Open Access Article
Huikun Yana,
Yuanyuan Zhub,
Gongming Yang
*a and
Shuangxi Gu
*a
aState Key Laboratory of Green and Efficient Development of Phosphorus Resources, Key Laboratory for Green Chemical Process of Ministry of Education, Hubei Key Laboratory of Novel Reactor and Green Chemical Technology, School of Chemical Engineering & Pharmacy, Wuhan Institute of Technology, Wuhan 430205, China. E-mail: ygm18@tsinghua.org.cn; shuangxigu@163.com
bSchool of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430205, China
First published on 29th October 2025
Chiral macrocyclic units are not only prevalent in natural products, bioactive molecules, and functional compounds, but also play significant roles in synthetic and host–guest chemistry. Although extensive efforts have been devoted to constructing chiral macrocycles, few methods have been disclosed to date. Consequently, the rapid enantioselective construction of optically active macrocycles remains a formidable challenge. N-Heterocyclic carbene (NHC) catalysis, a highly successful organocatalytic approach, has emerged as a powerful tool for rapidly constructing complex molecular architectures. However, only recently has this strategy been applied to achieve enantioselective synthesis of chiral macrocycles. This review highlights recent advances in NHC-catalyzed enantioselective synthesis of chiral macrocycles—including centrally chiral, planar chiral, and inherently chiral macrocycles, thereby providing a timely overview and foundation for future research.
In comparison with their corresponding linear analogues, the chiral macrocyclic molecules display distinct advantages due to their stable spatial conformations, which open avenues for scientists to explore new drugs, functional materials, as well as chiral catalysts and ligands.8–12 As outlined in Fig. 1b, centrally and planar chiral macrocyclic units have been widely found in many natural products and bioactive molecules,13,14 such as antituberculosis active pyridomycin, macrocyclic immunosuppressive agent FR252921,15 and planar chiral macrocyclic natural product darobactin A,16,17 which has selective inhibition against Gram-negative bacteria. Meanwhile, the planar chiral [2.2]paracyclophane skeletons have significant applications in materials science18 and asymmetric catalysis.19 Finally, the inherently chiral macrocyclic scaffolds, especially the calix[4]arenes, are of great importance in the area of enantioselective synthesis, chiral recognition,20 and host–guest chemisty.21,22 Owing to their widespread applications in drug discovery and synthetic chemistry, extensive efforts have been devoted to synthesizing enantiopure macrocycles. However, only limited methods have been reported to date for the assembly of chiral macrocycles when compared to the synthesis of other classes of chiral molecules. Overall, the rapid access to enantiomerically pure macrocyclic molecules in a highly enantioselective fashion is still in its infancy.
N-Heterocyclic carbene (NHC), one of the most successful organocatalysts, has been recognized to be a powerful tool for rapid construction of complex chiral scaffolds.23–25 Mechanistically, NHCs achieve highly enantioselective catalysis by taking advantage of their strong electron-donating ability, tunable steric environments, as well as capacity to generate diverse and reactive chiral intermediates (Fig. 2, such as Breslow, acyl azolium, azolium enolate, and Baylis–Hillman-type intermediates).26–28 Therefore, the NHC asymmetric catalysis has attracted extensive attentions from chemists and achieved significant developments in recent years. However, there are only very few methods have been reported to construct chiral macrocyclic systems to date when compared with the asymmetric preparation of other type of chiral molecules catalyzed by NHCs.
This review aims to summarize the recent progress in NHC-catalyzed enantioselective synthesis of chiral macrocycles, offering a critical overview of current research status while highlighting emerging strategies and providing future research directions in this rapid evolving field. In order to guide the reader through this topic, we categorize thesis reactions into four sections according to the asymmetric characteristics of the chiral macrocyclic products: (1) centrally chiral macrocycles; (2) planar chiral [2.2]-paracyclophanes; (3) planar chiral macrocycles; (4) inherently chiral macrocycles.
In 2016, Wang and colleagues33 pioneered a NHC asymmetric catalysis strategy for asymmetric synthesis of centrally chiral macrocycles. As shown in Fig. 3, the enantiopure macrolacton 2 was produced successfully in modern yield with high enantioselectivity via NHC-catalyzed intramolecular asymmetric macrocyclization of 1,3-diols 1. Furthermore, the optimization of the reaction conditions indicates that key to high enantioselectivity of this method was the employment of chiral phosphoric acid (CPA) C-I as a crucial cocatalyst, proposed to stabilize the transition state via hydrogen-bonding interactions.
![]() | ||
| Fig. 3 NHC-catalytic intramolecular enantioselective macrocyclization of 1,3-diol for the synthesis of centrally chiral macrolactone. | ||
Mechanistically, the catalytic cycle involves NHC addition to the aldehyde, forming a Breslow intermediate B-I, subsequent oxidation to generate a chiral acyl azolium specie A-I, then a highly enantioselective intramolecular esterification/macrocyclization by one of the 1,3-diol hydroxyl groups onto this active carbonyl, followed by the formation of the centrally chiral macrolactone and the release of catalyst to next catalytic cycle. Overall, this work represents the first highly catalytic enantioselective synthesis of centrally chiral macrocycles via NHC activation.
![]() | ||
| Fig. 4 Asymmetric synthesis of planar chiral [2.2]paracyclophanes via NHC-catalytic kinetic resolution strategy. | ||
The reaction tolerated diverse substituents on the [2.2]paracyclophane rings and sulfonyl groups, affording both optically pure carbonitrile and sulphonamide products in excellent yields with enantioselectivities (up to >98% ee). Furthermore, the resulting planar chiral [2.2]paracyclophanes demonstrated significant synthetic utility and bioactivity. As shown in Fig. 4b, carbonitriles were derivatized to thioamides, amides, and primary amines without erosion of enantiopurity. Notably, five products exhibited superior antibacterial activity against Xanthomonas oryzae pv Oryzae (Xoo) compared to commercial pesticide thiodiazole copper (TC), highlighting their potential in agrochemical development.
:
3).
![]() | ||
| Fig. 5 Asymmetric synthesis of planar chiral [2.2]paracyclophanes via NHC-catalytic chemodivergent parallel kinetic resolution strategy. | ||
To evaluate the practicality of this protocol, a series of synthetic transformation experiments of the obtained planar chiral pyridines and lactams were performed. As illustrated in Fig. 5b, planar chiral product 13a (98% ee) underwent Ts-deprotection, chlorination, and triflation to access C-2 functionalized planar chiral [2.2]paracyclophane 15 without erosion of ee value. To further expand synthetic utility, the authors conducted a Pd-catalyzed asymmetric substitution reaction between the alkene 16 and the malonate 17 by using planar chiral [2.2]paracyclophane 15 as ligand, producing the desired chiral product 18 in 72% yield with 84% ee value. Additionally, the optically active (+)-14 displayed potent antibacterial activity against Xoo (71% inhibition at 100 μg mL−1), outperforming its ent-enantiomer and racemate, underscoring the significant role of planar chirality in bioactivity.
![]() | ||
| Fig. 6 NHC-catalyzed asymmetric desymmetrization strategy for the synthesis of planar chiral [2.2]paracyclophanes. | ||
Finally, to elucidate the reaction mechanism and origin of stereocontrol, a series of control experiments were performed. As summarized in the left of Fig. 6a, for prochiral substrate 19a, the reaction involved reversible Breslow intermediate formation (KIE = 2.8), followed by oxidation to form chiral acyl azolium intermediate, which then underwent a enantioselective esterification process to yield the major enantiomer 20a, with the reaction rate of the favoured enantiomer being 7.6 times faster than that of its ent-enantiomer. Subsequently, the ent-20a underwent the second esterification via a kinetic resolution pathway catalysed by the same catalyst with the 5.4 times reaction rate than the major product 20a. In contrast, prochiral 21 underwent irreversible Breslow formation (KIE = 0.5), enabling direct enantioselective desymmetrization (kfast3/kslow3 > 400) to afford the major enantiomer 22a in 91% yield with >99% ee value (Fig. 6b, right).
![]() | ||
| Fig. 7 NHC-catalyzed intramolecular enantioselective macrocyclization for the synthesis of indole/pyrrole-based planar chiral macrocycles. | ||
To understand the relationship between ring size and the configurational stability of the planar chiral product, the 20-membered and 21-membered macrocycles were synthesized under the optimal conditions separately. As results, they were detected with 0% ee and without planar chirality separately, clearly indicating that the macrocyclic planar chirality is highly dependent on the ring size. In addition, this method also had been successfully used to the construction of novel planar chiral macrocyclic skeletons bearing multiple stereogenic elements. A series of enantiopure macrocycles with both planar and axial chiralities were obtained with promising enantioselectivities and excellent diastereoselectivities via asymmetric kinetic resolution process under the optimal conditions.
Finally, the practicality of this method was also been demonstrated by further investigation on thermal studies and synthetic transformations of the planar chiral products. Notably, a novel chiral NHC catalyst 29 (91%)—derived from 28d, was able to catalyze the atroposelective macrocyclization of 30, providing the desired planar chiral paracyclophane 29 in good yield with high enantioselectivity (Fig. 7, 72% yield, 84% ee). Overall, this method not only represents the first realization of catalytic enantioselective access to planar chiral indole/pyrrole-based macrocycles, but also opens up a new avenue for development of NHC asymmetric catalysis.
Meanwhile, Chi and co-workers51 reported another NHC-catalyzed enantioselective macrocyclization strategy for the synthesis of planar chiral paracyclophanes from achiral bifunctional hydroxyl-aldehyde substrates 32 (Fig. 8). The plausible key intermediate TS-II is generated through the reaction of NHC with substrates 32 under oxidative conditions. This key specie simultaneously facilitates the macrocyclization via intramolecular esterification and dictates the planar chirality by differentiating the prochiral faces of aromatic ring within the substrate. Notably, the co-catalyst C-II plays a crucial role in the high yield and stereocontrol of this reaction. This method demonstrates broad scope, efficiently producing enantioenriched planar chiral paracyclophanes bearing varied functional group-containing both anas chain and aromatic plane in good-to-excellent yields (up to 82%) with high enantioselectivities (up to 92% ee).
![]() | ||
| Fig. 8 NHC-catalyzed intramolecular macrocyclization for the synthesis of planar chiral paracyclophanes (Chi's work). | ||
In addition, the resulting planar chiral macrocycles exhibit significant configurational stability, resisting racemization even at elevated temperatures (150 °C). Finally, the utility of the products is highlighted by successful derivatization via transition-metal-catalyzed cross-coupling reactions (including Suzuki, Hech, Sonogshira), showcasing their potential as versatile chiral building blocks for further elaboration.
At the same time, Zhao and co-workers52 uncovered a similar strategy for the synthesis of planar chiral paracyclophanes (Fig. 9). The reaction utilizes bifunctional hydroxy-aldehyde substrates 36, where the NHC forms an acyl azolium intermediate TS-III under oxidative conditions, enabling intramolecular esterification to form the macrocycle while controlling planar stereogenicity. Crucially, high enantioselectivities (up to 99% ee) are achieved across diverse ring sizes (15- to 18-membered) and substituents on the aromatic moiety (aryl, heteroaryl, alkynyl). Furthermore, this method is also successfully applied to synthesize planar-chiral derivatives of pharmaceuticals (including gemfibrozil, telmisartan, indomethacin, etc.) and peptides. Despite above broad substrate scopes, the type of these substrates remains limited to the ortho-alkoxy group substituted aromatic aldehydes.
![]() | ||
| Fig. 9 NHC-catalyzed intramolecular macrocyclization for the synthesis of planar chiral paracyclophanes (Zhao's work). | ||
Finally, the origin enentiocontrol in this process was also elucidated by controlled experiments and density functional theory (DFT) calculations. As shown in Fig. 9c, the results indicate that the enentioselectivity primarily stems from a stabilizing cation–π interaction between the electron-deficient acyl azolium and electron-rich OR groups (e.g., methoxy groups) within the substrate. An additional π–π interaction between the substrate's aromatic ring and the NHC's N-aryl substituent further stabilizes the favored transition state.
Recently, following their continued interest in asymmetric construction of planar chirality,53–57 the Zhao group58 disclosed the first enantio-, atrop-, and diastereoselective macrocyclization enabled by the NHC and CPA co-catalysts, yielding type III planar chiral cyclophanes featuring chiral ansa chains (Fig. 10). This strategy centers on the NHC-catalyzed desymmetrization of prochiral 1,3-diols embedded within linear precursors. The NHC generates an acyl azolium specie that initiates macrocyclization, while the CPA co-catalyst engages the diols and acyl azolium intermediates via hydrogen bonding, critically enhancing the diastereoselectivity of the process.
![]() | ||
| Fig. 10 Enantio-, atrop-, and diastereoselective intramolecular macrocyclization for the synthesis of type III planar chiral cyclophanes enabled by NHC and CPA co-catalysis. | ||
Under the optimal conditions, a range of optically active macrocycles with varying ring sizes (17- to 19-membered) and diverse aromatic ring substituents (aryl, heteroaryl) were achieved effectively via this method. In addition, the thermodynamic studies and DFT calculations demonstrate that the chiral substituent significantly increases the rotational barrier of the benzene ring within the macrocycle compared to unsubstituted analogues. Computational analysis reveals that the chiral substituent shrinks the ansa chain by compressing the bond angle, thereby hindering the conformational rotation responsible for racemization. Finally, the use of 40—derived from 39a, in the asymmetric [4 + 4]cycloadditon as the organocatlyst demonstrates the potential utility of these type III planar chiral macrocycles (Fig. 10b). Overall, this work provides a catalytic alternative to substrate-controlled diastereoselective macrocyclizations for accessing Type III cyclophanes.
![]() | ||
| Fig. 11 Pd/NHC sequentially catalyzed intermolecular atroposelective macrocyclization for the synthesis of planar chiral macrocycles. | ||
This method features broad substrate scope and high functional group tolerance. For instance, the aryl- and heteroaryl-substituted VECs delivered corresponding macrocycles (46a–46b) with excellent enantioselectivities, while alkyl-substituted VECs required longer reaction times but maintained promising stereocontrol (46c). The aryl aldehyde scope accommodated electron-donating/withdrawing groups and heterocycles. Crucially, macrocycle ring size dictates planar chirality: 17- to 19-membered rings retained chirality, while 20-membered analogues racemized due to low rotational barriers. Furthermore, racemization studies confirmed high thermal stability of these obtained planar chiral macrocycles. In addition, gram-scale synthesis and derivatizations (Fig. 11b, e.g., Sonogashira coupling, epoxidation) underscored synthetic utility. Stepwise experiments validated the sequential mechanism, distinguishing this strategy from intramolecular approaches reliant on pre-functionalized substrates.
Based on the experimental results of the substrate scope, a central insight is the critical balance between ring size and aryl substituent size for effective DKR. While 17- to 19-membered substrates underwent efficient DKR, smaller sizes (e.g., 11- to 12-membered) led to kinetic resolution due to slow racemization, and bigger size (20-membered) with flexible ansa chain (52d) abolished planar chirality. In addition, the utility of this method was also been demonstrated by the diversified derivatizations of the products. Mechanistic studies confirmed rapid substrate racemization (KIE ≈ 1) and identified NHC addition as the enantiodetermining step. As shown in Fig. 12b, the DFT calculations revealed a 3.6 kcal mol−1 energy difference between transition states TS-IV and its enantiomer (ent-TS-IV), rationalizing selectivity. Furthermore, these rigid products exhibited exceptional configurational stability (no racemization after 7 days at 110 °C in toluene), highlighting their potential as chiral scaffolds.
![]() | ||
| Fig. 13 NHC-catalyzed asymmetric desymmetrization strategy for the synthesis of enantioenriched calix[4]arens with inherent chirality. | ||
This reaction exhibits exceptional scope and practicality. Lower-rim alkyl groups and diverse phenols—including electron-rich/deficient arenes, natural products (e.g., eugenol, capsaicin), and pharmaceuticals (e.g., estrone, ezetimibe)—delivered products with promising ee values. Gram-scale synthesis (55a, 79% yield, 98% ee) and post-functionalizations (Wittig olefination, reduction, Pinnick oxidation) afforded versatile chiral building blocks without erosion of optical purity (Fig. 13b, 56–58). Furthermore, the carboxylic acid derivative 58 served as a precursor for hybrid organocatalysts (Fig. 13c), which facilitated an aqueous aldol reaction (30% yield, 92% ee).
Mechanistic studies using deuterium labeling and KIE measurements (Fig. 13d, KIE = 1.5) indicate a rate-determining 1,2-proton shift in Breslow intermediate formation. Control experiments confirme desymmetrization governs enantioselectivity (for model reaction: kfast1/kslow1 = 43
:
1), not kinetic resolution, and a postulated mechanistic pathway is presented in Fig. 13e. Overall, this metal-free protocol offers a valuable platform for asymmetric molecular recognition and catalysis.
![]() | ||
| Fig. 14 NHC-catalyzed enantioselective (dynamic) kinetic resolution for synthesis of macrocycles with inherent chirality. | ||
Meanwhile, the key optimizations reveal elevated temperature accelerated substrate racemization (critical for DKR process), while diluted conditions enhanced enantiocontrol. The DKR reaction demonstrates broad scope across aromatic/heteroaromatic aldehydes and alkyl aldehydes, with the latter showing superior compatibility (Fig. 14a, 65a–65d). On the other hands, for configurationally stable macrocycles (Fig. 14b, 63e–63g), KR process affords enantiomerically enriched products (65e–65g) and recovered starting materials with high selectivity factors (s up to 106).
Furthermore, the synthetic utility of this protocol has been demonstrated by the gram-scale synthesis of 65c and a range of its post-functionalization (such as sonogashira coupling). Finally, the origins of the high enantioselectivity for the DKR process have been elucidated by DFT calculations. Overall, this work pioneers NHC-catalytic (D)KR strategies for the synthesis of inherently chiral macrocycles, offering robust access to enantiopure calix[4]heteroarenes, and providing a valuable platform for supramolecular chemisty.
Despite these progresses, NHC-catalytic enantioselective synthesis of chiral macrocycles is still in its infancy, and there remains several challenges and opportunities merit exploration: (1) design and construction of novel chiral macrocycles with multiple stereogenic elements—such as with both axial and planar chiralities, or with heteroatom stereogenic centers—including nitrogen-, and phosphorus-stereocenters.71 A broad range of natural products and bioactive compounds feature multiple stereogenic elements, which have significant influences on their biological and physical properties.72 (2) Exploration of new catalysts and novel catalytic mechanisms, which is highly desirable for advancing the field and is particularly important for improving stereocontrol. (3) Development of new strategies for NHC catalytic catalysis—such as integration with photochemical, electrochemical, or biocatalytic approaches. Exploring greener catalytic schemes would address the scalability and sustainability issues, and will be crucial for future industrial transformations. Overall, this review aims to offer a panoramic overview of research progress, challenges, and opportunities in catalytic enantioselective synthesis of chiral macrocycles enabled by NHC catalysis for potential researchers. We tentatively conclude that further exploration will foster sustainable developments in this burgeoning field.
| This journal is © The Royal Society of Chemistry 2025 |