1.6 Hz, 1H), 8.44 (dd, J = 8.4, 1.6 Hz, 1H), 8.41 (d, J = 9.2 Hz, 1H), 7.99 (d, J = 9.2 Hz, 1H), 7.83 (d, J = 9.2 Hz, 1H), 7.82–7.78 (m, 1H), 7.73–7.69 (m, 1H), 7.54–7.50 (m, 1H), 7.36–7.32 (m, 1H), 7.27–7.23 (m, 3H), 7.20–7.17 (m, 2H), 7.10–7.06 (m, 2H), 6.54 (s, 1H), 6.48 (d, J = 2.4 Hz, 1H), 5.47 (s, 1H), 5.03 (d, J = 12.4 Hz, 1H), 5.00 (d, J = 12.4 Hz, 1H), 3.43 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 159.0, 153.6, 148.2, 136.8, 135.7, 134.4, 131.9, 131.3, 130.3, 129.9, 128.5, 128.3, 128.3, 128.0, 127.7, 127.0, 126.8, 126.3, 125.0, 124.8, 124.7, 123.6, 123.0, 122.7, 117.5, 117.2, 115.3, 109.0, 103.8, 67.1, 55.1. HRMS (ESI) calcd for [M + H] C33H26NO4, m/z: 500.1857, found: 500.1859.
Methyl-6-(((benzyloxy)carbonyl)amino)-5-(10-hydroxyphenanthren-9-yl)-2-naphth- oate (3x) Yellowish solid. Yield: 92%. 1H-NMR (400 MHz, CDCl3) δ 8.79 (d, J = 7.2 Hz, 1H), 8.73 (dd, J = 8.4, 1.2 Hz, 1H), 8.68 (d, J = 8.8 Hz, 1H), 8.57 (d, J = 2.0 Hz, 1H), 8.46 (dd, J = 8.0, 1.6 Hz, 1H), 8.13 (d, J = 9.6 Hz, 1H), 7.84–7.79 (m, 1H), 7.77 (dd, J = 8.8, 1.6 Hz, 1H), 7.74–7.70 (m, 1H), 7.55–7.50 (m, 1H), 7.35–7.30 (m, 1H), 7.28–7.24 (m, 3H), 7.23–7.19 (m, 3H), 6.96 (dd, J = 8.4, 1.6 Hz, 1H), 6.70 (s, 1H), 5.65 (s, 1H), 5.04 (d, J = 12.4 Hz, 1H), 5.01 (d, J = 12.4 Hz, 1H), 3.90 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 167.0, 153.3, 148.5, 138.5, 135.4, 135.4, 132.0, 131.9, 131.3, 131.3, 129.7, 128.5, 128.4, 128.4, 128.2, 127.8, 127.1, 126.9, 126.8, 126.6, 125.4, 125.0, 124.9, 124.6, 123.6, 123.1, 122.7, 119.9, 116.3, 108.1, 67.3, 52.3. HRMS (ESI) calcd for [M + H] C34H26NO5, m/z: 528.1806, found: 528.1807.
Benzyl-(1-(2-hydroxyphenyl)naphthalen-2-yl)carbamate (3y) Yellowish solid. Yield: 75%. 1H-NMR (400 MHz, CDCl3) δ 8.36 (d, J = 8.4 Hz, 1H), 7.90–7.86 (m, 2H), 7.59–7.53 (m, 1H), 7.42–7.38 (m, 2H), 7.33–7.24 (m, 9H), 6.42 (s, 1H), 5.36 (s, 1H), 5.08 (s, 2H). 13C-NMR (100 MHz, CDCl3) δ 153.5, 151.3, 137.4, 135.8, 133.0, 132.0, 130.8, 130.1, 129.2, 128.5, 128.3, 128.3, 128.3, 127.3, 124.4, 124.1, 123.8, 122.7, 120.5, 117.8, 115.3, 67.0. HRMS (ESI) calcd for [M + H] C24H20NO3, m/z: 370.1438, found: 370.1435.
Benzyl-(1-(5-bromo-2-hydroxyphenyl)naphthalen-2-yl)carbamate (3z) White solid. Yield: 72%. 1H-NMR (400 MHz, CDCl3) δ 8.20 (d, J = 8.8 Hz, 1H), 7.87 (d, J = 9.2 Hz, 1H), 7.80 (d, J = 7.6 Hz, 1H), 7.45–7.29 (m, 8H), 7.28–7.22 (m, 2H), 6.92 (d, J = 8.4 Hz, 1H), 6.58 (s, 1H), 5.47 (s, 1H), 5.05 (d, J = 12.4 Hz, 1H), 5.00 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.8, 153.2, 135.7, 134.4, 134.0, 133.5, 132.5, 130.8, 130.2, 128.7, 128.5, 128.4, 128.2, 127.4, 125.4, 124.9, 123.2, 120.3, 119.7, 118.5, 113.2, 67.4. HRMS (ESI) calcd for [M + H] C24H19BrNO3, m/z: 448.0543, found: 448.0540.
Benzyl-(1-(3-fluoro-6-hydroxy-2,4-dimethylphenyl)naphthalen-2-yl)carbamate (3aa) Yellowish solid. Yield: 72%. 1H-NMR (400 MHz, CDCl3) δ 7.96 (d, J = 7.6 Hz, 1H), 7.91-7.87 (m, 2H), 7.42–7.14 (m, 9H), 6.16 (s, 1H), 5.45 (s, 1H), 5.05 (d, J = 12.0 Hz, 1H), 5.01 (d, J = 12.0 Hz, 1H), 2.41 (d, J = 2.4 Hz, 3H), 1.86 (d, J = 2.8 Hz, 3H). 13C-NMR (100 MHz, CDCl3) δ 156.8 (d, J = 240.3 Hz), 153.8, 151.4, 135.8, 132.6, 131.0, 129.3, 128.5, 128.5, 128.3, 128.2, 127.5, 126.1 (d, J = 19.0 Hz), 125.8 (d, J = 17.9 Hz), 123.9, 123.5, 122.1, 121.0, 117.9, 113.5, 113.4, 67.0, 15.1 (d, J = 3.7 Hz), 11.9 (d, J = 4.4 Hz). 19F-NMR (375 MHz, CDCl3) δ −123.4. HRMS (ESI) calcd for [M + H] C26H23FNO3, m/z: 416.1657, found: 416.1658.
Benzyl-(2-(2-hydroxynaphthalen-1-yl)-3,5-dimethylphenyl)carbamate (3ab) White solid. Yield: 62%. 1H-NMR (400 MHz, CDCl3) δ 8.03 (s, 1H), 7.90 (d, J = 8.8 Hz, 1H), 7.88–7.85 (m, 1H), 7.41–7.30 (m, 6H), 7.27–7.15 (m, 3H), 7.02 (d, J = 1.6 Hz, 1H), 7.02 (s, 1H), 6.25 (s, 1H), 5.04 (s, 2H), 2.47 (s, 3H), 1.90 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 153.5, 151.4, 139.9, 139.3, 137.3, 135.9, 132.7, 130.8, 129.4, 128.5, 128.5, 128.3, 128.2, 127.3, 126.7, 123.8, 123.7, 118.8, 118.3, 117.7, 114.1, 66.9, 21.7, 19.8. HRMS (ESI) calcd for [M + H] C26H24NO3, m/z: 398.1751, found: 398.1750.
Benzyl-(2-(2-hydroxynaphthalen-1-yl)phenyl)carbamate (3ac) Yellowish solid. Yield: 68%. 1H-NMR (400 MHz, CDCl3) δ 8.28 (d, J = 9.2 Hz, 1H), 7.84 (d, J = 9.2 Hz, 1H), 7.76 (d, J = 8.0 Hz, 1H), 7.35–7.21 (m, 9H), 7.08–6.99 (m, 3H), 6.99 (s, 1H), 5.05 (s, 1H), 5.09 (d, J = 12.0 Hz, 1H), 5.04 (d, J = 12.0 Hz, 1H). 13C-NMR (100 MHz, CDCl3) δ 153.9, 153.7, 135.9, 134.6, 132.7, 131.7, 130.7, 130.6, 129.8, 128.6, 128.4, 128.4, 128.2, 127.1, 125.1, 125.1, 121.5, 120.7, 120.0, 119.8, 116.6, 67.2. HRMS (ESI) calcd for [M + H] C24H20NO3, m/z: 370.1438, found: 370.1438.
Benzyl-(1-(2-hydroxy-2′-iodo-6,6′-dimethyl-[1,1′-biphenyl]-3-yl)naphthalen-2-yl)- carbamate (3ad) Yellowish solid. 72% yield, dr = 1.2/1. First diastereomer: 1H-NMR (400 MHz, CDCl3) δ 8.48 (d, J = 9.2 Hz, 1H), 7.96 (d, J = 9.2 Hz, 1H), 7.90–7.88 (m, 1H), 7.80 (d, J = 8.0 Hz, 1H), 7.50–7.30 (m, 9H), 7.19–7.10 (m, 2H), 7.02 (t, J = 8.0 Hz, 1H), 6.94 (s, 1H), 5.16–5.09 (m, 2H), 4.62 (s, 1H), 2.20 (s, 3H), 2.06 (s, 3H). 13C-NMR (100 MHz, CDCl3) δ 153.5, 149.9, 140.2, 138.8, 138.2, 136.9, 135.8, 134.6, 133.1, 131.6, 131.6, 130.5, 130.1, 129.8, 129.4, 128.6, 128.5, 128.3, 128.1, 126.7, 125.1, 124.7, 123.2, 120.2, 119.3, 118.6, 102.2, 67.1, 21.4, 19.6. HRMS (ESI) calcd for [M + H] C32H27INO3, m/z: 600.1030, found: 600.1031. Second diastereomer: 1H-NMR (400 MHz, CDCl3) δ 8.37 (d, J = 8.4 Hz, 1H), 7.95 (d, J = 9.2 Hz, 1H), 7.89–7.87 (m, 1H), 7.84 (d, J = 8.0 Hz, 1H), 7.66–7.62 (m, 1H), 7.48–7.31 (m, 8H), 7.19 (d, J = 7.6 Hz, 1H), 7.11 (d, J = 7.6 Hz, 1H), 7.02 (t, J = 8.0 Hz, 1H), 6.79 (s, 1H), 5.21–5.14 (m, 2H), 4.65 (s, 1H), 2.08 (s, 6H). 13C-NMR (100 MHz, CDCl3) δ 153.5, 149.9, 140.3, 138.8, 138.3, 137.0, 136.0, 134.1, 132.8, 131.6, 131.4, 130.7, 130.2, 129.8, 129.5, 128.6, 128.3, 128.2, 126.9, 125.7, 124.9, 123.2, 121.0, 119.6, 118.7, 115.6, 101.9, 67.1, 21.3, 19.6. HRMS (ESI) calcd for [M + H] C32H27INO3, m/z: 600.1030, found: 600.1030.
N2-methyl-[1,1′-binaphthalene]-2,2′-diamine (5a) Yellowish solid. Yield: 52%. 1H-NMR (400 MHz, CDCl3) δ 7.92 (d, J = 8.8 Hz, 1H), 7.83–7.79 (m, 3H), 7.27–7.24 (m, 2H), 7.22–7.17 (m, 3H), 7.15 (d, J = 8.4 Hz, 1H), 7.04 (d, J = 8.4 Hz, 1H), 7.02 (dd, J = 9.2, 2.8 Hz, 1H), 3.85 (brs, 1H), 2.85 (s, 3H), 2.55 (brs, 2H). 13C-NMR (100 MHz, CDCl3) δ 144.9, 142.5, 133.5, 133.0, 129.3, 129.1, 128.1, 127.8, 127.7, 127.2, 126.4, 126.3, 123.6, 123.2, 122.0, 121.4, 117.9, 113.0, 112.1, 111.5, 30.7. HRMS (ESI) calcd for [M + H] C21H19N2, m/z: 299.1543, found: 299.1544.
N2-phenyl-[1,1′-binaphthalene]-2,2′-diamine (5b) White solid. Yield: 50%. 1H-NMR (400 MHz, CDCl3) δ 7.93–7.84 (m, 4H), 7.76 (d, J = 9.2 Hz, 1H), 7.48–7.06 (m, 11H), 6.98 (t, J = 7.6 Hz, 1H), 5.67 (brs, 1H), 3.22 (brs, 2H). 13C-NMR (100 MHz, CDCl3) δ 142.9, 142.8, 140.2, 134.0, 133.8, 129.8, 129.5, 129.3, 129.2, 128.5, 128.3, 128.2, 127.1, 126.9, 124.6, 123.9, 123.4, 122.6, 122.0, 119.8, 118.4, 118.0, 116.9, 112.1. HRMS (ESI) calcd for [M + H] C26H21N2, m/z: 361.1699, found: 361.1699.
N2-benzyl-[1,1′-binaphthalene]-2,2′-diamine (5c) Yellow solid. Yield: 50%. 1H-NMR (400 MHz, CDCl3) δ 7.86–7.78 (m, 4H), 7.31–7.18 (m, 11H), 7.15–7.13 (m, 1H), 7.08–7.05 (m, 1H), 4.44 (s, 2H), 3.92 (brs, 3H). 13C-NMR (100 MHz, CDCl3) δ 143.9, 143.0, 139.8, 134.0, 133.6, 129.7, 129.6, 128.6, 128.5, 128.2, 128.2, 127.8, 126.9, 126.9, 126.9, 126.8, 124.2, 123.8, 122.5, 122.0, 118.4, 114.4, 112.5, 112.3, 47.7. HRMS (ESI) calcd for [M + H] C27H23N2, m/z: 375.1856, found: 375.1855.
N2-benzyl-6′-fluoro-[1,1′-binaphthalene]-2,2′-diamine (5d) Light brown solid. Yield: 56%. 1H-NMR (400 MHz, CDCl3) δ 7.84 (d, J = 9.2 Hz, 1H), 7.82–7.79 (m, 1H), 7.78 (d, J = 8.8 Hz, 1H), 7.46 (dd, J = 10.0, 2.8 Hz, 1H), 7.28–7.20 (m, 9H), 7.12–7.09 (m, 1H), 7.05–7.00 (m, 2H), 4.44 (s, 2H), 4.19 (brs, 1H), 3.72 (brs, 2H). 13C-NMR (100 MHz, CDCl3) δ 158.9 (d, J = 241.0 Hz), 143.9, 142.4 (d, J = 2.0 Hz), 139.8, 133.5, 130.9, 129.7, 128.9 (d, J = 8.0 Hz), 128.8 (d, J = 5.0 Hz), 128.5, 128.2, 127.7, 127.0, 126.9, 126.8, 126.5 (d, J = 8.0 Hz), 123.6, 122.1, 119.6, 116.7 (d, J = 24.0 Hz), 114.3, 112.7, 112.1, 111.3 (d, J = 20.0 Hz), 47.7. 19F-NMR (376 MHz, CDCl3) δ −120.50. HRMS (ESI) calcd for [M + H] C27H22FN2, m/z: 393.1762, found: 393.1762

4. Conclusions

In summary, a copper-catalyzed domino reaction toward NOBIN and BINAM derivatives has been established employing diaryliodonium salts as arylation reagents. The results from the control experiments substantiated that the copper catalyst played a key role in improving the yield during the arylation process. This reaction consisting of facile O-/N-arylation and [3,3]-sigmatropic rearrangement sequence proceeds under mild conditions and displayed good substrate generality and excellent efficiency. In addition, a group of biaryl amino alcohols (including a diaxial structure) and BINAM derivatives were synthesized in moderate or good yields under identical conditions.

Supplementary Materials

The following are available online: Table S1. Solvent and catalysts screenings for the reaction with N-naphthylhydroxylamine; Table S2. Base and loading screenings for the reaction with N-naphthylhydroxylamine; III. Supplementary Experimental Procedures; V. Copies of NMR spectra.

Author Contributions

B.T. and S.-H.X. conceived and directed the project. J.-W.Z. performed experiments and prepared the Supplementary Materials. S.L. helped with characterizing some new compounds. J.-W.Z., S.-H.X. and B.T. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Shenzhen Science and Technology Innovation Committee (ZDSYS20190902093215877).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Kočovský, P.; Vyskočil, Š.; Smrčina, M. Non-Symmetrically Substituted 1,1‘-Binaphthyls in Enantioselective Catalysis. Chem. Rev. 2003, 103, 3213–3246. [Google Scholar] [CrossRef]
  2. Ding, K.; Li, X.; Ji, B.; Guo, H.; Kitamura, M. Ten Years of Research on NOBIN Chemistry. Curr. Org. Synth. 2005, 2, 499–545. [Google Scholar] [CrossRef]
  3. 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] [CrossRef]
  4. Zhou, Q.-L. (Ed.) Privileged Chiral Ligands and Catalysts; Wiley VCH: Weinheim, Germany, 2011; pp. 1–462. ISBN 978-3-527-32704-1. [Google Scholar]
  5. Carreira, E.M.; Lee, W.; Singer, R.A. Catalytic, Enantioselective Acetone Aldol Additions with 2-Methoxypropene. J. Am. Chem. Soc. 1995, 117, 3649–3650. [Google Scholar] [CrossRef]
  6. Van Veldhuizen, J.J.; Garber, S.B.; Kingsbury, J.S.; Hoveyda, A.H. A Recyclable Chiral Ru Catalyst for Enantioselective Olefin Metathesis. Efficient Catalytic Asymmetric Ring-Opening/Cross Metathesis in Air. J. Am. Chem. Soc. 2002, 124, 4954–4955. [Google Scholar] [CrossRef]
  7. Yan, Y.; Zhang, X. A Hybrid Phosphorus Ligand for Highly Enantioselective Asymmetric Hydroformylation. J. Am. Chem. Soc. 2006, 128, 7198–7202. [Google Scholar] [CrossRef]
  8. Uraguchi, D.; Kinoshita, N.; Ooi, T. Catalytic Asymmetric Protonation of α-Amino Acid-Derived Ketene Disilyl Acetals UsingP-Spiro Diaminodioxaphosphonium Barfates as Chiral Proton. J. Am. Chem. Soc. 2010, 132, 12240–12242. [Google Scholar] [CrossRef] [PubMed]
  9. Tan, B.; Candeias, N.R.; Barbas, C.F., III. Construction of Bispirooxindoles Containing Three Quaternary Stereocentres in a Cascade Using a Single Multifunctional Organocatalyst. Nat. Chem. 2011, 3, 473–477. [Google Scholar] [CrossRef] [PubMed]
  10. Vallavoju, N.; Selvakumar, S.; Jockusch, S.; Sibi, M.P.; Sivaguru, J. Enantioselective Organo-Photocatalysis Mediated by Atropisomeric Thiourea Derivatives. Angew. Chem. Int. Ed. 2014, 53, 5604–5608. [Google Scholar] [CrossRef] [PubMed]
  11. Lipshutz, B.H.; Buzard, D.; Olsson, C.; Noson, K. A modular route to nonracemic cyclo-NOBINs. Preparation of the parent ligand for homo- and heterogeneous catalysis. Tetrahedron 2004, 60, 4443–4449. [Google Scholar] [CrossRef]
  12. Li, Q.; Green, L.; Venkataraman, N.; Shiyanovskaya, I.; Khan, A.; Urbas, A.; Doane, J.W. Reversible Photoswitchable Axially Chiral Dopants with High Helical Twisting Power. J. Am. Chem. Soc. 2007, 129, 12908–12909. [Google Scholar] [CrossRef] [PubMed]
  13. Ritter, N.; Senkovska, I.; Kaskel, S.; Weber, J. Towards Chiral Microporous Soluble Polymers-Binaphthalene-Based Polyi-mides. Macromol. Rapid Commun. 2011, 32, 438–443. [Google Scholar] [CrossRef] [PubMed]
  14. Singer, R.A.; Buchwald, S.L. Preparation of 2-Amino-2′-hydroxy-1,1’-binaphthyl and N-arylated 2-Amino-1,1’-binaphthyl Derivatives via Palladium-Catalyzed Amination. Tetrahedron Lett. 1999, 40, 1095–1098. [Google Scholar] [CrossRef]
  15. Patel, D.; Breitbach, Z.S.; Woods, R.M.; Lim, Y.; Wang, A.; Foss, F.W.; Armstrong, D.W. Gram Scale Conversion ofR-BINAM toR-NOBIN. J. Org. Chem. 2016, 81, 1295–1299. [Google Scholar] [CrossRef] [Green Version]
  16. Chang, X.; Zhang, Q.; Guo, C. Switchable Smiles Rearrangement for Enantioselective O-Aryl Amination. Org. Lett. 2019, 21, 4915–4918. [Google Scholar] [CrossRef]
  17. Körber, K.; Tang, W.; Hu, X.; Zhang, X. A Practical Synthesis of 2-Amino-2′-hydroxy-1,1′-binaphthyl (NOBIN). Tetrahedron Lett. 2002, 43, 7163–7165. [Google Scholar] [CrossRef]
  18. Smrcina, M.; Lorenc, M.; Hanuš, V.; Kocovsky, P. A Facile Synthesis of 2-Amino-2?-hydroxy-1,1?-binaphthyl and 2,2?-Diamino-1,1?-binaphthyl by Oxidative Coupling Using Copper(II) Chloride. Synlett 1991, 1991, 231–232. [Google Scholar] [CrossRef]
  19. 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] [CrossRef]
  20. 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] [CrossRef]
  21. Smrcina, M.; Vyskocil, S.; Maca, B.; Polasek, M.; Claxton, T.A.; Abbott, A.P.; Kocovsky, P. Selective Cross-Coupling of 2-Naphthol and 2-Naphthylamine Derivatives. A Facile Synthesis of 2,2’,3-Trisubstituted and 2,2’,3,3’-Tetrasubstituted 1,1’-Binaphthyls. J. Org. Chem. 1994, 59, 2156–2163. [Google Scholar] [CrossRef]
  22. Ding, K.; Xu, Q.; Wang, Y.; Liu, J.; Yu, Z.; Du, B.; Wu, Y.; Koshima, H.; Matsuura, T. Novel two-phase oxidative cross-coupling of the two-component molecular crystal of 2-naphthol and 2-naphthylamine. Chem. Commun. 1997, 693–694. [Google Scholar] [CrossRef]
  23. Singer, R.A.; Brock, J.R.; Carreira, E.M. Synthesis of A Tridentate Ligand for Use in TiIV-Catalyzed Acetate Aldol Addition Reactions. Helv. Chim. Acta 2003, 86, 1040–1044. [Google Scholar] [CrossRef]
  24. Zhao, X.-J.; Li, Z.-H.; Ding, T.-M.; Tian, J.-M.; Tu, Y.-Q.; Wang, A.-F.; Xie, Y.-Y. Enantioselective Synthesis of 3,3′-Disubstituted 2-Amino-2′-hydroxy-1,1′-binaphthyls by Copper-Catalyzed Aerobic Oxidative Cross-Coupling. Angew. Chem. Int. Ed. 2021, 60, 7061–7065. [Google Scholar] [CrossRef]
  25. Zhang, J.-W.; Jiang, F.; Chen, Y.-H.; Xiang, S.-H.; Tan, B. Synthesis of Structurally Diversified BINOLs and NOBINs via Palladium-Catalyzed C-H Arylation with Diazoquinones. Sci. China Chem. 2021. [Google Scholar] [CrossRef]
  26. Sheradsky, T.; Avramovici-Grisaru, S. Hydrolytic cleavages of 7-nitro-2-phenyl-1,2-benzisoxazol-3-one. Evidence for the occurrence of a benzidine-like rearrangement of N,O-diphenylhydroxylamines. J. Heterocycl. Chem. 1980, 17, 189–190. [Google Scholar] [CrossRef]
  27. Gao, H.; Ess, D.H.; Yousufuddin, M.; Kürti, L. Transition-Metal-Free Direct Arylation: Synthesis of Halogenated 2-Amino-2′-hydroxy-1,1′-biaryls and Mechanism by DFT Calculations. J. Am. Chem. Soc. 2013, 135, 7086–7089. [Google Scholar] [CrossRef]
  28. De, C.K.; Pesciaioli, F.; List, B. Catalytic Asymmetric Benzidine Rearrangement. Angew. Chem. Int. Ed. 2013, 52, 9293–9295. [Google Scholar] [CrossRef]
  29. Li, G.-Q.; Gao, H.; Keene, C.; Devonas, M.; Ess, D.H.; Kürti, L. Organocatalytic Aryl-Aryl Bond Formation: An Atro-poselective [3,3]-Rearrangement Approach to BINAM Derivatives. J. Am. Chem. Soc. 2013, 135, 7414–7417. [Google Scholar] [CrossRef]
  30. Yanagi, T.; Otsuka, S.; Kasuga, Y.; Fujimoto, K.; Murakami, K.; Nogi, K.; Yorimitsu, H.; Osuka, A. Metal-Free Approach to Biaryls from Phenols and Aryl Sulfoxides by Temporarily Sulfur-Tethered Regioselective C–H/C–H Coupling. J. Am. Chem. Soc. 2016, 138, 14582–14585. [Google Scholar] [CrossRef] [PubMed]
  31. Hori, M.; Guo, J.-D.; Yanagi, T.; Nogi, K.; Sasamori, T.; Yorimitsu, H. Sigmatropic Rearrangements of Hyperva-lent-Iodine-Tethered Intermediates for the Synthesis of Biaryls. Angew. Chem. Int. Ed. 2018, 57, 4663–4667. [Google Scholar] [CrossRef] [PubMed]
  32. Guo, L.; Liu, F.; Wang, L.; Yuan, H.; Feng, L.; Kürti, L.; Gao, H. Cascade Approach to Highly Functionalized Biaryls by a Nucleophilic Aromatic Substitution with Arylhydroxylamines. Org. Lett. 2019, 21, 2894–2898. [Google Scholar] [CrossRef] [PubMed]
  33. Yuan, H.; Du, Y.; Liu, F.; Guo, L.; Sun, Q.; Feng, L.; Gao, H. Tandem Approach to NOBIN Analogues from Arylhydrox-ylamines and Diaryliodonium Salts via [3,3]-Sigmatropic Rearrangement. Chem. Commun. 2020, 56, 8226–8229. [Google Scholar] [CrossRef]
  34. Zhang, J.; Qi, L.; Li, S.; Xiang, S.; Tan, B. Direct Construction of NOBINs via Domino Arylation and Sigmatropic Rearrangement Reactions. Chin. J. Chem. 2020, 38, 1503–1514. [Google Scholar] [CrossRef]
  35. Deprez, N.R.; Sanford, M.S. Reactions of Hypervalent Iodine Reagents with Palladium: Mechanisms and Applications in Organic Synthesis. Inorg. Chem. 2007, 46, 1924–1935. [Google Scholar] [CrossRef]
  36. Phipps, R.J.; Gaunt, M. A Meta-Selective Copper-Catalyzed C-H Bond Arylation. Science 2009, 323, 1593–1597. [Google Scholar] [CrossRef] [PubMed]
  37. Duong, H.A.; Gilligan, R.E.; Cooke, M.L.; Phipps, R.J.; Gaunt, M. Copper(II)-Catalyzed meta-Selective Direct Arylation of α-Aryl Carbonyl Compounds. Angew. Chem. Int. Ed. 2010, 50, 463–466. [Google Scholar] [CrossRef]
  38. Ghosh, M.K.; Rzymkowski, J.; Kalek, M. Transition-Metal-Free Aryl-Aryl Cross-Coupling: C-H Arylation of 2-Naphthols with Diaryliodonium Salts. Chem. Eur. J. 2019, 25, 9619–9623. [Google Scholar] [CrossRef]
  39. Phipps, R.J.; Grimster, N.P.; Gaunt, M. Cu(II)-Catalyzed Direct and Site-Selective Arylation of Indoles Under Mild Conditions. J. Am. Chem. Soc. 2008, 130, 8172–8174. [Google Scholar] [CrossRef]
  40. Aradi, K.; Tóth, B.L.; Tolnai, G.L.; Novák, Z. Diaryliodonium Salts in Organic Syntheses: A Useful Compound Class for Novel Arylation Strategies. Synlett 2016, 27, 1456–1485. [Google Scholar] [CrossRef]
  41. Qi, L.-W.; Mao, J.-H.; Zhang, J.; Tan, B. Organocatalytic asymmetric arylation of indoles enabled by azo groups. Nat. Chem. 2017, 10, 58–64. [Google Scholar] [CrossRef] [PubMed]
  42. Qi, L.-W.; Li, S.; Xiang, S.-H.; Wang, J.J.; Tan, B. Asymmetric construction of atropisomeric biaryls via a redox neutral cross-coupling strategy. Nat. Catal. 2019, 2, 314–323. [Google Scholar] [CrossRef]
  43. Ding, W.-Y.; Yu, P.; An, Q.-J.; Bay, K.L.; Xiang, S.-H.; Li, S.; Chen, Y.; Houk, K.N.; Tan, B. DFT-Guided Phosphor-ic-Acid-Catalyzed Atroposelective Arene Functionalization of Nitrosonaphthalene. Chemistry 2020, 6, 2046–2059. [Google Scholar] [CrossRef]
  44. Yan, S.; Xia, W.; Li, S.; Song, Q.; Xiang, S.-H.; Tan, B. Michael Reaction Inspired Atroposelective Construction of Axially Chiral Biaryls. J. Am. Chem. Soc. 2020, 142, 7322–7327. [Google Scholar] [CrossRef] [PubMed]
  45. Phipps, R.J.; McMurray, L.; Ritter, S.; Duong, H.A.; Gaunt, M.J. Copper-Catalyzed Alkene Arylation with Diaryliodonium Salts. J. Am. Chem. Soc. 2012, 134, 10773–10776. [Google Scholar] [CrossRef]
  46. Allen, A.E.; Macmillan, D.W.C. Enantioselective α-Arylation of Aldehydes via the Productive Merger of Iodonium Salts and Organocatalysis. J. Am. Chem. Soc. 2011, 133, 4260–4263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Harvey, J.S.; Simonovich, S.P.; Jamison, C.R.; Macmillan, D.W.C. Enantioselective α-Arylation of Carbonyls via Cu(I)-Bisoxazoline Catalysis. J. Am. Chem. Soc. 2011, 133, 13782–13785. [Google Scholar] [CrossRef] [Green Version]
  48. Zhu, S.; MacMillan, D.W.C. Enantioselective Copper-Catalyzed Construction of Aryl Pyrroloindolines via an Aryla-tion-Cyclization Cascade. J. Am. Chem. Soc. 2012, 134, 10815–10818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ribas, X.; Jackson, D.A.; Donnadieu, B.; Mahía, J.; Parella, T.; Xifra, R.; Hedman, B.; Hodgson, K.O.; Llobet, A.; Stack, T.D.P. Aryl C-H Activation by CuII To Form an Organometallic Aryl-CuIII Species: A Novel Twist on Copper Disproportionation. Angew. Chem. Int. Ed. 2002, 41, 2991–2994. [Google Scholar] [CrossRef]
  50. Yao, B.; Wang, D.-X.; Huang, Z.-T.; Wang, M.-X. Room-temperature aerobic formation of a stable aryl-Cu(III) complex and its reactions with nucleophiles: Highly efficient and diverse arene C-H functionalizations of azacalix[1]arene[3]pyridine. Chem. Commun. 2009, 20, 2899–2901. [Google Scholar] [CrossRef]
  51. Ribas, X.; Calle, C.; Poater, A.; Casitas, A.; Gómez, L.; Xifra, R.; Parella, T.; Benet-Buchholz, J.; Schweiger, A.; Mitrikas, G.; et al. Facile C-H Bond Cleavage via a Proton-Coupled Electron Transfer Involving a C-H···CuII Interaction. J. Am. Chem. Soc. 2010, 132, 12299–12306. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Representative molecules deriving from NOBIN or BINAM.
Figure 1. Representative molecules deriving from NOBIN or BINAM.
Molecules 26 03223 g001
Figure 2. Our strategies for synthesis of NOBIN derivatives via copper-catalyzed domino arylation and rearrangement.
Figure 2. Our strategies for synthesis of NOBIN derivatives via copper-catalyzed domino arylation and rearrangement.
Molecules 26 03223 g002
Figure 3. Substrate scope of N-naphthylhydroxylamines and diaryliodonium salts. All reactions were performed with Cu(TFA)2 (10 mol%), 1a (0.20 mmol), 2a (0.24 mmol) and base (0.26 mmol) in DCM (4.0 mL) at room temperature; Isolated yields were provided.
Figure 3. Substrate scope of N-naphthylhydroxylamines and diaryliodonium salts. All reactions were performed with Cu(TFA)2 (10 mol%), 1a (0.20 mmol), 2a (0.24 mmol) and base (0.26 mmol) in DCM (4.0 mL) at room temperature; Isolated yields were provided.
Molecules 26 03223 g003
Figure 4. (a) Synthesis of biaryl amino alcohols under standard conditions; (b) Synthesis of BINAM derivatives under standard conditions.
Figure 4. (a) Synthesis of biaryl amino alcohols under standard conditions; (b) Synthesis of BINAM derivatives under standard conditions.
Molecules 26 03223 g004
Figure 5. Control experiments: (a) transition metal-free; (b) aluminum(III), magnesium(II), zinc(II), or nickel(II) trifluoromethanesulfonate instead of copper(II) trifluoroacetate; (c) base free.
Figure 5. Control experiments: (a) transition metal-free; (b) aluminum(III), magnesium(II), zinc(II), or nickel(II) trifluoromethanesulfonate instead of copper(II) trifluoroacetate; (c) base free.
Molecules 26 03223 g005
Figure 6. Proposed reaction pathway.
Figure 6. Proposed reaction pathway.
Molecules 26 03223 g006
Table 1. Optimization of the reaction conditions involving N-naphthylhydroxylamine a.
Table 1. Optimization of the reaction conditions involving N-naphthylhydroxylamine a.
Molecules 26 03223 i001
EntryVariation from the Optimized ConditionsYield (%) b
1none98
2toluene instead of DCM80
3THF instead of DCM90
4MeCN instead of DCM88
5EA instead of DCM86
6Cu(OAc)2 instead of Cu(TFA)289
7Cu(OTf)2 instead of Cu(TFA)282
8CuOTf instead of Cu(TFA)289
9CuI instead of Cu(TFA)286
10K2CO3 instead of Na2CO393
11Cs2CO3 instead of Na2CO390
12NaOH instead of Na2CO375
13NaOtBu instead of Na2CO366
14Et3N instead of Na2CO373
a All reactions were performed with Cu(TFA)2 (10 mol%), 1a (0.10 mmol), 2a (0.12 mmol), and base (0.13 mmol) in DCM (2.0 mL) at room temperature; b Yield was determined by 1H-NMR analysis of the crude reaction mixture using 1,3,5-trimethoxybenzene as the internal standard.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.-W.; Xiang, S.-H.; Li, S.; Tan, B. Copper-Catalyzed Synthesis of Axially Chiral Biaryls with Diaryliodonium Salts as Arylation Reagents. Molecules 2021, 26, 3223. https://doi.org/10.3390/molecules26113223

AMA Style

Zhang J-W, Xiang S-H, Li S, Tan B. Copper-Catalyzed Synthesis of Axially Chiral Biaryls with Diaryliodonium Salts as Arylation Reagents. Molecules. 2021; 26(11):3223. https://doi.org/10.3390/molecules26113223

Chicago/Turabian Style

Zhang, Ji-Wei, Shao-Hua Xiang, Shaoyu Li, and Bin Tan. 2021. "Copper-Catalyzed Synthesis of Axially Chiral Biaryls with Diaryliodonium Salts as Arylation Reagents" Molecules 26, no. 11: 3223. https://doi.org/10.3390/molecules26113223

Article Metrics

Back to TopTop