Abstract
Dynamic kinetic resolution (DKR) has emerged as an elegant and powerful tool for enantioselective synthesis, enabling the transformation of racemic compounds into enantiomerically enriched products with theoretically quantitative yields. Despite its widespread success, the dynamic kinetic asymmetric C−O cross-coupling has presented significant challenges and remains unexplored. In this study, we report a dynamic kinetic asymmetric C−O cross-coupling of oximes and phenols via copper/BOX-catalysed enantioselective O-arylation with diaryliodonium salts. This method efficiently produces a wide range of inherently chiral oxime ethers, as well as axially chiral styrenes, with high yields and excellent regio- and enantioselectivities. Through controlled experiments and Density Functional Theory (DFT) studies, we have elucidated the dynamic kinetic resolution process and gained insights into the origins of regio- and enantioselectivity.
Similar content being viewed by others
Introduction
Aryl ethers are important structural components prevalent in numerous natural products and pharmaceuticals, including vancomycin, asenapine, and riccardin C (Fig. 1a)1. Moreover, aryl ether structures exhibit a wide range of bioactivities, such as antiviral, antimicrobial, and amyloidogenesis inhibition properties, and are also utilized as chiral receptors and chiral ligands2. Over the past few decades, significant effort has been devoted to developing mild and general methods for synthesizing aryl ethers in both industry and academia3. Many catalytic systems, particularly those employing transition-metal-catalysed C−O cross-coupling, have been developed to operate on an increasingly broad scope of aryl halides, alcohols, or phenols4,5,6. Among these, copper-mediated C−O cross-coupling has garnered substantial attention due to its low cost, low toxicity, and broad applicability7,8,9,10. Despite these advancements, enantioselective techniques for synthesizing chiral aryl ethers remain significantly underdeveloped, even though the optically active products produced are highly valuable. In 2013, Beaudry and co-workers reported Cu-catalysed enantioselective synthesis of planar chiral meta-cyclophanes related to diarylether heptanoid natural products via enantioselective Ullmann cross-coupling reactions11. However, the enantioselectivity achieved was only up to 44%. Subsequently, Cai and co-workers reported copper-catalysed, highly enantioselective intramolecular O-arylation coupling reaction, forming chiral 2,3-dihydrobenzofurans using an asymmetric desymmetrization strategy (Fig. 1b)12. In 2020, Gu and co-workers developed a borinic acid-activated site-selective O-arylation of diols in a copper-catalysed asymmetric ring-opening reaction of cyclic diaryliodonium salts13. In this reaction, borinic acids nucleophilically activate diols by forming borolanuide, although this also limits the range of alcohol nucleophiles. More recently, Miller developed atroposelective Cu-catalysed C−O cross-coupling of phenols utilizing a guanidiny-lated peptidic ligand based on remote symmetry-breaking processes in diarylmethine substrates14,15,16. However, the reaction is limited to arylation reagents with o-trifluoromethanesulfonamide substitution. Despite these achievements, asymmetric O-arylation reactions are currently limited to desymmetrization strategies, and the range of applicable substrates remains greatly restricted. Thus, there is a strong appeal for the development of more facile and direct O-arylation routes.
a Representative natural products, bioactive molecules, ligands and functional molecules. b Previous copper-catalysed asymmetric O-arylation reaction. c Dynamic kinetic resolution of carbonyl, oximino, and hydroxyl compounds. d This design: Copper-catalysed dynamic kinetic resolution O-arylation with diaryliodonium salts.
Dynamic kinetic resolution (DKR), which can transform racemic compounds into enantiomerically enriched products with a theoretically quantitative yield, has proven to be an elegant and powerful tool for enantioselective synthesis (Fig. 1c)17,18,19. While DKR is a potent approach for establishing absolute configuration, it generally requires a mechanistic pathway for racemization. The DKR of carbonyl compounds bearing two substituents at the α-carbon has garnered considerable attention over the past decades20,21,22,23,24,25,26,27, especially since Noyori’s groundbreaking work on the asymmetric hydrogenation of β-ketoesters28. The success of these reactions relies on the acidity of the α-carbon, which facilitates the interconversion of enantiomers through enol tautomerization. However, despite the widespread application of tautomeric dynamic kinetic resolution in asymmetric synthesis, extending this strategy to other stereochemical motifs has been challenging. Oximes are attractive building blocks in organic synthesis due to their ability to readily interconvert into a variety of functional groups, such as nitrones, nitriles, amides, isoxazolines, and oxime ethers29,30. Due to the acidity of the oxime hydroxyl group, oximes can interconvert with nitroso compounds through double bond migration under acidic or basic conditions, offering the potential for dynamic kinetic resolution of oximes. Despite significant advancements by Takemoto31, You32, Carreira33, Zhang34, Liu35,36, and Zhang37 in the field of enantioselective and chemoselective allylic substitution, hydroamination, heterodifunctionalization, and carboetherification reactions—catalysed by Ir, Cu, and Pd complexes to achieve O- or N-allylation and alkylation of oxime derivatives—all of these reactions involve the construction of carbon-centered chiral compounds using achiral oximes as nucleophiles. To date, dynamic kinetic resolution achieved by prochiral oximes remains unexplored. Besides the interconversion of two enantiomers through tautomerization, another strategy to achieve dynamic kinetic resolution involves the ring-opening/ring-closing equilibrium of biaryl acetals, as developed by Akiyama38,39,40,41,42,43,44,45. Importantly, this DKR strategy can reduce the electron density on the phenolic aromatic ring, thereby avoiding the competition of phenol carbon arylation. We hypothesized that constructing axially chiral styrene compounds via copper-catalysed O-arylation of prochiral phenol hydroxyl groups might be achievable. Herein, we report a copper-catalysed dynamic kinetic asymmetric O-arylation of oximes and phenols with diaryliodonium salts46,47,48,49,50,51,52,53,54,55,56,57,58, delivering inherently and axially chiral aryl ethers in good yields with exceptional regio- and enantioselectivities (Fig. 1d)59,60,61,62,63,64,65,66.
Results
Optimization studies
At the outset of our studies, we selected tribenzotropone oxime (1a), derived from tribenzotropone, as an appropriate substrate for the catalytic enantioselective arylation process. Tribenzotropone oxime exhibits the favorable property of existing in equilibrium with its nitroso form in solution under basic conditions, facilitating tautomeric dynamic kinetic transformation. This hypothesis, however, faces significant challenges in selectively regulating O-, N-, or C-arylation67,68,69,70. To explore this, we first studied the arylation of tribenzotropone oxime 1a with Ph₂IPF₆ 2a (Table 1). In the presence of 10 mol % CuI, 12 mol % chiral bisoxazoline ligand L1, and NaHCO₃ as the base in dichloromethane, the selectively O-arylated product 3a was obtained in 61% yield with 42% ee at 25 oC (entry 1), confirming that copper salts could indeed catalyse this DKR transformation. Encouraged by this initial result, we investigated several copper catalysts, including Cu(OAc)₂, Cu(OTf)₂, and CuBr (Table 1, entries 2–4). However, when the reaction was performed with Cu(OAc)₂ or Cu(OTf)₂ as catalysts, almost no product 3a was obtained, and the main by-product was tribenzotropone, resulting from the decomposition of oxime 1a (entries 2–3). CuBr, on the other hand, provided the best result, yielding 3a in 72% yield with 43% ee (entry 4). We then examined several chiral bisoxazolines with different substituents (entries 5–13). The use of tBu-BOX ligand L2 proved inefficient (entry 5). The desired product was obtained in 70% yield with 38% ee when the cis-diphenyl-substituted oxazoline ligand BOX L3 was employed (entry 6). Next, we explored substituents on the methylene linker between the two oxazoline rings in the BOX ligand. Enantioselectivity was significantly improved with the gem-dibenzyl-substituted (S,S)-diphenylbisoxazoline L5, yielding 3a in 71% yield and 92% ee (entry 8). Further modification of the benzyl substituents on the oxazoline ligands did not enhance enantioselectivity (entries 9–10). Interestingly, the enantioselectivity of 3a was greatly reduced when methyl-benzyl-substituted oxazoline L8 was used, compared to the gem-dibenzyl-substituted ligand L5 (entry 11), highlighting the importance of alkyl substituents for stereoselective control. The cyclopentane-substituted (S,S)-diphenylbisoxazoline ligand L9 was not effective, providing only moderate yield and enantioselectivity (entry 12). When the sterically encumbered indaBOX ligand L10 was employed, product 3a was afforded in 52% yield with only 9% ee (entry 13). Screening different reaction media identified acetone as the best solvent (entries 14–17). Using L5 as the ligand in acetone, we then examined the effect of the counteranion of the diaryliodonium salt on the reaction outcome (X=BF₄, OTf). A comparable result was obtained with Ph₂IBF₄ 2a’ as the arylation reagent (entry 18). However, when Ph₂IOTf 2a” was used, although the enantioselectivity was 92%, the yield significantly decreased (entry 19). Additionally, the use of molecular sieves was tested, but they did not have a significant impact on the reaction outcome (entry 20).
Scope of the reaction
With the optimized conditions in hand, we then explored the substrate scope of this enantioselective O-arylation reaction (Fig. 2). First, we examined the scope of diaryliodonium salts. A variety of functional groups were well-tolerated as meta or para substitutions on the aryl ring of the diaryliodonium salts, including halogens (F, 3b, 3l; Cl, 3f, 3m; Br, 3c, 3n), methyl (3g, 3j), trifluoromethyl (3i, 3o), trifluoromethoxy (3d, 3k), ester (3e, 3p), and nitro (3h) groups. The inherently chiral oxime ethers 3 were obtained in 65%–85% yields and 81%–97% ee. The absolute configurations of 3a and 3f were determined by X-ray crystallographic analysis (CCDC: 3a 2372246; 3f 2372247). Notably, aryl substrates bearing C−Br bonds remained unaltered by the Cu(I)-catalysed arylation, providing a complementary platform for further elaboration through conventional metal-catalysed coupling reactions (3c, 3n). Diaryliodonium salts with either electron-donating or electron-withdrawing groups at the ortho-position of the aryl ring also proved to be suitable substrates, although the yields were slightly reduced when these groups were incorporated (3q-3t). Furthermore, a series of diaryliodonium salts containing multi-substituted aryl groups were successfully employed in this arylation reaction, affording the target oxime ethers in 62%–90% yields and 91%–99% ee (3u-3ab).
Encouraged by the previous results, we next investigated the substrate scope of tribenzotropone oximes 1. As shown in Fig. 3, a variety of electronically diverse tribenzotropone oximes, with substitutions on either the bottom or edge aromatic rings, performed well in this process, delivering the O-arylated products 3ac-3ao in 65%–87% yields with diastereoselectivities ranging from 1:1 to 1.4:1. The enantioselectivity of the diastereomers was nearly identical, with most achieving enantioselectivities above 90%. Notably, the reaction is not limited to polyphenyl-substituted oximes; it also proceeds smoothly with the introduction of nitrogen-containing heterocycles (3ap). However, the enantioselectivity of the product was reduced. We assessed the racemization of 3ap at various time intervals (see Supplementary Information Figure S5 for details). Experimental results indicated that the configurational stability of product 3ap was relatively low at 25 °C in iPrOH, and the measured rotational energy barrier was 24.1 kcal/mol for 3ap (at 25 °C in iPrOH). This suggests that the substrate 3ap has relatively low activation energy, and extended reaction times lead to slow racemization.
Having examined the asymmetric C−O coupling of oximes with diaryliodonium salts, we next examined the dynamic kinetic asymmetric arylation of phenols (see Supplementary information Table S2 for detail optimization of the reaction conditions). Figure 4 shows that a diverse array of diaryliodonium salts and 1-enal substituted 2-naphthols were successfully converted into axially chiral styrene products with good yields and enantioselectivities71,72,73,74,75,76. First, we examined the reactivity of a simple diaryliodonium salt 2 with a series of analogs that included methyl, halogen groups, including F, Cl, Br and trifluoromethyl substituents, which successfully yielded the targeted products 5a–5f. An X-ray diffraction analysis of a single crystal of 5a confirmed that its absolute configuration was S. Substrates carrying an electron-donating or electron-withdrawing group at the meta-positions on phenyl rings (5g–5j) resulted in high yields and excellent enantioselectivities. It is worth noting that the product 5a (24.4 kcal/mol, iPrOH, 25 °C) underwent slightly racemization during column chromatography. Consequently, the aldehyde was converted into an alcohol with a higher energy barrier, allowing us to conduct further substrate scope tests. When an electron-donating methoxy group was used, the corresponding product 6k was obtained in 74% yield with 92% ee. Moreover, substrates featuring electron-withdrawing functional groups such as ester (6l), trifluoromethoxy (6m and 6o), and even nitro (6n) groups have emerged as appropriate choices, affording good-to-high yields while maintaining high levels of enantioselectivity. A significant exception was noted for the diaryliodonium salt 6p, which had a methyl group at its ortho position. This alteration resulted in a decrease in the enantioselectivity, emphasizing the pronounced influence of steric hindrance at this location on stereochemical control. Double-halogen-substituted diaryliodonium salts (6q–6t), demonstrating good substrate compatibility and enantioselectivities ranging from 90% to 93%, were also efficiently synthesized via this method. Subsequently, we investigated the broad spectrum of 1-enal substituted 2-naphthols 4 by introducing various substituents on aromatic rings R1 and R2. Various labile functionalities were efficiently incorporated into the R1 of naphthols, including bromine (6u and 6y), ester (6v), nitrile (6w), phenyl (6x), and methoxy (6z) groups. This indicates the high tolerance of different functional groups in the current catalytic approach. Notably, the nitrile-substituted substrate was suitable for the described reaction, although it resulted in slightly reduced enantioselectivity (6w). Subsequently, the aromatic ring R2 featuring various ortho- and meta-substituents, such as phenyl, halogen, nitro, and trifluoromethyl groups, were effectively utilized, producing the desired the product (6aa–6af) in yields of 76–94% and excellent enantioselectivities. Para-substituent substrates, whether with electron-donating or electron-withdrawing substituents, smoothly participated in the reaction, showing good enantioselectivity and resulting in the corresponding products 6ag–6ak in yields of 51–95%. Importantly, this method was effective for the transformation of electron-rich naphthyl aldehyde substrate (6al). Similarly, substitution with heteroaromatic furan and thiophene groups resulted in seamless reactions affording the products 6am and 6an, respectively, achieving yields of 52–56% and >82% ee.
Reaction Conditions: a4 (0.12 mmol), 2 (0.10 mmol), CuBr (10 mol%), L7 (12 mol%), Na2CO3 (1.5 equiv.) in DCE (1.0 mL) at −20 °C for 48 h; isolated yields; ee values were determined by chiral HPLC. bAfter reaction, NaBH4 (1.5 equiv.) in 1.0 mL MeOH was added and the reaction was stirred at −20 °C for 4 h. cReacted for 5 days.
The effectiveness of our method was further demonstrated through the modification of drug and natural product derivatives (Fig. 5a). Substrates derived from febuxostat, lithocholic acid, and linoleic acid were successfully transformed into the corresponding products 8a–8c with respectable yields and enantio- and diastereoselectivities. To further highlight the synthetic value of this method, a series of experiments was performed (Fig. 5b) using 4a and 2a as substrates. The synthesis was smoothly conducted on a 2.0 mmol scale under standard conditions. This process resulted in the production of product 6a with 75% yield while maintaining the enantioselectivity. Subsequently, allylic alcohol product 6a successfully reacted with acetic anhydride and iodomethane, resulting in the production of the corresponding esterification and alkylation 9a and 9b, respectively, with a slight decrease in the enantiomeric ratio. Additionally, the silylation of 6a was performed, leading to the synthesis of 9c with a high enantioselectivity.
Stereochemical stability study and control experiments
To assess the stereochemical stability of these chiral aryl ether compounds, we conducted racemization experiments on oxime 1a, oxime ether 3a, and axially chiral styrene 6ak over time at various temperatures. As depicted in Fig. 6a, the ee value of 3a decreased slightly to 83% after 3.5 h in iPrOH at 70 °C. When the solution was maintained at 80 °C, the ee value of 3a gradually dropped to 75% over 3.5 h. Based on these experimental results, the rotational barrier for 3a was calculated to be 28.9 kcal/mol at 80 °C. Interestingly, oxime 1a exhibited similar stereochemical stability to oxime ether 3a, with a rotational barrier of 27.7 kcal/mol at 60 °C. Further experiments showed that axially chiral styrene 6ak remained relatively stable at 50 °C in iPrOH, with a measured rotational energy barrier of 28.1 kcal/mol at 70 °C. Given the relatively high inversion energy barrier of oximes, several controlled experiments were conducted to investigate their dynamic kinetic transformation at room temperature (Fig. 6b). We assessed the stability of enantioenriched oxime (+)-1a under base or base/copper conditions similar to those in the optimal process. In these cases, the remaining oxime 1a was partially or almost fully racemized. Additionally, we demonstrated that the reaction of enantioenriched (+)-1a under standard conditions, but without the chiral ligand L5, yielded 3a in 60% yield with 66% ee after 12 h. Furthermore, the reaction of (+)-1a with (S,S)-L5 as a ligand resulted in the productive, matched formation of the inherently chiral 3a with high ee. Conversely, the mismatched experiment using (R,R)-L5 as the ligand produced the opposite enantiomer of 3a in a similar yield but with 80% ee. These results confirmed that during the reaction, the optically active oxime 1a underwent racemization and then reacted with the matching ligand to achieve highly enantioselective arylation products. Collectively, these experiments elucidated the process of dynamic kinetic resolution in oximes 1.
a Racemization experiments on oxime 1a, oxime ether 3a, and axially chiral styrene 6ak. b Control experiments for the dynamic kinetic resolution of 1a. c Kinetic experiment towards a dynamic kinetic resolution (DKR) pathway. d Non-linear experiments. e Hammett plot for selected diaryliodonium salts.
To clarify the mechanism of the asymmetric C-O coupling reaction of phenols, a sequence of initial experiments was performed. Initially, a Cu-catalyzed O-arylation reaction between substrates 4a and 2a was performed under standard conditions to study the dynamic kinetic resolution (DKR) process (Fig. 6c). During the reaction, the ee value of 4a remained racemic, which indicated that the reaction underwent a transition state involving a five-membered aryl alkene lactol under the influence of sodium carbonate (base). This suggests that rapid dynamic kinetic resolution occurred under these reaction conditions. Initially, the ee value of 5a increased during the reaction, reaching a peak (91% ee) and stabilizing after 11 h. This pattern could be due to a strong background reaction that occurred during the early stages of the process. Additionally, experiments on nonlinear effects were carried out (Fig. 6d). The observed positive non-linear effect between 5a and L7 implied that the actual catalyst in the catalytic cycle might not be a mononuclear CuBr·L7 complex77,78,79. To better understand the enantiodetermining step of the O-arylation reaction, we examined the Hammett electronic parameters of different para- and meta-substituted diaryliodonium salts and enantioselectivities of the resulting products (Fig. 6e). Hammett plots exhibited a linear correlation (R2 = 0.9655) with a negative slope (ρ = −0.3500), suggested that copper intermediates generated from electron-rich diaryliodonium salts react more effectively with 1-enal substituted 2-naphthol 1a than those from electron-deficient diaryliodonium salts.
To further elucidate the origins of the regio- and enantioselectivity for the C−O cross-coupling reaction, density functional theory (DFT) calculations were performed with the B3LYP-D3 functional (see SI for computational details). As shown in Fig. 7a, starting from the CuI(Br)(L5) complex int1, the oxidative addition of the diaryliodonium salt80,81 2 has an activation free energy barrier of 24.2 kcal/mol via TS_OA and is slightly exergonic to lead to the CuIII(Br)(Ph)(L5) intermediate int2. Following that, the ligand exchange between oximes (+)-1, NaHCO3, and BF4− is thermodynamically favorable to provide int3, in which the O atom of the substrate binds to the Cu. Subsequently, the reductive elimination via transition state TS1 would afford the O-Ph coupling product. Alternatively, int3 would first isomerize to a less stable intermediate int4, in which the N atom binds to the Cu, and followed by the reductive elimination through TS2 to produce the N-Ph coupling product. By comparison, the pathway leading to the experimentally observed O-Ph coupling product (via TS1) is 3.2 kcal/mol more favorable than that for the N-Ph coupling product (via TS2). Thus, computations are consistent with experiment on the regioselectivity. It should be noted that we have also considered other possible pathways for the coupling step64,71 (e.g., the transmetalation-like pathway) but none of them were more favorable than the reductive elimination mechanism (see SI for details).
Calculations were conducted to elucidate the enantioselectivity of the reaction. Starting from intermediate int2, the ligand exchange between oximes (-)-1, NaHCO3, and BF4− leads to the formation of int3’. However, this pathway is thermodynamically less favorable compared to the formation of int3. The subsequent reductive elimination occurs via transition state TS1’, which would generate the enantiomer opposite to the experimentally observed product. This minor pathway, progressing from int3’ to TS1’, has a higher energy barrier by 1.4 kcal/mol compared to the major pathway from int3 to TS1. These findings are consistent with the experimentally observed enantioselectivity.
A detailed analysis of the selectivity-determining transition states (Fig. 7b) reveals key factors influencing the reaction’s enantioselectivity. In the N-Ph coupling transition state TS2, significant steric repulsions occur between the ligand and the substrates, with the closest H-H distances measured at 2.06 and 2.12 Å, respectively. In contrast, no such steric repulsion is observed in the O-Ph coupling transition state TS1. Additionally, TS1 benefits from favorable C-H…π interactions between the phenyl groups of the ligand and oximes, as well as within the ligand itself, with distances of 2.64 and 2.63 Å, respectively. In addition, there is also a favorable C-H…Br interaction between the substrate and the Br atom (see SI for details of NCI analysis). In the minor enantioselective transition state TS1’, there is only one C-H…π interaction were found. Consequently, the observed enantioselectivity is primarily attributed to these non-covalent interactions, while both non-covalent interactions and steric effects contribute to the regioselectivity of the reaction.
Discussion
In conclusion, we reported a dynamic kinetic asymmetric C−O cross-coupling of oximes and phenols via copper/BOX-catalysed enantioselective O-arylation with diaryliodonium salts. This approach efficiently prepared a wide range of inherently and axially chiral oxime ethers and axially chiral styrenes in high yields with excellent regio- and enantioselectivities. Control experiments revealed the dynamic kinetic resolution process of oximes, while DFT studies provided insights into the origins of regio- and enantioselectivity. The reaction features mild conditions, good functional group tolerance, readily available starting materials, and commercial bisoxazoline ligands, indicating its potential for broad applications in organic synthesis.
Methods
General procedure for the synthesis of chiral aryl oxime ethers 3
Under air atmosphere, to a mixture of 1 (0.24 mmol, 1.2 equiv.), 2 (0.20 mmol, 1.0 equiv.), CuBr (2.9 mg, 0.02 mmol, 10 mol%), L5 (11.7 mg, 0.024 mmol, 12 mol%), NaHCO3 (16.8 mg, 0.20 mmol, 1.0 equiv.) in a sealed vial was added acetone (1.0 mL) at 25 °C. Upon completion of the reaction, the solvent was removed under vacuum, and the crude product was purified directly by column chromatography to yield the desired product 3.
General procedure for the synthesis of chiral aryl diaryl ethers 5
Under an air atmosphere, a mixture of compound 4 (0.12 mmol, 1.2 equiv.), compound 2 (0.10 mmol, 1.0 equiv.), CuBr (1.4 mg, 0.01 mmol, 10 mol%), L7 (7.2 mg, 0.012 mmol, 12 mol%), and Na₂CO₃ (15.9 mg, 0.15 mmol, 1.5 equiv.) was added in a sealed vial. This was followed by the addition of 1.0 mL of DCE. The reaction mixture was then stirred at −20 °C for 48 h. Upon completion, the solvent was removed under vacuum, and the crude product was purified directly by column chromatography to obtain the desired product 5.
Additional information
Supplementary information and chemical compound information are available in the online version of the paper. Reprints and permission information is available online at http://www.nature.com/reprints/.
Data availability
The data generated in this study are provided in the Supplementary Information file. The experimental procedures, data of NMR, and HRMS have been deposited in the Supplementary Information file and source data are provided with this paper. Crystallographic data for the structures reported in this Article have been deposited at the Cambridge Crystallographic Data Centre under deposition numbers CCDC 2372246 (3a), 2372247 (4l), and 2358843 (5a). Copies of the data can be obtained free of charge via www.ccdc.cam.ac.uk/getstructures. All other data supporting the findings of the study, including experimental procedures and compound characterization, are available within the paper and its Supplementary Information, or from the corresponding author upon request. Source data are provided with this paper.
References
Scott, K. A., Cox, P. B. & Njardarson, J. T. Phenols in pharmaceuticals: analysis of a recurring motif. J. Med. Chem. 65, 7044–7072 (2022).
Yang, Q., Zhao, Y. & Ma, D. Cu-mediated Ullmann-type cross-coupling and industrial applications in route design, process development, and scale-up of pharmaceutical and agrochemical processes. Org. Process Res. Dev. 26, 1690–1750 (2022).
Mandal, S. et al. A review on the advancement of ether synthesis from organic solvent to water. RSC Adv. 6, 69605–69614 (2016).
Stambuli, J. P. Transition metal-catalyzed formation of C-O and C-S bonds. In New Trends In Cross-coupling: Theory And Application (ed. Colacot, T. J.) 254−275 (Royal Society of Chemistry, 2014).
Tasker, S. Z., Standley, E. A. & Jamison, T. F. Recent advances inhomogeneous nickel catalysis. Nature 509, 299–309 (2014).
Hao, Y.-J., Hu, X.-S., Zhou, Y., Zhou, J. & Yu, J.-S. Catalytic enantioselective α-arylation of carbonyl enolates and related compounds. ACS Catal. 10, 955–993 (2020).
Bhunia, S., Pawar, G. G., Kumar, S. V., Jiang, Y. & Ma, D. Selected copper-based reactions for C-N, C-O, C-S, and C-C bond formation. Angew. Chem. Int. Ed. 56, 16136–16179 (2017).
Cheng, L.-J. & Mankad, N. P. C-C and C-X coupling reactions of unactivated alkyl electrophiles using copper catalysis. Chem. Soc. Rev. 49, 8036–8064 (2020).
Ma, D. & Cai, Q. Copper/amino acid catalyzed cross-couplings of aryl and vinyl halides with nucleophiles. Acc. Chem. Res. 41, 1450–1460 (2008).
Cook, A. & Newman, S. G. Alcohols as substrates in transition-metal-catalyzed arylation, alkylation, and related reactions. Chem. Rev. 124, 6078–6144 (2024).
Salih, M. Q. & Beaudry, C. M. Enantioselective Ullmann ether couplings: syntheses of (−)-myricatomentogenin, (−)-jugcathanin, (+)-galeon, and (+)-pterocarine. Org. Lett. 15, 4540–4543 (2013).
Yang, W., Liu, Y., Zhang, S. & Cai, Q. Copper-catalyzed intramolecular desymmetric aryl C−O coupling for the enantioselective construction of chiral dihydrobenzofurans and dihydrobenzopyrans. Angew. Chem. Int. Ed. 54, 8805–8808 (2015).
Zhao, K., Yang, S., Gong, Q., Duan, L. & Gu, Z. Diols activation via Cu/Borinic acids synergistic catalysis in atroposelective ring‐opening of cyclic diaryliodoniums. Angew. Chem. Int. Ed. 60, 5788–5793 (2021).
Chinn, A. J., Kim, B., Kwon, Y. & Miller, S. J. Enantioselective intermolecular C−O bond formation in the desymmetrization of diarylmethines employing a guanidinylated peptide-based catalyst. J. Am. Chem. Soc. 139, 18107–18114 (2017).
Kwon, Y., Chinn, A. J., Kim, B. & Miller, S. J. Divergent control of point and axial stereogenicity: catalytic enantioselective C−N bond-forming cross-coupling and catalyst-controlled atroposelective cyclodehydration.Angew. Chem. Int. Ed. 57, 6251–6255 (2018).
Hwang, J., Mercado, B. Q. & Miller, S. J. Chirality-matched catalyst-controlled macrocyclization reactions. Proc. Natl. Acad. Sci. USA 118, e2113122118 (2021).
Huerta, F. F., Minidis, A. B. E. & Backvall, J.-E. Racemisation in asymmetric synthesis. Dynamic kinetic resolution and related processes in enzyme and metal catalysis. Chem. Soc. Rev. 30, 321–331 (2001).
Verho, O. & Backvall, J.-E. Chemoenzymatic dynamic kinetic resolution: a powerful tool for the preparation of enantiomerically pure alcohols and amines. J. Am. Chem. Soc. 137, 3996–4009 (2015).
Bhat, V., Welin, E. R., Guo, X. & Stoltz, B. M. Advances in stereoconvergent catalysis from 2005 to 2015: transition-metal-mediated stereoablative reactions, dynamic kinetic resolutions, and dynamic kinetic asymmetric transformations. Chem. Rev. 117, 4528–4561 (2017).
Zhang, Z., Butt, N. A. & Zhang, W. Asymmetric hydrogenation of nonaromatic cyclic substrates. Chem. Rev. 116, 14769–14827 (2016).
Bartlett, S. L. & Johnson, J. S. Synthesis of complex glycolates by enantioconvergent addition reactions. Acc. Chem. Res. 50, 2284–2296 (2017).
Wang, H., Wen, J. & Zhang, X. Chiral tridentate ligands in transition metal-catalyzed asymmetric hydrogenation. Chem. Rev. 121, 7530–7567 (2021).
Cotman, A. E., Cahard, D. & Mohar, B. Stereoarrayed CF3-substituted 1,3-diols by dynamic kinetic resolution: ruthenium(II)-catalyzed asymmetric transfer hydrogenation. Angew. Chem. Int. Ed. 55, 5294–5298 (2016).
Bin, H.-Y. et al. Scalable enantioselective total synthesis of (−)goniomitine. Angew. Chem. Int. Ed. 58, 1174–1177 (2019).
DeHovitz, J. S. et al. Static toinducibly dynamic stereocontrol: the convergent use of racemicβ-substituted ketones. Science 369, 1113–1118 (2020).
Chen, T. et al. Dynamic kinetic resolution of β-substituted α-diketones via asymmetric transfer hydrogenation. J. Am. Chem. Soc. 145, 585–599 (2023).
Ruan, L.-X., Sun, B., Liu, J.-M. & Shi, S.-L. Dynamic kinetic asymmetric arylationand alkenylation of ketones. Science 379, 662–670 (2023).
Noyori, R., Tokunaga, M. & Kitamura, M. Stereoselective organic synthesis via dynamic kinetic resolution. Bull. Chem. Soc. Jpn 68, 36–55 (1995).
Kölmel, D. K. & Kool, E. T. Oximes and hydrazones in bioconjugation: mechanism and catalysis. Chem. Rev. 117, 10358–10376 (2017).
Bolotin, D. S., Bokach, N. A., Demakova, M. Y. & Kukushkin, V. Y. Metal-involving synthesis and reactions of oximes. Chem. Rev. 117, 13039–13122 (2017).
Miyabe, H., Matsumura, A., Moriyama, K. & Takemoto, Y. Utility of the iridium complex of the pybox ligand in regio- and enantioselective allylic substitution. Org. Lett. 6, 4631–4634 (2004).
Wang, Y., Zhang, W.-Y. & You, S.-L. Ketones and aldehydes as O-nucleophiles in iridium-catalyzed intramolecular asymmetric allylic substitution reaction. J. Am. Chem. Soc. 141, 2228–2232 (2019).
Sandmeier, T. & Carreira, E. M. Enantioselective synthesis of cyclic nitrones by chemoselective intramolecular allylic alkylationof oximes. Angew. Chem. Int. Ed. 60, 9913–9918 (2021).
Li, Z. et al. Asymmetric nitrone synthesis via ligand-enabled copper-catalyzed cope-type hydroamination of cyclopropene with oxime. J. Am. Chem. Soc. 139, 11702–11705 (2017).
Li, X.-T., Gu, Q.-S., Dong, X.-Y., Meng, X. & Liu, X.-Y. A copper catalyst with a cinchona-alkaloid-based sulfonamide ligand for asymmetric radical oxytrifluoromethylation of alkenyl oximes. Angew. Chem. Int. Ed. 57, 7668–7672 (2018).
Li, X.-T. et al. Diastereo- and enantioselective catalytic radical oxysulfonylation of alkenes in β,γ-unsaturated ketoximes. Chem 6, 1692–1706 (2020).
Wang, L. et al. Enantioselective synthesis of isoxazolines enabled by palladium-catalyzed carboetherification of alkenyl oximes. Angew. Chem. Int. Ed. 59, 4421–4427 (2020).
Mori, K., Itakura, T. & Akiyama, T. Enantiodivergent atroposelective synthesis of chiral biaryls by asymmetric transfer hydro-genation: chiral phosphoric acid catalyzed dynamic kinetic resolution. Angew. Chem. Int. Ed. 55, 11642–11646 (2016).
Carmona, J. A., Rodríguez-Franco, C., Fernández, R., Hornillos, V. & Lassaletta, J. M. Atroposelective transformation of axially chiral (hetero)biaryls. From desymmetrization to modern resolution strategies. Chem. Soc. Rev. 50, 2968–2983 (2021).
Liu, C.-X., Zhang, W.-W., Yin, S.-Y., Gu, Q. & You, S.-L. Synthesis of atropisomers by transition-metal-catalyzed asymmetric C−H functionalization reactions. J. Am. Chem. Soc. 143, 14025–14040 (2021).
Guo, F., Fang, S., He, J., Su, Z. & Wang, T. Enantioselective organocatalytic synthesis of axially chiral aldehyde-containing styrenes via SNAr reaction-guided dynamic kinetic resolution. Nat. Commun. 14, 5050 (2023).
Cheng, J. K., Xiang, S.-H., Li, S., Ye, L. & Tan, B. Recent advances in catalytic asymmetric construction of atropisomers. Chem. Rev. 121, 4805–4902 (2021).
Zhang, J. & Wang, J. Atropoenantioselective redox-neutral amination of biaryl compounds through borrowing hydrogen and dynamic kinetic resolution. Angew. Chem. Int. Ed. 57, 465–469 (2018).
Lv, Y. et al. Catalytic atroposelective synthesis of axially chiral benzonitriles via chirality control during bond dissociation and CN group formation. Nat. Commun. 13, 36 (2022).
Hao, X. et al. Atroposelective synthesis of axial biaryls by dynamic kinetic resolution using engineered imine reductases. Angew. Chem. Int. Ed. 63, e202410112 (2024).
Yoshimura, A. & Zhdankin, V. V. Advances in synthetic applications of hypervalent iodine compounds. Chem. Rev. 116, 3328–3435 (2016).
Phipps, R. J. & Gaunt, M. J. A meta-selective copper-catalyzed C-H bond arylation. Science 323, 1593–1597 (2009).
Bigot, A., Williamson, A. E. & Gaunt, M. J. Enantioselective α-arylation of N-acyloxazolidinones with copper(II)-bisoxazoline catalysts and diaryliodonium salts. J. Am. Chem. Soc. 133, 13778–13781 (2011).
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. 133, 13782–13785 (2011).
Kieffer, M. E., Chuang, K. V. & Reisman, S. E. Copper-catalyzed diastereoselective arylation of tryptophan derivatives: total synthesis of (+)-naseseazines A and B. J. Am. Chem. Soc. 135, 5557–5560 (2013).
Liu, C. et al. Copper(I)-catalyzed asymmetric dearomatization of indole acetamides with 3-indolylphenyliodonium salts. Chem. Eur. J. 22, 10813–10816 (2016).
Zhang, F. et al. Cu-catalyzed cascades to carbocycles: union of diaryliodonium salts with alkenes or alkynes exploiting remote carbocations. J. Am. Chem. Soc. 136, 8851–8854 (2014).
Yang, Y., Li, R., Zhao, Y., Zhao, D. & Shi, Z. Cu-catalyzed direct C6-arylation of indoles. J. Am. Chem. Soc. 138, 8734–8737 (2016).
Wu, H., Wang, Q. & Zhu, J. Copper-catalyzed enantioselective arylative desymmetrization of prochiral cyclopentenes with diaryliodonium salts. Angew. Chem. Int. Ed. 57, 2721–2725 (2018).
Zhao, K. et al. Enhanced reactivity by torsional strain of cyclic diaryliodonium in cu-catalyzed enantioselective ring-opening reaction. Chem 4, 599–612 (2018).
Frey, J. et al. Enantioselective synthesis of N-C axially chiral compounds by Cu-catalyzed atroposelective aryl amination. Angew. Chem. Int. Ed. 59, 8844–8848 (2020).
Escudero-Casao, M., Licini, G. & Orlandi, M. Enantioselective α-arylation of ketones via a novel Cu(I)−bis(phosphine) dioxide catalytic system. J. Am. Chem. Soc. 143, 3289–3294 (2021).
Zhang, X. et al. Atroposelective ring opening of cyclic diaryliodonium salts with bulky anilines controlled by a chiral cobalt(III) anion. Angew. Chem. Int. Ed. 59, 19899–19904 (2020).
Tang, M. & Yang, X. Catalytic enantioselective synthesis of inherently chiral molecules: recent advances. Eur. J. Org. Chem. 26, e202300738 (2023).
Tong, S. et al. Catalytic enantioselective synthesis and switchable chiroptical property of inherently chiral macrocycles. J. Am. Chem. Soc. 142, 14432–14436 (2020).
Zhang, Y. Z., Xu, M. M., Si, X. G., Hou, J. L. & Cai, Q. Enantioselective synthesis of inherently chiral calix[4]arenes via palladium-catalyzed asymmetric intramolecular C–H arylations. J. Am. Chem. Soc. 144, 22858–22864 (2022).
Zhang, X., Tong, S., Zhu, J.-P. & Wang, M.-X. Inherently chiral calixarenes by a catalytic enantioselective desymmetrizing cross-dehydrogenative coupling. Chem. Sci. 14, 827–832 (2023).
Luo, Y. et al. A new saddle-shaped aza analog of tetraphenylene: atroposelective synthesis and application as a chiral acylating reagent. CCS Chem. 4, 2897–2905 (2022).
Luo, Y. et al. Inherently chiral 6,7-diphenyldibenzo[e,g][1,4]diazocine: enantioselective synthesis and application as a ligand platform. CCS Chem. 5, 982–993 (2023).
Li, J.-H. et al. Organocatalytic enantioselective synthesis of seven-membered ring with inherent chirality. Angew. Chem. Int. Ed. 63, e202319289 (2024).
Zhang, D., Zhou, J., Qin, T. & Yang, X. Asymmetric synthesis of saddle-shaped eight-membered azaheterocycles via (dynamic) kinetic resolution. Chem. Catal. 4, 100827 (2024).
Ma, X. P. et al. Synthesis of α,β-unsaturated N-aryl ketonitrones from oximes and diaryliodonium salts: observation of a metal-free N-arylation process. J. Org. Chem. 80, 10098–10107 (2015).
Gao, H., Xu, Q. L., Keene, C. & Kürti, L. Scalable, transition-metal-free direct oxime O-arylation: rapid access to O-arylhydroxylamines and substituted benzo[b]furans. Chem. Eur. J. 20, 8883–8887 (2014).
Mo, D. L. & Anderson, L. L. Copper-catalyzed rearrangement of N-aryl nitrones into epoxyketimines. Angew. Chem. Int. Ed. 52, 6722–6725 (2013).
Maimone, T. J. & Buchwald, S. L. Pd-catalyzed O-arylation of ethyl acetohydroximate: synthesis of O-arylhydroxylamines and substituted benzofurans. J. Am. Chem. Soc. 132, 9990 (2010).
Wu, S., Xiang, S.-H., Cheng, J.-K. & Tan, B. Axially chiral alkenes: atroposelective synthesis and applications. Tetrahedron Chem. 1, 100009 (2022).
Qian, P.-F., Zhou, T. & Shi, B.-F. Transition-metal-catalyzed atroposelective synthesis of axially chiral styrenes. Chem. Comm. 59, 12669–12684 (2023).
Zhang, H.-H. & Shi, F. Organocatalytic atroposelective synthesis of indole derivatives bearing axial chirality: strategies and applications. Acc. Chem. Res. 18, 2562–2580 (2022).
Chen, B., Hou, X.-L., Li, Y.-X. & Wu, Y.-D. Mechanistic understanding of the unexpected meta selectivity in copper-catalyzed anilide CH bond arylation. J. Am. Chem. Soc. 133, 7668–7671 (2011).
Shao, Y.-D. et al. Construction of axially chiral styrene-type allylamines via chiral phosphoric acid-catalyzed asymmetric reductive amination. Org. Chem. Front. 9, 764–770 (2022).
Fang, S. et al. Atherton−todd reaction-guided enantioselective synthesis of axially chiral olefins via bifunctional phosphonium salt-regulatingketone-enol tautomerism. ACS. Catal. 13, 13077–13088 (2023).
Satyanarayana, T., Abraham, S. & Kagan, H. B. Nonlinear effects in asymmetric catalysis. Angew. Chem. Int. Ed. 48, 456–494 (2009).
Bryliakov, K. P. Dynamic nonlinear effects in asymmetric catalysis. ACS Catal. 9, 5418–5438 (2019).
Ali, C., Blackmond, D. G. & Burés, J. Kinetic rationalization of nonlinear effects in asymmetric catalytic cascade reactions under curtin–hammett conditions. ACS Catal. 12, 5776–5785 (2022).
Ichiishi, N., Canty, A. J., Yates, B. F. & Sanford, M. S. Cu-catalyzed fluorination of diaryliodonium salts with KF. Org. Lett. 15, 5134–5137 (2013).
Ichiishi, N., Canty, A. J., Yates, B. F. & Sanford, M. S. Mechanistic investigations of Cu-catalyzed fluorination of diaryliodonium salts: elaborating the CuI/CuIII manifold in copper catalysis. Organometallics 33, 5525–5534 (2014).
Acknowledgements
This work was supported by the financial support from the National Natural Science Foundation of China (22371152 R.L.), Taishan Scholar Youth Expert Program in Shandong Province (tsqn 201909096 R.L., tsqn202312066 L.X.), Natural Science Foundation of Shandong (ZR2023JQ006 R.L., ZR2022MB021 L.X., ZR2024YQ016 L.X., ZR2024QB091 L.X.), Qingdao Natural Science Foundation 23-2-1-244-zyyd-jch R.L., the Youth Innovation Team Program in Colleges and Universities of Shandong Province (2022KJ228 L.X.), and the Qilu Young Scholar of Shandong University.
Author information
Authors and Affiliations
Contributions
M.Z, H.W., Y.T., L.X., C.L., X.D., and C.S. performed the experiments. All authors contributed to analyzing the experimental resulted. H.S. and L.X. performed the DFT study, L.X. and R.L. supervised the project and wrote the paper.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Zhishan Su and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Zhang, MR., Wang, HR., Shan, HM. et al. Copper-catalysed dynamic kinetic asymmetric C−O cross-coupling to access chiral aryl oxime ethers and diaryl ethers. Nat Commun 16, 2505 (2025). https://doi.org/10.1038/s41467-025-57804-8
Received:
Accepted:
Published:
Version of record:
DOI: https://doi.org/10.1038/s41467-025-57804-8