Next Article in Journal
Construction of Hollow TiO2/ZnS Heterojunction Photocatalysts for Highly Enhanced Photodegradation of Tetracycline Hydrochloride
Previous Article in Journal
Osteogenic Differentiation in Chitosan-Based Scaffolds via P28 and VEGF Delivery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes

Key Laboratory of Green Chemistry & Technology of Ministry of Education, College of Chemistry, Sichuan University, 29 Wangjiang Road, Chengdu 610064, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(17), 3646; https://doi.org/10.3390/molecules30173646
Submission received: 9 August 2025 / Revised: 1 September 2025 / Accepted: 5 September 2025 / Published: 7 September 2025
(This article belongs to the Section Organic Chemistry)

Abstract

Macrocyclic arenes are rich-electron macrocycles bridged by methylene or methyl groups from aromatic rings substituted by hydroxyl or alkoxy groups. It has attracted great interest in host–guest chemistry and supramolecular self-assembly due to its clear cavity, adjustable structure and multifunctional binding ability. In particular, nitrogen-containing macrocyclic arenes including (hetero) aromatic moieties—constructed from building blocks such as pyrrole, carbazole, phenothiazine, and imidazole—have undergone rapid development, yielding a new generation of functional macrocycles, including calix[4]carbazoles, Tröger’s base-derived macrocycles, and phenothiazine-based architectures. These nitrogen-functionalized macrocycles feature rich chemical derivatization potential, unique structural and host–guest characteristics, and exceptional photophysical properties. They show great promise in molecular recognition, selective adsorption and separation, and the development of advanced functional materials. This review summarizes recent advances in the design, synthesis, and application of nitrogen-containing macrocyclic arenes, with a particular focus on structure–property relationships and emerging functions.

1. Introduction

Supramolecular chemistry studies molecular recognition and self-assembly processes governed by non-covalent interactions, encompassing a wide range of disciplines including chemistry, biology, and materials science [1,2,3]. Within this domain, macrocyclic hosts have served as essential structural scaffolds [4,5,6]. Since the seminal discovery of crown ethers, the design and synthesis of macrocyclic hosts have remained central to the advancement of this field [7]. Despite the extensive development of various macrocyclic frameworks [8,9,10,11,12,13,14,15], the exploration of novel macrocyclic systems continues to pose a formidable challenge. Macrocyclic arenes are rich-electron macrocycles bridged by methylene or methyl groups from aromatic rings substituted by hydroxyl or alkoxy groups [16,17,18]. These compounds typically feature electron-rich aromatic moieties connected via sp3-hybridized carbon atoms, conferring relatively rigid frameworks, unique electronic properties, and conformational dynamics. Recent advances in synthetic methodologies have enabled the construction of increasingly sophisticated macrocyclic arenes with well-defined geometries and functional potential [16,19,20,21,22,23,24]. Macrocyclic arenes are primarily prepared by macrocyclization with repeat aromatic units, such as phenyl/biphenyl-based derivatives (e.g., pillar[n]arenes, biphen[n]arenes, tiara[5]arenes) [25,26,27,28,29,30] and fused-ring compounds (e.g., prism[n]arenes, pagoda[n]arenes, naphthotubes) [31,32,33,34,35,36]. These macrocyclic hosts expand the research scope in supramolecular chemistry and provide new molecular platforms for the development of functional materials.
Compared to oxygen-containing analogs [16,37], nitrogen atoms exhibit superior functionalization versatility, enabling formation of amides, polysubstituted amines, quaternary ammonium salts, and N-oxides, while their variable valence states and stimulus responsiveness impart unique structural flexibility and functional diversity to macrocyclic arenes [38,39]. These characteristics establish nitrogen incorporation as a critical research focus in contemporary macrocyclic molecular engineering [40,41]. The most fundamental synthetic approach involves direct incorporation of nitrogenous units into the macrocyclic backbone. Friedel–Crafts alkylation efficiently links nitrogenous fragments to macrocyclic frameworks, while amide condensation routes show significant promise for ion-recognition applications, yielding macrocyclic arenes with distinctive peptide-bond architectures. Complementary strategies include Mannich condensation, transition metal-catalyzed couplings, coordination-directed assembly, and nucleophilic substitutions, each offering distinct features: Friedel–Crafts alkylation bridges electron-rich arenes; Mannich reactions require specific aromatic amine/carbonyl components; amide condensations usually employ amino acid derivatives; nucleophilic substitutions utilize halocarbon substrates. These synthetic advances enable macrocyclic arenes to demonstrate superior performance in molecular recognition, selective adsorption, and functional materials science. In addition, there are some methods for the synthesis of nitrogen-containing macrocyclic compounds [42,43], such as the synthesis of macrocyclic compounds containing 1,10-phenanthroline [44,45]. Asymmetric synthesis methods are also often used for nitrogen-containing compounds [46,47]. Finally, some macrocyclic compounds without catalysis [48,49,50,51] and light-controlled synthesis [52] also play an important role.
This review mainly summarizes the research progress of nitrogen-containing macrocyclic aromatic hydrocarbons in the past five years. Firstly, the synthesis strategies are classified according to the structural position of the nitrogen atom, and the system with nitrogen as the bridging site and the system with nitrogen in the non-bridging position are distinguished. It includes three types: Friedel–Crafts alkylation reaction, nucleophilic reaction, and Mannich reaction. Subsequently, the structural features and physicochemical properties of nitrogen-functionalized macrocycles are systematically compared with those of traditional phenol-based macrocycles. Finally, the current progress in their applications—particularly in molecular recognition, selective adsorption and separation, and the development of functional materials—is summarized. This review aims to offer both theoretical insight and practical guidance for the rational design of next-generation macrocyclic arene systems.

2. Synthesis of Nitrogen-Containing Macrocyclic Arenes

Despite substantial advances in macrocyclic arene research over the past decade, current synthetic systems predominantly utilize benzenoid cores with hydroxy/alkoxy substituents [16]. Research on extended π-systems (e.g., biphenyl, acene) remains underdeveloped, primarily hindered by availability of synthetic monomers, synthetic challenges, poor solubility, and reaction site selectivity issues. Nitrogenous compounds are pivotal in medicinal chemistry and functional materials, owing to structural diversity and facile functionalization. These properties render them ideal precursors for macrocycle construction, enabling development of structurally unique macrocyclic arenes. Based on nitrogen-atom positioning within the macrocyclic framework, these systems classify into two categories: (1) nitrogen is a bridging atom of aromatic subunits (e.g., Tröger’s base-functionalized macrocycles [53], phenothiazine-based macrocycles [54]); (2) nitrogen is not the bridging atom, but serves as a heteroatom within a heterocycle, such as calix[n]pyrroles [55,56], calix[n]carbazoles [57], cyclocarbazoles [58,59,60,61], phenothiazine derivatives [62], imidazole-based macrocycles [63], and indole-containing architectures. Macrocyclic arenes featuring bridging nitrogen atoms are often synthesized through aniline derivatives. These nitrogen centers exhibit versatile reactivity for further functionalization. Macrocyclic arenes having non-bridging nitrogen are typically constructed via Friedel–Crafts alkylation, which generally have component units of relatively large p plane and could show unique photophysical properties.

2.1. Macrocyclic Arenes with Non-Bridging Nitrogen

Certain macrocyclic arenes incorporate nitrogen atoms at non-bridging sites. Their synthetic approach closely resembles that of phenol-based macrocyclic arenes, involving one-pot condensation of electron-rich aromatics with paraformaldehyde catalyzed by Brønsted acids, particularly for constructing macrocycles comprising identical subunits.
Research on nitrogen-containing macrocycles originated in the late 19th century. In 1886, Baeyer and co-workers described the HCl-catalyzed condensation of pyrrole with acetone (Figure 1a) [64]. As a representative system, calix[4]pyrrole 1 mediates halide anion recognition via N−H···X hydrogen bonding, while simultaneously binding ion pairs through cation–π interactions [55]. Pioneered by the research groups of Sessler and co-workers, derivatives based on the calix[4]pyrrole scaffold have been extensively developed. Despite a century of research on calix[4]pyrrole, its straightforward synthesis, structural tunability, and superior host–guest properties sustain its prominence in modern supramolecular chemistry.
Yang and co-workers achieved a breakthrough in carbazole macrocycle synthesis, pioneering the methylene-bridged carbazole macrocycle (Figure 1b) [57]. Employing carbazole monomer 2 and paraformaldehyde as starting materials with FeCl3·6H2O as catalyst, they achieved one-step cyclization in CH2Cl2 at ambient temperature. This reaction afforded product 3, predominantly comprising calix[3]carbazole (20% yield), with minor amounts of calix[4–6]carbazole byproducts.
Wang and co-workers efficiently synthesized a series of novel nitrogen- or oxygen-bridged calix[2]arene[2]triazine macrocycles (Figure 2) [65]. Their stepwise approach involved the nucleophilic substitution of a triazine derivative 4 to form trimeric intermediates 58, followed by a [3 + 1] cyclization that yielded macrocycles 917. X-ray analysis confirmed a 1,3-alternating conformation in the solid state, with the bridging heteroatoms arranged coplanarly. The researchers found that these heteroatoms critically modulate the cavity geometries through electronic, conjugative, and steric effects, which, in turn, dictate distinct supramolecular assemblies. This work significantly expanded the known diversity of heteroatom-bridged calixarene macrocycles.
The same group reported the synthesis of O6-corona[3]arene[3]tetrazine macrocycles via aromatic nucleophilic substitution (Figure 3) [66]. Macrocycles 2734 (Figure 3a) were synthesized in one step through mild nucleophilic substitution between 1,4-dihydroxybenzene derivatives 1825 and 3,6-dichlorotetrazine 26, with yields ranging from 25% to 80%. Single-crystal analyses of 27 and 28 revealed hexagonal macrocycles featuring nearly planar arrangements of oxygen bridges and tetrazine rings (Figure 3b). Variable-temperature NMR spectroscopy indicated conformational changes between solid and solution states. Owing to their electron-deficient character, these macrocycles exhibit chloride binding via anion–π interactions, as demonstrated by 33 (Figure 3c). This study demonstrates that aromatic nucleophilic substitution is a viable strategy for macrocyclization, thereby broadening the structural scope of macrocyclic chemistry.
Esser and co-workers reported a novel methylene-bridged macrocycle derived from phenothiazine (Figure 4) [62]. The synthesis was carried out using N-substituted phenothiazine derivatives and paraformaldehyde in 1,2-dichloroethane, with trifluoroacetic acid as the catalyst at 80 °C for 120 h. When alkyl-substituted monomer 35 was employed, both calix[3]phenothiazine 37 (15% yield) and calix[4]phenothiazine 38 (4% yield) were obtained. In contrast, the p-methylphenyl-substituted monomer 36 selectively produced calix[3]phenothiazine 39 in 30% yield. X-ray crystallographic analysis confirmed a non-pyramidal AAB configuration for 39. Quantum chemical calculations indicated that the cone conformation (AAA) is 1.2–1.3 kcal/mol higher in energy than the observed non-cone form.
Yang and co-workers reported bisindole[3]arenes—a novel class of indole-based macrocycles (Figure 5) [67]. The initial direct cyclization of indole with paraformaldehyde afforded the methylene-bridged cyclic trimer 40, characterized by (i) a small cavity and (ii) methylene linkages at the 2′- and 3′-positions of the indole rings (Figure 5a). An optimized one-pot condensation of bisindole derivatives with paraformaldehyde in CH2Cl2, catalyzed by AlCl3 at room temperature, yielded target compounds 41a (4.8%) and 41b (10.3%) (Figure 5b). X-ray crystallographic analysis confirmed that all bisindole[3]arenes possess C3v symmetry. Notably, 41a displays alternating up/down orientations of the indole units relative to the macrocyclic plane, whereas the methyl-substituted 41b adopts a petal-like conformation (Figure 5c).
Shi and co-workers reported a series of DPA[n] macrocycles (4248) derived from methyldiphenylamine units, representing a significant advancement in the synthesis and self-assembly of aniline-based macrocycles (Figure 6a) [68]. These compounds were efficiently synthesized via a one-pot procedure and exhibit structural features distinct from classical aniline derivatives. All benzene rings adopt either coplanar or perpendicular conformations. X-ray crystallographic analysis revealed that macrocycle 42 adopts a planar structure stabilized by intermolecular edge-to-face π–π interactions. In contrast, 43 displays a V-shaped conformation with offset face-to-face π–π stacking within its molecular assemblies (Figure 6b,c). This stacking mode facilitates hierarchical self-assembly: (i) 42 and 43 form columnar structures; (ii) van der Waals interactions at the benzene ring edges promote staggered layering; and (iii) ultimately, three-dimensional porous frameworks are formed.
Yang and co-workers synthesized novel aniline-based macrocycles 4951 via a one-pot condensation of diphenylamine with paraformaldehyde (Figure 7a) [69]. Structural analysis revealed that An[2] comprises six benzene rings adopting distinct conformations, with four in coplanar orientation, forming a semi-open cavity of approximately 9 Å in diameter.
Building upon earlier studies of aniline macrocycles, significant progress has been made in the cyclization of triphenylamine derivatives. Zhao and co-workers employed bromotriphenylamine as a precursor to synthesize C3-symmetric TPA[3] macrocycles (Figure 7b,c). Brominated derivatives improved yields compared to conventional substrates, while halogen substituents facilitated further post-functionalization. The resulting macrocyclic framework consists of six benzene rings connected through alternating methylene and nitrogen bridges, forming a cavity (~10 Å) comparable in size to that of pillar[6]arene [70].

2.2. Macrocyclic Arenes with Nitrogen-Bridging Macrocycles

2.2.1. Nucleophilic Reaction

Research on paraquat-based macrocycles began in the 1980s. Hünig and co-workers synthesized paraquat macrocycles 5662 (Figure 8a) [71], followed by systematic investigations of their radical cation properties [72]. Stoddart and co-workers prepared cyclobis(paraquat-p-phenylene) (CBPQT4+, 55; Figure 8b), demonstrating its host–guest complexation with electron-rich species. This finding laid the foundation for the development of mechanically interlocked molecules. Despite over three decades of study, CBPQT4+ remains a focus of active research due to its accessible synthesis and unique electronic characteristics. Recent efforts have yielded structural derivatives (e.g., 5658), thereby broadening the scope of supramolecular applications.
Chun and co-workers reported the first calixarene-like system derived from imidazolium salts, marking a significant advance in heterocyclic macrocycle chemistry (Figure 9a) [63]. Using methylene-bridged imidazole 63 as a precursor, they obtained calix[4]imidazolium 64 and calix[5]imidazolium 65 via nucleophilic substitution. This strategy overcomes the structural uniformity of traditional calixarenes, offering greater design versatility. Subsequently, Szumna and co-workers reported the synthesis of pillar[4]pyridinium macrocycles (Figure 9b), contributing to the development of pyridinium-based systems [73]. They employed pyridinium salt 66 as the precursor, achieving macrocycle 67 in 50% yield through an efficient one-step nucleophilic substitution. Notably, the choice of counterion significantly influenced the solubility of the resulting macrocycles. X-ray crystallographic analysis revealed that pillar[4]pyridinium 67 adopts a square, box-like conformation (side length: 6.2 Å; cavity diameter: 3 Å) (Figure 9c,d), a preorganized geometry well-suited for applications in molecular recognition and self-assembly.
Kim and co-workers synthesized triangular porphyrin macrocycle P3L3 by imine condensation reaction using cis-diaminophenyl porphyrin and isophthalaldehyde linker as building blocks (Figure 10a) [74]. P3L3 molecules have two conformations A and B with different cis–trans arrangements in the solid state, and stack with each other through π-π and CH-π interactions to form nanotubes with a diameter of 1.7 nm (Figure 10b). Gas adsorption experiments and powder X-ray diffraction analysis show that the pores in the P3L3 crystal are permanent and will not collapse due to the reversibility of the imine bond. By adjusting the angle of the bending connector, some columns of large ring structures with different sizes can also be obtained.

2.2.2. Mannich-Type Macrocyclization

Tröger’s base (TB), a structurally distinctive class of organic compounds, was first reported by Carl Julius Ludwig Tröger in 1887 [75,76]. Research in this field has spanned nearly 140 years, with a notable resurgence over the past two decades—approximately 60% of all related publications have appeared during this period [77,78]. Structurally, TB features a characteristic V-shaped geometry defined by (1) methylene-bridged aromatic rings, (2) primary amine nitrogen atoms located at the V-shaped termini, and (3) a rigid, highly symmetric framework. This unique structure imparts three key properties: (i) intrinsic chirality, with enantiopure forms accessible via chiral HPLC resolution; (ii) utility as chiral building blocks for macrocyclic systems; and (iii) pronounced steric rigidity. The established synthetic route (Figure 11) follows a Mannich-type reaction mechanism: (1) condensation of aniline with paraformaldehyde forms imine intermediate 68; (2) electrophilic addition of 68 to aniline generates diamine adduct 69; (3) condensation with paraformaldehyde constructs the nitrogen bridge at the amine sites of 69; and (4) a final electrophilic addition completes assembly of the TB scaffold. These attributes make TB derivatives promising candidates for supramolecular chemistry and functional materials development.
Wang and co-workers reported the synthesis of TB-derived macrocycles (Figure 12a) [79]. To address the longstanding challenges of low yields and poor stereoselectivity in TB scaffold construction, the authors employed ethylene glycol-bridged phenylenediamine precursors of varying chain lengths to synthesize macrocycles 75a75c via a one-step TB reaction. The length of the ethylene glycol linker was found to play a critical role in modulating macrocyclic chirality: (1) short chains (75a) exclusively yielded homochiral R4N/S4N isomers; (2) medium chains (75b) produced a mixture of homochiral (R4N/S4N) and heterochiral (R2NS2N) isomers; and (3) long chains (75c) generated only heterochiral configurations (Figure 12b). Although the mechanistic basis for this chain length–chirality relationship remains to be fully elucidated, this work provides a valuable experimental foundation for the stereoselective synthesis of TB-based macrocycles and advances the development of chiral supramolecular systems.
Yang and co-workers reported a synthetic advancement in the construction of Tröger base (TB)-derived macrocycles [53]. Utilizing diphenylamine precursors, compounds 76 and 77 were synthesized via single-step TB formation reactions (Figure 13a), establishing the TBn (n = 2–3) macrocyclic series. The resulting methylene-bridged TB2 and TB3 macrocycles were obtained as racemic mixtures, each consisting of homochiral configurations (R4N/S4N and R6N/S6N, respectively). Enantiopure forms were successfully isolated via chiral HPLC. Subsequent oxidation of the resolved enantiomers with m-CPBA afforded water-soluble N-oxides 78 and 79, with complete retention of the original TB chirality (Figure 13b,c).
More recently, Yang and co-workers developed a one-pot synthesis of phenacetin-derived[3]arene macrocycle 80 via a Mannich-type condensation (Figure 14) [54]. This acid-catalyzed reaction between phenacetin and paraformaldehyde establishes a novel cyclization pathway, forming methylene bridges between aromatic amines and aryl carbons (NAm-CH2-CAr). Compound 80 exhibits unique structural features, tunable chemical functionality, and molecular recognition capabilities. Phenacetin, a commercially available p-aminophenol derivative and widely used antipyretic/analgesic, served as the starting material. Under BF3·OEt2 catalysis, the one-pot reaction afforded 80, with uncyclized oligomers as the primary byproducts. Homologs of phenacetin underwent analogous cyclization to yield compounds 8182, enabling gram-scale production. Subsequent reduction, dealkylation, or hydrolysis of 80 afforded derivatives 8385, offering multiple sites for further functionalization.

3. Applications of N-Containing Macrocyclic Arenes

3.1. Molecular Recognition

3.1.1. Recognition of Cationic Guests

Nitrogen-containing macrocyclic arenes exhibit selective recognition toward electron-deficient guests, particularly cations, due to their electron-rich aromatic cavities. A representative example is provided by the Motuo and co-workers: their conical calix[3]phenoxazine 88a (Figure 15) forms a stable 1:1 complex with the cationic guest G1, with a binding constant of (1.66 ± 0.15) × 103 M−1. This result highlights the superior cation-binding affinity of such macrocyclic systems [80].
Calix[3]carbazole 41 (Figure 16) exhibits distinctive molecular recognition behavior. Its enlarged, semi-open, π-electron-rich cavity selectively binds tetraethylammonium (TEA+) through cation–π interactions, affording stable 1:1 complexes. In contrast, its binding affinity for larger guests such as tetrapropyl- and tetrabutylammonium is significantly diminished, highlighting its stringent size selectivity [57].
Phenacetin[3]arene 80 [54] exhibits a pronounced affinity toward primary ammonium salts. Host–guest complexes were characterized via 1H NMR, 13C NMR, ITC, circular dichroism, and DFT calculations (Figure 17). Compound 80 forms 1:2 complexes, wherein the initial 1:1 association is governed by triple hydrogen bonding between the three carbonyl groups of 80 and the NH groups of the guest, yielding binding constants (K1) on the order of 105~106 M−1, as determined by ITC. The resulting 1:2 complex displays allosteric behavior: Rim A stabilizes the guest through two C=O···H–N hydrogen bonds and one NH···π interaction, whereas rim B engages one C=O···H–N bond and two NH···π interactions. The stepwise binding constant for the second guest (K2) is approximately 103 M−1.

3.1.2. Enantioselective Recognition of Chiral Molecules

Chiral recognition—a fundamental molecular recognition process in biological systems [14,15,29,81,82,83,84,85,86,87,88,89]—holds significant practical value in asymmetric catalysis [90,91], chiral drug development [92], and the design of functional materials [93]. Macrocyclic arenes serve as ideal platforms for investigating stereoselectivity due to their tunable cavity geometries and well-defined stereochemical environments. Recent advances in macrocyclic arene receptors with rigid frameworks or internally modified cavities have yielded systems with exceptional chiral recognition capabilities. A notable example is the water-soluble nitrogen-bridged macrocycles 7879 (Figure 18a), reported by Yang and co-workers. These macrocycles not only preserve the chiral configuration of the tert-butyl substituents but also exhibit pronounced enantioselectivity toward chiral guests in aqueous media, attributed to their structural rigidity. Remarkably, the selectivity coefficient toward naphthyl phosphate guest G11 reaches 41.0—the highest reported for macrocyclic arenes to date (Figure 18b). These results offer valuable guidance for the development of next-generation chiral separation materials and asymmetric catalysts [53].

3.2. Adsorption and Separation

Iodine is a common nuclear contaminant of significant concern. The radioisotope 129I, characterized by its long half-life (1.57 × 107 years), high environmental mobility, and toxicity [94], poses persistent ecological risks. In contrast, 131I, though short-lived (t1/2 = 8.02 days), presents serious radiological hazards due to its propensity to accumulate in the thyroid during human metabolism [95,96]. Consequently, efficient radioactive iodine capture is vital, particularly in the context of radiological emergencies. Yang and co-workers conducted a systematic study on iodine vapor adsorption using bisindole[3]arene derivatives 41a41c (Figure 19). These macrocyclic powders exhibited pronounced chromogenic responses upon iodine uptake. Among them, compound 41c showed the highest adsorption capacity (5.12 g/g), outperforming 41a (4.49 g/g) and 41b (4.73 g/g). Mechanistic investigations revealed that these initially non-porous materials undergo substantial structural reorganization during adsorption, with the exceptional iodine capture efficiency attributed to intra- and intermolecular synergistic effects. Furthermore, fluorescence quenching observed in 41c confirms charge transfer interactions between the indole moieties and iodine. These findings provide key molecular design insights for the development of next-generation iodine capture materials [67].
In addition, macrocyclic diphenylamine 4951 exhibited outstanding iodine adsorption performance (Figure 20), with a maximum uptake capacity of 7.35 g/g in both gas-phase and solution-phase environments [69]. The study revealed a distinctive reverse dissolution–adsorption mechanism: upon exposure to iodine, the crystalline forms of 4951 undergo a phase transition to amorphous states, substantially enhancing their iodine-loading capacity. Notably, iodine can be reversibly desorbed through thermal treatment or exposure to organic solvents, and the materials retain stable adsorption performance over multiple cycles. Despite their limited apparent specific surface areas, the electron-rich nitrogen-containing frameworks of 4951 facilitate stable iodine capture via efficient electron transfer. These findings offer a novel strategy for the molecular design of advanced adsorption materials.
Zeng and co-workers designed novel electron-deficient pillar[6]arene derivatives featuring a biphenyltetracarboxylic diimide-extended structure that forms a nanoscale electron-deficient cavity [97]. This macrocycle demonstrated exceptional separation performance as a gas chromatography stationary phase, efficiently resolving benzene/cyclohexane and toluene/methylcyclohexane mixtures. Compared to conventional capillary columns, this packed column offers distinct advantages: simplified preparation, reduced column length, lower separation temperatures (e.g., 40 °C), and significantly shorter retention times. These properties provide a promising alternative for chromatographic separation technologies.

3.3. Novel Photophysical Properties

Significant progress has recently been made in the development of luminescent materials based on novel N-containing macrocyclic arenes, particularly within the realm of thermally activated delayed fluorescence (TADF) systems. Chen and co-workers pioneered a supramolecular TADF (STADF) strategy [98], employing host–guest charge transfer between electron-donating calix[3]acridine 89 and electron-accepting 1,2-dicyanobenzene (G12) to achieve efficient STADF emission (Figure 21a). This co-crystal system achieves a remarkable PLQY of 70%—the highest reported for intermolecular donor–acceptor TADF materials. Building on this work, the team further developed calix[3]acridine 91 (Figure 21b), which possesses intrinsic TADF properties by incorporating acridine donor and s-triphenyltriazine acceptor units [99]. Compound 91 exhibits an exceptionally small ΔEST (12 meV) and a high solution PLQY (80% in toluene), demonstrating outstanding potential as an efficient TADF emitter.
Liu and co-workers recently developed an amorphous supramolecular system comprising calix[3]phenothiazine 92 and dicyanobenzene derivatives G12/G13 (Figure 22a), exhibiting unique organic room-temperature phosphorescence (RTP) properties [100]. These complexes demonstrate unprecedented RTP behavior arising solely from host–guest complexes. Upon 430 nm excitation, C–H···π interactions facilitate the formation of dense supramolecular networks, resulting in phosphorescence emissions at 566 nm (92/G12) and 524 nm (92/G13). In contrast, the pristine crystals of 92 display only fluorescence at 508 nm. Mechanistic investigations revealed that G12 and G13 serve as “molecular rivets,” stabilizing the conformation of 92 and restricting molecular vibrations, thereby reducing non-radiative decay pathways. This structural stabilization leads to high phosphorescence quantum yields ranging from 20.8% to 36.8%. Capitalizing on this guest-activated RTP phenomenon, the authors further demonstrated practical applications, including multicolor quick-response (QR) codes and information encryption (Figure 22b).
Recently, Chen and co-workers developed novel dual-emission TADF polymer systems by solution-processing calix[3]acridine-functionalized polymer 93-P with electron-accepting guests G13G16 of varying electron affinity (Figure 23a) [101]. The TADF properties originate from through-space charge transfer (TSCT) between the macrocyclic donor in the polymer and electron-accepting guests. Precise modulation of guest electron-accepting strength enables tunable emission colors and achieves photoluminescence quantum yields up to 40%. These materials exhibit superior processability, tunable luminescence, and excellent mechanical stability. Leveraging these advantages, the team applied the systems to information security, developing substrate-independent anti-counterfeiting tags and smartphone-readable QR codes (Figure 23b). This provides novel solutions for advanced information encryption.

4. Conclusions and Perspectives

Recent advances in macrocyclic arenes research have significantly advanced supramolecular chemistry [68]. In particular, nitrogen-incorporated macrocycles have garnered broad interdisciplinary interest due to their distinctive host–guest recognition capabilities [53,54]. This perspective categorizes recent synthetic strategies and applications of N-containing macrocyclic arenes based on their macrocyclization approaches: (1) nitrogen-bridging macrocycles and (2) non-nitrogen-bridging macrocycles. These developments not only expand the architectural diversity of macrocyclic arenes but also establish new design paradigms relevant to chemistry, biology, and materials science.
Despite notable progress in the chemistry of macrocyclic arenes [16], the design, synthesis, and application of structurally novel N-containing macrocyclic arenes continue to present fundamental challenges. Owing to their unique structural features, these macrocycles hold great promise for the development of N-containing functional molecules. One-pot condensation of polycyclic aromatic or heteroaromatic units offers a promising route to constructing N-containing macrocyclic arenes due to its synthetic efficiency. However, the presence of active hydrogens in primary and secondary amines often leads to side reactions, complicating the reaction landscape. Tertiary amines are frequently employed as macrocyclization units. Furthermore, the judicious selection of building blocks, in combination with the use of directing groups, significantly enhances synthetic efficiency. For unmodified polycyclic aromatic systems, post-synthetic modification of macrocyclic frameworks provides an alternative and complementary synthetic strategy.
Macrocyclic arenes serve as key supramolecular functional materials owing to their tunable cavity dimensions and superior electronic properties. Through rational molecular design, these macrocycles enable highly efficient cation recognition, offering promising applications in chiral separations, environmental remediation, and optoelectronic devices. Nonetheless, several critical challenges must be addressed to achieve practical, high-performance materials. In the context of chiral separations, macrocyclic arene-based platforms offer promising solutions for the enantioselective interaction of pharmaceuticals and bioactive molecules. While successful chiral recognition examples have been reported, systems that simultaneously exhibit high enantioselectivity, strong affinity, and long-term cycling stability remain limited. For environmental applications, these macrocycles demonstrate strong affinity for aqueous organic pollutants and heavy metal ions via specific host–guest interactions. However, their implementation in real-world wastewater treatment requires further validation. As optoelectronic materials, macrocyclic arenes suppress non-radiative decay through their rigid skeletal frameworks, resulting in enhanced fluorescence quantum yields. Moreover, host–guest complexation modulates inter- and intramolecular charge transfer processes, allowing precise control over luminescence color and enabling long-lived emission (e.g., RTP). These properties support a wide range of applications in sensors, organic light-emitting diodes (OLEDs), and advanced anti-counterfeiting technologies. Nevertheless, the field still faces major challenges in conducting systematic mechanistic studies and improving material stability.
Although the practicality of macrocyclic aromatic hydrocarbons has been proven, their functionalization, scalable synthesis, and conversion applications still face significant obstacles. Future research should combine computational simulation, rational molecular design, and multidisciplinary approaches to accelerate the transition from basic research to practical applications. This rapidly evolving field offers abundant opportunities at the interfaces of supramolecular chemistry, life sciences, and advanced materials. Overcoming existing technical limitations and harnessing the full potential of macrocyclic arenes will pave the way for transformative advances in both fundamental science and technological innovation.

Author Contributions

Conceptualization: J.H. and C.Y.; investigation: J.H. and C.Y.; resources, J.H. and C.Y.; data curation: J.H. and C.Y.; writing—original draft preparation, J.H.; writing—review and editing: J.H. and C.Y.; visualization, J.H. and C.Y.; supervision: W.W. and C.Y.; project administration: W.W. and C.Y.; funding acquisition: W.W. and C.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 22471182, 22271201, 22422108, 22171194), the Science and Technology Department of Sichuan Province (No. 2025ZNSFSC0125).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cram, D.J. The Design of Molecular Hosts, Guests, and Their Complexes (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 1009–1020. [Google Scholar] [CrossRef]
  2. Lehn, J.-M. Supramolecular Chemistry—Scope and Perspectives Molecules, Supermolecules, and Molecular Devices (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 89–112. [Google Scholar] [CrossRef]
  3. Pedersen, C.J. The Discovery of Crown Ethers (Noble Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 1021–1027. [Google Scholar] [CrossRef]
  4. Feringa, B.L. The Art of Building Small: From Molecular Switches to Motors (Nobel Lecture). Angew. Chem. Int. Ed. 2017, 56, 11060–11078. [Google Scholar] [CrossRef]
  5. Sauvage, J.-P. From Chemical Topology to Molecular Machines (Nobel Lecture). Angew. Chem. Int. Ed. 2017, 56, 11080–11093. [Google Scholar] [CrossRef]
  6. Stoddart, J.F. Mechanically Interlocked Molecules (MIMs)—Molecular Shuttles, Switches, and Machines (Nobel Lecture). Angew. Chem. Int. Ed. 2017, 56, 11094–11125. [Google Scholar] [CrossRef]
  7. Pedersen, C.J. Cyclic polyethers and their complexes with metal salts. J. Am. Chem. Soc. 1967, 89, 7017–7036. [Google Scholar] [CrossRef]
  8. Gutsche, C.D.; Muthukrishnan, R. Calixarenes. 1. Analysis of the product mixtures produced by the base-catalyzed condensation of formaldehyde with para-substituted phenols. J. Org. Chem. 1978, 43, 4905–4906. [Google Scholar] [CrossRef]
  9. Freeman, W.A.; Mock, W.L.; Shih, N.Y. Cucurbituril. J. Am. Chem. Soc. 1981, 103, 7367–7368. [Google Scholar] [CrossRef]
  10. Kim, K.; Selvapalam, N.; Ko, Y.H.; Park, K.M.; Kim, D.; Kim, J. Functionalized cucurbiturils and their applications. Chem. Soc. Rev. 2007, 36, 267–279. [Google Scholar] [CrossRef]
  11. Rekharsky, M.V.; Inoue, Y. Complexation Thermodynamics of Cyclodextrins. Chem. Rev. 1998, 98, 1875–1918. [Google Scholar] [CrossRef]
  12. Harada, A. Cyclodextrin-Based Molecular Machines. Acc. Chem. Res. 2001, 34, 456–464. [Google Scholar] [CrossRef] [PubMed]
  13. Ogoshi, T.; Kanai, S.; Fujinami, S.; Yamagishi, T.-A.; Nakamoto, Y. para-Bridged Symmetrical Pillar[5]arenes: Their Lewis Acid Catalyzed Synthesis and Host–Guest Property. J. Am. Chem. Soc. 2008, 130, 5022–5023. [Google Scholar] [CrossRef]
  14. Zhao, T.; Yi, J.; Liu, C.; Liang, X.; Shen, Y.; Wei, L.; Xie, X.; Wu, W.; Yang, C. “First Come, First Served” and Threshold Effects in a Central-to-Planar-to-Helical Hierarchical Chiral Induction. Angew. Chem. Int. Ed. 2023, 62, e202302232. [Google Scholar] [CrossRef] [PubMed]
  15. Ji, J.; Wei, X.; Wu, W.; Fan, C.; Zhou, D.; Kanagaraj, K.; Cheng, G.; Luo, K.; Meng, X.-G.; Yang, C. The More the Slower: Self-Inhibition in Supramolecular Chirality Induction, Memory, Erasure, and Reversion. J. Am. Chem. Soc. 2022, 144, 1455–1463. [Google Scholar] [CrossRef]
  16. Han, X.-N.; Han, Y.; Chen, C.-F. Recent advances in the synthesis and applications of macrocyclic arenes. Chem. Soc. Rev. 2023, 52, 3265–3298. [Google Scholar] [CrossRef]
  17. Wang, Y.; Ping, G.; Li, C. Efficient complexation between pillar[5]arenes and neutral guests: From host–guest chemistry to functional materials. Chem. Commun. 2016, 52, 9858–9872. [Google Scholar] [CrossRef]
  18. Yang, L.-P.; Liu, W.-E.; Jiang, W. Naphthol-based macrocyclic receptors. Tetrahedron Lett. 2016, 57, 3978–3985. [Google Scholar] [CrossRef]
  19. Wu, J.-R.; Yang, Y.-W. New opportunities in synthetic macrocyclic arenes. Chem. Commun. 2019, 55, 1533–1543. [Google Scholar] [CrossRef] [PubMed]
  20. Wu, J.-R.; Wu, G.; Yang, Y.-W. Pillararene-Inspired Macrocycles: From Extended Pillar[n]arenes to Geminiarenes. Acc. Chem. Res. 2022, 55, 3191–3204. [Google Scholar] [CrossRef]
  21. Dale, E.J.; Vermeulen, N.A.; Juríček, M.; Barnes, J.C.; Young, R.M.; Wasielewski, M.R.; Stoddart, J.F. Supramolecular Explorations: Exhibiting the Extent of Extended Cationic Cyclophanes. Acc. Chem. Res. 2016, 49, 262–273. [Google Scholar] [CrossRef]
  22. Chi, X.; Tian, J.; Luo, D.; Gong, H.-Y.; Huang, F.; Sessler, J.L. “Texas-Sized” Molecular Boxes: From Chemistry to Applications. Molecules 2021, 26, 2426. [Google Scholar] [CrossRef]
  23. Wang, Y.; Wu, H.; Stoddart, J.F. Molecular Triangles: A New Class of Macrocycles. Acc. Chem. Res. 2021, 54, 2027–2039. [Google Scholar] [CrossRef] [PubMed]
  24. Kanagaraj, K.; Rebek, J.; Yu, Y. Self-assembled hexameric capsule: A highly β-selective O-glycosylation reaction enabled via proton wire mechanism. Green Synth. Catal. 2023, 4, 7–9. [Google Scholar] [CrossRef]
  25. Zhang, Z.-Y.; Li, C. Biphen[n]arenes: Modular Synthesis, Customizable Cavity Sizes, and Diverse Skeletons. Acc. Chem. Res. 2022, 55, 916–929. [Google Scholar] [CrossRef]
  26. Yang, W.; Samanta, K.; Wan, X.; Thikekar, T.U.; Chao, Y.; Li, S.; Du, K.; Xu, J.; Gao, Y.; Zuilhof, H.; et al. Tiara[5]arenes: Synthesis, Solid-State Conformational Studies, Host–Guest Properties, and Application as Nonporous Adaptive Crystals. Angew. Chem. Int. Ed. 2020, 59, 3994–3999. [Google Scholar] [CrossRef] [PubMed]
  27. Wu, J.-R.; Mu, A.U.; Li, B.; Wang, C.-Y.; Fang, L.; Yang, Y.-W. Desymmetrized Leaning Pillar[6]arene. Angew. Chem. Int. Ed. 2018, 57, 9853–9858. [Google Scholar] [CrossRef]
  28. Wu, J.-R.; Yang, Y.-W. Geminiarene: Molecular Scale Dual Selectivity for Chlorobenzene and Chlorocyclohexane Fractionation. J. Am. Chem. Soc. 2019, 141, 12280–12287. [Google Scholar] [CrossRef]
  29. Yao, J.; Wu, W.; Xiao, C.; Su, D.; Zhong, Z.; Mori, T.; Yang, C. Overtemperature-protection intelligent molecular chiroptical photoswitches. Nat. Commun. 2021, 12, 2600. [Google Scholar] [CrossRef]
  30. Yang, J.; Lou, X.-Y.; Dai, D.; Shi, J.; Yang, Y.-W. Desymmetrized pillar[8]arenes: High-yield synthesis, functionalization, and host-guest chemistry. Chin. Chem. Lett. 2025, 36, 109818. [Google Scholar] [CrossRef]
  31. Della Sala, P.; Del Regno, R.; Talotta, C.; Capobianco, A.; Hickey, N.; Geremia, S.; De Rosa, M.; Spinella, A.; Soriente, A.; Neri, P.; et al. Prismarenes: A New Class of Macrocyclic Hosts Obtained by Templation in a Thermodynamically Controlled Synthesis. J. Am. Chem. Soc. 2020, 142, 1752–1756. [Google Scholar] [CrossRef] [PubMed]
  32. Han, X.-N.; Han, Y.; Chen, C.-F. Pagoda[4]arene and i-Pagoda[4]arene. J. Am. Chem. Soc. 2020, 142, 8262–8269. [Google Scholar] [CrossRef] [PubMed]
  33. Du, X.-S.; Zhang, D.-W.; Guo, Y.; Li, J.; Han, Y.; Chen, C.-F. Towards the Highly Efficient Synthesis and Selective Methylation of C(sp3)-Bridged[6]Cycloparaphenylenes from Fluoren[3]arenes. Angew. Chem. Int. Ed. 2021, 60, 13021–13028. [Google Scholar] [CrossRef]
  34. Wang, J.-Q.; Han, Y.; Chen, C.-F. 3,6-Fluoren[5]arenes: Synthesis, structure and complexation with fullerenes C60 and C70. Chem. Commun. 2021, 57, 3987–3990. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, X.; Jia, F.; Yang, L.-P.; Zhou, H.; Jiang, W. Conformationally adaptive macrocycles with flipping aromatic sidewalls. Chem. Soc. Rev. 2020, 49, 4176–4188. [Google Scholar] [CrossRef]
  36. Yang, L.-P.; Wang, X.; Yao, H.; Jiang, W. Naphthotubes: Macrocyclic Hosts with a Biomimetic Cavity Feature. Acc. Chem. Res. 2020, 53, 198–208. [Google Scholar] [CrossRef]
  37. Shi, Q.; Wang, X.; Liu, B.; Qiao, P.; Li, J.; Wang, L. Macrocyclic host molecules with aromatic building blocks: The state of the art and progress. Chem. Commun. 2021, 57, 12379–12405. [Google Scholar] [CrossRef]
  38. Ma, C.; Chang, Y.; Yang, J.; Chen, L.; Wu, D.; Go, Y.; Wang, B.; Shi, L.; Li, B. Recent advances in photocatalytic C-H amination to nitrogenous structures. Green Synth. Catal. 2024. [Google Scholar] [CrossRef]
  39. Liu, J.; Yang, Y.; Ouyang, K.; Zhang, W.-X. Transition-metal-catalyzed transformations of C–N single bonds: Advances in the last five years, challenges and prospects. Green Synth. Catal. 2021, 2, 87–122. [Google Scholar] [CrossRef]
  40. Han, X.-N.; Han, Y.; Chen, C.-F. Fluorescent Macrocyclic Arenes: Synthesis and Applications. Angew. Chem. Int. Ed. 2025, 64, e202424276. [Google Scholar] [CrossRef]
  41. Mao, Y.; Ma, J.; Ji, J.; Wang, Y.; Wu, W.; Yang, C. Crown aldoxime ethers: Their synthesis, structure, acid-catalyzed/photo-induced isomerization and adjustable guest binding. Chin. Chem. Lett. 2024, 35, 109927. [Google Scholar] [CrossRef]
  42. Smart, J.E.; Emerson-King, J.; Jeans, R.J.; Hood, T.M.; Lau, S.; Bara-Estaún, A.; Hintermair, U.; Pringle, P.G.; Chaplin, A.B. A Resorcin[4]arene-Based Phosphite-Phosphine Ligand for the Branched-Selective Hydroformylation of Alkyl Alkenes. ACS Catal. 2024, 14, 11803–11807. [Google Scholar] [CrossRef]
  43. Ayuso-Carrillo, J.; Bonifazi, D. Catalyst-Transfer Macrocyclization Protocol: Synthesis of π-Conjugated Azaparacyclophanes Made Easy. JACS Au 2025, 5, 1482–1498. [Google Scholar] [CrossRef]
  44. Espinosa, M.R.; Guerrero, F.; Kazmierczak, N.P.; Oyala, P.H.; Hong, A.; Hadt, R.G.; Agapie, T. Slow Electron Spin Relaxation at Ambient Temperatures with Copper Coordinated by a Rigid Macrocyclic Ligand. J. Am. Chem. Soc. 2025, 147, 14036–14042. [Google Scholar] [CrossRef]
  45. Lemport, P.S.; Petrov, V.S.; Matveev, P.I.; Leksina, U.M.; Roznyatovsky, V.A.; Gloriozov, I.P.; Yatsenko, A.V.; Tafeenko, V.A.; Dorovatovskii, P.V.; Khrustalev, V.N.; et al. First 24-Membered Macrocyclic 1,10-Phenanthroline-2,9-Diamides—An Efficient Switch from Acidic to Alkaline Extraction of f-Elements. Int. J. Mol. Sci. 2023, 24, 10261. [Google Scholar] [CrossRef]
  46. Li, Z.; Zhang, J.; Zhu, W.; Wang, T.; Tang, Y.; Wang, J. N-Heterocyclic carbene-catalyzed enantioselective (dynamic) kinetic resolution for the assembly of inherently chiral macrocycles. Chem. Sci. 2025, 16, 11021–11026. [Google Scholar] [CrossRef] [PubMed]
  47. Ibragimova, R.; Lu, Q.-L.; Wang, M.-X.; Zhu, J.; Tong, S. Asymmetric Synthesis of Inherently Chiral Tetraoxacalix[2]arene[2]pyridines via SNAr-Based Cross-Cyclotetramerization. Chem. Asian J. 2025, 20, e202401621. [Google Scholar] [CrossRef]
  48. Flerin, M.; Duarte, F.; Langton, M.J. Chloride Selective, Nonprotonophoric Ion Transport with Macrocyclic Halogen Bonding Anionophores. Chem. A Eur. J. 2025, 31, e202502033. [Google Scholar] [CrossRef]
  49. Kurashov, I.A.; Kharlamova, A.D.; Abel, A.S.; Averin, A.D.; Beletskaya, I.P. Polyoxa- and Polyazamacrocycles Incorporating 6,7-Diaminoquinoxaline Moiety: Synthesis and Application as Tunable Optical pH-Indicators in Aqueous Solution. Molecules 2023, 28, 512. [Google Scholar] [CrossRef]
  50. Kharlamova, A.D.; Ermakova, E.V.; Abel, A.S.; Gontcharenko, V.E.; Cheprakov, A.V.; Averin, A.D.; Beletskaya, I.P.; Andraud, C.; Bretonnière, Y.; Bessmertnykh-Lemeune, A. Quinoxaline-based azamacrocycles: Synthesis, AIE behavior and acidochromism. Org. Biomol. Chem. 2024, 22, 5181–5192. [Google Scholar] [CrossRef] [PubMed]
  51. Qian, H.; Cheng, Z.P.; Luo, Y.; Lv, L.; Chen, S.; Li, Z. Pd/IPrBIDEA-Catalyzed Hydrodefluorination of gem-Difluorocyclopropanes: Regioselective Synthesis of Terminal Fluoroalkenes. J. Am. Chem. Soc. 2024, 146, 24–32. [Google Scholar] [CrossRef]
  52. Sahoo, D.; Bera, A.; Vennapusa, S.R.; De, S. Light-Triggered Reversible Helicity Switching of a Rotor by a Photo-Responsive Plier. Chem. A Eur. J. 2025, 31, e202404771. [Google Scholar] [CrossRef]
  53. Liang, X.; Zhao, T.; Shen, Y.; Fang, L.; Chen, L.; Zhou, D.; Wu, W.; Yang, C. Nitrogen-Oxidized Tröger’s Base Macrocyclic Arenes: Unprecedented Enantioselective Recognition in Water. Angew. Chem. Int. Ed. 2025, 64, e202416975. [Google Scholar] [CrossRef]
  54. Shen, Y.; Liang, X.; Ma, T.; Zhou, D.; Liu, W.; Ma, J.; Wu, W.; Yu, Z.; Yang, C. Phenacetin[3]Arenes: Mannich-Type Macrocyclization, Unique Structure, Versatile Functionalization, and Strong Allosteric Binding. Angew. Chem. Int. Ed. 2025, 137, e202504211. [Google Scholar] [CrossRef]
  55. Gale, P.A.; Anzenbacher, P., Jr.; Sessler, J.L. Calixpyrroles II. Coord. Chem. Rev. 2001, 222, 57–102. [Google Scholar] [CrossRef]
  56. Kim, S.K.; Sessler, J.L. Calix[4]pyrrole-Based Ion Pair Receptors. Acc. Chem. Res. 2014, 47, 2525–2536. [Google Scholar] [CrossRef]
  57. Yang, P.; Jian, Y.; Zhou, X.; Li, G.; Deng, T.; Shen, H.; Yang, Z.; Tian, Z. Calix[3]carbazole: One-Step Synthesis and Host–Guest Binding. J. Org. Chem. 2016, 81, 2974–2980. [Google Scholar] [CrossRef] [PubMed]
  58. Ostrauskaite, J.; Strohriegl, P. Formation of Macrocycles in the Synthesis of Poly(N-(2-ethylhexyl)carbazol-3,6-diyl). Macromol. Chem. Phys. 2003, 204, 1713–1718. [Google Scholar] [CrossRef]
  59. Coumont, L.S.; Veinot, J.G.C. Ni(COD)2 coupling of 3,6-dibromocarbazoles as a route to all-carbazole shape persistent macrocycles. Tetrahedron Lett. 2015, 56, 5595–5598. [Google Scholar] [CrossRef]
  60. Zhu, H.; Shi, B.; Chen, K.; Wei, P.; Xia, D.; Mondal, J.H.; Huang, F. Cyclo[4]carbazole, an Iodide Anion Macrocyclic Receptor. Org. Lett. 2016, 18, 5054–5057. [Google Scholar] [CrossRef] [PubMed]
  61. Zhu, H.; Badía-Domínguez, I.; Shi, B.; Li, Q.; Wei, P.; Xing, H.; Ruiz Delgado, M.C.; Huang, F. Cyclization-Promoted Ultralong Low-Temperature Phosphorescence via Boosting Intersystem Crossing. J. Am. Chem. Soc. 2021, 143, 2164–2169. [Google Scholar] [CrossRef] [PubMed]
  62. Schmidt, M.; Hermann, M.; Otteny, F.; Esser, B. Calix[n]phenothiazines: Optoelectronic and Structural Properties and Host–Guest Chemistry. Org. Mater. 2020, 02, 235–239. [Google Scholar] [CrossRef]
  63. Chun, Y.; Jiten Singh, N.; Hwang, I.-C.; Woo Lee, J.; Yu, S.U.; Kim, K.S. Calix[n]imidazolium as a new class of positively charged homo-calix compounds. Nat. Commun. 2013, 4, 1797. [Google Scholar] [CrossRef]
  64. Baeyer, A. Ueber ein Condensationsproduct von Pyrrol mit Aceton. Berichte Der Dtsch. Chem. Ges. 1886, 19, 2184–2185. [Google Scholar] [CrossRef]
  65. Wang, M.-X.; Yang, H.-B. A General and High Yielding Fragment Coupling Synthesis of Heteroatom-Bridged Calixarenes and the Unprecedented Examples of Calixarene Cavity Fine-Tuned by Bridging Heteroatoms. J. Am. Chem. Soc. 2004, 126, 15412–15422. [Google Scholar] [CrossRef]
  66. Guo, Q.-H.; Fu, Z.-D.; Zhao, L.; Wang, M.-X. Synthesis, Structure, and Properties of O6-Corona[3]arene[3]tetrazines. Angew. Chem. Int. Ed. 2014, 53, 13548–13552. [Google Scholar] [CrossRef]
  67. Yu, X.; Wu, W.; Zhou, D.; Su, D.; Zhong, Z.; Yang, C. Bisindole[3]arenes—Indolyl Macrocyclic Arenes Having Significant Iodine Capture Capacity. CCS Chem. 2021, 4, 1806–1814. [Google Scholar] [CrossRef]
  68. Mao, L.; Hu, Y.; Tu, Q.; Jiang, W.-L.; Zhao, X.-L.; Wang, W.; Yuan, D.; Wen, J.; Shi, X. Highly efficient synthesis of non-planar macrocycles possessing intriguing self-assembling behaviors and ethene/ethyne capture properties. Nat. Commun. 2020, 11, 5806. [Google Scholar] [CrossRef]
  69. Jin, P.; Liang, W.; Rong, Y.; Wu, W.; Gou, M.; Tang, Y.; Yang, C. Remarkable iodine uptake by aniline-based macrocyclic arenes through a reverse dissolution mechanism. J. Mater. Chem. A 2023, 11, 11126–11132. [Google Scholar] [CrossRef]
  70. Fang, W.; Zhang, J.; Guo, M.; Zhao, Y.; Sue, A.C.H. Triphenylamine[3]arenes: Streamlining Synthesis of a Versatile Macrocyclic Platform for Supramolecular Architectures and Functionalities. Angew. Chem. Int. Ed. 2024, 63, e202409120. [Google Scholar] [CrossRef] [PubMed]
  71. Geuder, W.; Hünig, S.; Suchy, A. Phanes with Two 4,4′-Bipyridinium Moieties—A New Class of Compounds. Angew. Chem. Int. Ed. Engl. 1983, 22, 489–490. [Google Scholar] [CrossRef]
  72. Geuder, W.; Hünig, S.; Suchy, A. Single and double bridged viologenes and intramolecular pimerization of their cation radicals. Tetrahedron 1986, 42, 1665–1677. [Google Scholar] [CrossRef]
  73. Kosiorek, S.; Rosa, B.; Boinski, T.; Butkiewicz, H.; Szymański, M.P.; Danylyuk, O.; Szumna, A.; Sashuk, V. Pillar[4]pyridinium: A square-shaped molecular box. Chem. Commun. 2017, 53, 13320–13323. [Google Scholar] [CrossRef]
  74. Kim, I.; Dhamija, A.; Hwang, I.-C.; Lee, H.; Ko, Y.H.; Kim, K. One-pot Synthesis of a Truncated Cone-shaped Porphyrin Macrocycle and Its Self-assembly into Permanent Porous Material. Chem. Asian J. 2021, 16, 3209–3212. [Google Scholar] [CrossRef]
  75. Sergeyev, S. Recent Developments in Synthetic Chemistry, Chiral Separations, and Applications of Tröger’s Base Analogues. Helv. Chim. Acta 2009, 92, 415–444. [Google Scholar] [CrossRef]
  76. Rúnarsson, Ö.V.; Artacho, J.; Wärnmark, K. The 125th Anniversary of the Tröger’s Base Molecule: Synthesis and Applications of Tröger’s Base Analogues. Eur. J. Org. Chem. 2012, 2012, 7015–7041. [Google Scholar] [CrossRef]
  77. Dolenský, B.; Havlík, M.; Král, V. Oligo Tröger’s bases—New molecular scaffolds. Chem. Soc. Rev. 2012, 41, 3839–3858. [Google Scholar] [CrossRef] [PubMed]
  78. Shi, C.; Xu, G.; Qiu, H.; Li, Y.; Lu, X.; Jiang, J.; Wang, L. Tröger’s base-embedded macrocycles with chirality. Chem. Commun. 2025, 61, 2450–2467. [Google Scholar] [CrossRef] [PubMed]
  79. Feng, H.; Chen, Y.; Wang, R.; Niu, P.; Shi, C.; Yang, Z.; Cheng, M.; Jiang, J.; Wang, L. Chiral selection of Tröger’s base-based macrocycles with different ethylene glycol chains length in crystallization. Chin. Chem. Lett. 2023, 34, 108038. [Google Scholar] [CrossRef]
  80. Mao, L.; Li, F.; Huang, L.; Qu, X.; Wang, K.; Zhang, Y.; Ma, D. Structurally Diversified Calix[3]phenoxazines: Synthesis, Solid-State Conformational Investigation, and Host–Guest Chemistry. Org. Lett. 2023, 25, 5597–5601. [Google Scholar] [CrossRef] [PubMed]
  81. Wei, X.; Liu, J.; Xia, G.-J.; Deng, J.; Sun, P.; Chruma, J.J.; Wu, W.; Yang, C.; Wang, Y.-G.; Huang, Z. Enantioselective photoinduced cyclodimerization of a prochiral anthracene derivative adsorbed on helical metal nanostructures. Nat. Chem. 2020, 12, 551–559. [Google Scholar] [CrossRef]
  82. Tu, C.; Wu, W.; Liang, W.; Zhang, D.; Xu, W.; Wan, S.; Lu, W.; Yang, C. Host–Guest Complexation-Induced Aggregation Based on Pyrene-Modified Cyclodextrins for Improved Electronic Circular Dichroism and Circularly Polarized Luminescence. Angew. Chem. Int. Ed. 2022, 61, e202203541. [Google Scholar] [CrossRef]
  83. Wei, X.; Ji, J.; Nie, Y.; Tang, L.; Rao, M.; Wang, X.; Wu, W.; Su, D.; Zhong, Z.; Yang, C. Synthesis of cyclodextrin derivatives for enantiodifferentiating photocyclodimerization of 2-anthracenecarboxylate. Nat. Protoc. 2022, 17, 2494–2516. [Google Scholar] [CrossRef]
  84. Ji, J.; Wei, X.; Wu, W.; Yang, C. Asymmetric Photoreactions in Supramolecular Assemblies. Acc. Chem. Res. 2023, 56, 1896–1907. [Google Scholar] [CrossRef] [PubMed]
  85. Fang, L.; Guan, X.; Shen, Y.; Zhou, D.; Chen, L.; Chen, X.; Wu, W.; Wang, L.; Yang, C. Multi-step chirality transfer and racemization kinetics of pillar[5]arenes by tuning the halogen substituents on the rims. Org. Chem. Front. 2024, 11, 7229–7234. [Google Scholar] [CrossRef]
  86. Zhao, T.; Yi, J.; Liu, C.; Fang, L.; Chen, L.; Shen, Y.; Liang, X.; Li, K.; Wu, W.; Yang, C. Hierarchical chirality transfer of perylene diimide-tethered pillar[5]arenes for configuration and type differentiation of amino acid derivatives. New J. Chem. 2024, 48, 1856–1859. [Google Scholar] [CrossRef]
  87. Gao, F.; Yu, X.; Liu, L.; Chen, J.; Lv, Y.; Zhao, T.; Ji, J.; Yao, J.; Wu, W.; Yang, C. Chiroptical switching of molecular universal joint triggered by complexation/release of a cation: A stepwise synergistic complexation. Chin. Chem. Lett. 2023, 34, 107558. [Google Scholar] [CrossRef]
  88. Lv, Y.; Xiao, C.; Ma, J.; Zhou, D.; Wu, W.; Yang, C. Solvent and guest-binding-controlled chiroptical inversion of molecular devices based on pseudo[1]catenane-type pillar[5]arene derivatives. Chin. Chem. Lett. 2024, 35, 108757. [Google Scholar] [CrossRef]
  89. Huang, R.; Wei, X.; Wang, P.; Ma, J.; Mao, Y.; Zhou, D.; Wu, W.; Ji, J.; Yang, C. Chirality Induction and Memory of Pillar[4]arene[1]quinone Derivatives in Visible-Light Range. Org. Lett. 2024, 26, 1405–1409. [Google Scholar] [CrossRef] [PubMed]
  90. Sivaguru, J.; Natarajan, A.; Kaanumalle, L.S.; Shailaja, J.; Uppili, S.; Joy, A.; Ramamurthy, V. Asymmetric Photoreactions within Zeolites:  Role of Confinement and Alkali Metal Ions. Acc. Chem. Res. 2003, 36, 509–521. [Google Scholar] [CrossRef] [PubMed]
  91. Shaw, S.; White, J.D. Asymmetric Catalysis Using Chiral Salen–Metal Complexes: Recent Advances. Chem. Rev. 2019, 119, 9381–9426. [Google Scholar] [CrossRef]
  92. Cabré, A.; Verdaguer, X.; Riera, A. Recent Advances in the Enantioselective Synthesis of Chiral Amines via Transition Metal-Catalyzed Asymmetric Hydrogenation. Chem. Rev. 2022, 122, 269–339. [Google Scholar] [CrossRef]
  93. Kuang, H.; Xu, C.; Tang, Z. Emerging Chiral Materials. Adv. Mater. 2020, 32, 2005110. [Google Scholar] [CrossRef]
  94. Subrahmanyam, K.S.; Sarma, D.; Malliakas, C.D.; Polychronopoulou, K.; Riley, B.J.; Pierce, D.A.; Chun, J.; Kanatzidis, M.G. Chalcogenide Aerogels as Sorbents for Radioactive Iodine. Chem. Mater. 2015, 27, 2619–2626. [Google Scholar] [CrossRef]
  95. Küpper, F.C.; Feiters, M.C.; Olofsson, B.; Kaiho, T.; Yanagida, S.; Zimmermann, M.B.; Carpenter, L.J.; Luther Iii, G.W.; Lu, Z.; Jonsson, M.; et al. Commemorating Two Centuries of Iodine Research: An Interdisciplinary Overview of Current Research. Angew. Chem. Int. Ed. 2011, 50, 11598–11620. [Google Scholar] [CrossRef]
  96. Saiz-Lopez, A.; Plane, J.M.C.; Baker, A.R.; Carpenter, L.J.; von Glasow, R.; Gómez Martín, J.C.; McFiggans, G.; Saunders, R.W. Atmospheric Chemistry of Iodine. Chem. Rev. 2012, 112, 1773–1804. [Google Scholar] [CrossRef] [PubMed]
  97. Zeng, F.; Xiao, X.-S.; Gong, S.-F.; Yuan, L.; Tang, L.-L. An electron-deficient supramolecular macrocyclic host for the selective separation of aromatics and cyclic aliphatics. Org. Chem. Front. 2022, 9, 4829–4833. [Google Scholar] [CrossRef]
  98. Zhou, H.-Y.; Zhang, D.-W.; Li, M.; Chen, C.-F. A Calix[3]acridan-Based Host–Guest Cocrystal Exhibiting Efficient Thermally Activated Delayed Fluorescence. Angew. Chem. Int. Ed. 2022, 61, e202117872. [Google Scholar] [CrossRef]
  99. Zhou, H.-Y.; Zhang, D.-W.; Han, X.-N.; Han, Y.; Chen, C.-F. A novel thermally activated delayed fluorescence macrocycle. Chem. Commun. 2022, 58, 12180–12183. [Google Scholar] [CrossRef] [PubMed]
  100. Sun, Y.; Jiang, L.; Liu, L.; Chen, Y.; Xu, W.-W.; Niu, J.; Qin, Y.; Xu, X.; Liu, Y. Two Calix[3]Phenothiazine-Based Amorphous Pure Organic Room-Temperature Phosphorescent Supramolecules Mediated by Guest. Adv. Opt. Mater. 2023, 11, 2300326. [Google Scholar] [CrossRef]
  101. Ma, J.-H.; Han, Y.; Guo, W.-C.; Lu, H.-Y.; Chen, C.-F. Color-Tunable Thermally Activated Delayed Fluorescence Polymeric Materials Constructed by Host–Guest Complexation Revealing Rewritable Advanced Information Encryption. CCS Chem. 2025, 7, 2623–2632. [Google Scholar] [CrossRef]
Figure 1. (a) Synthesis of calix[4]pyrrole 1; (b) synthesis of calix[3]carbazole compounds [57]. Copyright © 2016, American Chemical Society.
Figure 1. (a) Synthesis of calix[4]pyrrole 1; (b) synthesis of calix[3]carbazole compounds [57]. Copyright © 2016, American Chemical Society.
Molecules 30 03646 g001
Figure 2. Synthesis of calix[2]arene[2]triazine macrocyclic 917 [65]. Copyright © 2004, American Chemical Society.
Figure 2. Synthesis of calix[2]arene[2]triazine macrocyclic 917 [65]. Copyright © 2004, American Chemical Society.
Molecules 30 03646 g002
Figure 3. (a) The synthetic route of O6-corona[3]arene[3]tetrazines; (b) single-crystal structure of O6-corona[3]arene[3]tetrazines; (c) single-crystal structure of Cl complex with O6-corona[3]arene[3]tetrazines [66]. Copyright © 2014 WILEY-VCH Verlag GmbH&Co.KGaA, Weinheim.
Figure 3. (a) The synthetic route of O6-corona[3]arene[3]tetrazines; (b) single-crystal structure of O6-corona[3]arene[3]tetrazines; (c) single-crystal structure of Cl complex with O6-corona[3]arene[3]tetrazines [66]. Copyright © 2014 WILEY-VCH Verlag GmbH&Co.KGaA, Weinheim.
Molecules 30 03646 g003
Figure 4. The synthetic route of macrocyclic 3739.
Figure 4. The synthetic route of macrocyclic 3739.
Molecules 30 03646 g004
Figure 5. (a,b) Synthesis of indole[3]arene and bisindole[3]arenes; (c) crystal structures of 41a41b [67]. Copyright © 2021 Chinese Chemical Society.
Figure 5. (a,b) Synthesis of indole[3]arene and bisindole[3]arenes; (c) crystal structures of 41a41b [67]. Copyright © 2021 Chinese Chemical Society.
Molecules 30 03646 g005
Figure 6. (a) Synthesis and structure diagram of DPA[n] macrocycle 42–48; (b) single-crystal stacking diagram of 42; (c) single-crystal stacking diagram of 43 [68].
Figure 6. (a) Synthesis and structure diagram of DPA[n] macrocycle 42–48; (b) single-crystal stacking diagram of 42; (c) single-crystal stacking diagram of 43 [68].
Molecules 30 03646 g006
Figure 7. (a) Synthesis and structure of macrocycle 4951 and single crystal of macrocycle 50; (b) synthesis and structure diagram of 5255; (c) the single-crystal structure diagram of 5255. Copyright © 2024 Wiley-VCH GmbH.
Figure 7. (a) Synthesis and structure of macrocycle 4951 and single crystal of macrocycle 50; (b) synthesis and structure diagram of 5255; (c) the single-crystal structure diagram of 5255. Copyright © 2024 Wiley-VCH GmbH.
Molecules 30 03646 g007
Figure 8. (a) The chemical structures of 4,4 ‘-bipyridine macrocycles 5658; (b) the chemical structure of bis(paraquat-p-phenylene) 59 and its derivatives macrocyclic 6062.
Figure 8. (a) The chemical structures of 4,4 ‘-bipyridine macrocycles 5658; (b) the chemical structure of bis(paraquat-p-phenylene) 59 and its derivatives macrocyclic 6062.
Molecules 30 03646 g008
Figure 9. (a) Synthesis of calix[4]imidazoline 64 and calix[5]imidazoline 65; (b) synthetic route of macrocyclic 67; (c,d) single-crystal structure of 67.
Figure 9. (a) Synthesis of calix[4]imidazoline 64 and calix[5]imidazoline 65; (b) synthetic route of macrocyclic 67; (c,d) single-crystal structure of 67.
Molecules 30 03646 g009
Figure 10. (a) The schematic diagram of the synthesis of imine porphyrin macrocycle P3L3; (b) conformation and stacking of P3L3 in the crystal structure.
Figure 10. (a) The schematic diagram of the synthesis of imine porphyrin macrocycle P3L3; (b) conformation and stacking of P3L3 in the crystal structure.
Molecules 30 03646 g010
Figure 11. The possible reaction mechanism of TB proposed by Wagner and Farrar.
Figure 11. The possible reaction mechanism of TB proposed by Wagner and Farrar.
Molecules 30 03646 g011
Figure 12. (a) Synthesis of ethylene glycol-link-TB macrocyclic path diagram; (b) Schematic diagram of the mirror relationship between ethylene glycol-linked TB macrocyclic. Copyright © 2023 Published by Elsevier B.V. on behalf of Chinese Chemical Society and Institute of Materia Medica, Chinese Academy of Medical Sciences.
Figure 12. (a) Synthesis of ethylene glycol-link-TB macrocyclic path diagram; (b) Schematic diagram of the mirror relationship between ethylene glycol-linked TB macrocyclic. Copyright © 2023 Published by Elsevier B.V. on behalf of Chinese Chemical Society and Institute of Materia Medica, Chinese Academy of Medical Sciences.
Molecules 30 03646 g012
Figure 13. (a) Synthesis of macrocycle 7677 and (b,c) nitrogen oxygen macrocycle 7879 path diagram. Copyright © 2024 Wiley-VCH GmbH.
Figure 13. (a) Synthesis of macrocycle 7677 and (b,c) nitrogen oxygen macrocycle 7879 path diagram. Copyright © 2024 Wiley-VCH GmbH.
Molecules 30 03646 g013
Figure 14. (a) Synthesis of Phenacetin[3]arenes 8085; (b) derivatization of the alkoxy end group of 86.
Figure 14. (a) Synthesis of Phenacetin[3]arenes 8085; (b) derivatization of the alkoxy end group of 86.
Molecules 30 03646 g014
Figure 15. The synthetic route of calix[3]phenoxazine 88a88h and the chemical structures of guests G1G6. Copyright © 2023, American Chemical Society.
Figure 15. The synthetic route of calix[3]phenoxazine 88a88h and the chemical structures of guests G1G6. Copyright © 2023, American Chemical Society.
Molecules 30 03646 g015
Figure 16. (a) synthesis of calix[3]carbazole compounds; (b) guest of the test. Copyright © 2016, American Chemical Society.
Figure 16. (a) synthesis of calix[3]carbazole compounds; (b) guest of the test. Copyright © 2016, American Chemical Society.
Molecules 30 03646 g016
Figure 17. The equilibrium diagrams of C3-80 and F-80 single host, 1:1, and 1:2 host–guest complexes. Copyright © 2025 Wiley-VCH GmbH.
Figure 17. The equilibrium diagrams of C3-80 and F-80 single host, 1:1, and 1:2 host–guest complexes. Copyright © 2025 Wiley-VCH GmbH.
Molecules 30 03646 g017
Figure 18. (a) Structure diagram of nitrogen oxide macrocycle 7879; (b) the binding diagram of naphthyl phosphate G11. Copyright © 2024 Wiley-VCH GmbH.
Figure 18. (a) Structure diagram of nitrogen oxide macrocycle 7879; (b) the binding diagram of naphthyl phosphate G11. Copyright © 2024 Wiley-VCH GmbH.
Molecules 30 03646 g018
Figure 19. (a) Structures of bisindolo[3]arene 41a41c; (b) time-dependent I2 vapor adsorption curve. Copyright © 2021 Chinese Chemical Society.
Figure 19. (a) Structures of bisindolo[3]arene 41a41c; (b) time-dependent I2 vapor adsorption curve. Copyright © 2021 Chinese Chemical Society.
Molecules 30 03646 g019
Figure 20. (a) at 348.15 K, the I2 vapor absorption rate of 4951, Ans, P-Ns, and ACs versus time curves; (b) photos of 4951, P-Ns, and ACs absorbing I2 vapor at different time intervals. (c) FT-IR spectra of 50 and I2-50.
Figure 20. (a) at 348.15 K, the I2 vapor absorption rate of 4951, Ans, P-Ns, and ACs versus time curves; (b) photos of 4951, P-Ns, and ACs absorbing I2 vapor at different time intervals. (c) FT-IR spectra of 50 and I2-50.
Molecules 30 03646 g020
Figure 21. (a) Preparation and photos of 89 and G12 co-crystals; (b) S-triphenyltriazine-derived calix[3]acridine 91; the PL spectra of 91 before and after nitrogen purge and the transient PL attenuation curve of 91 in toluene under saturated nitrogen were studied. Copyright © 2022 Wiley-VCH GmbH.
Figure 21. (a) Preparation and photos of 89 and G12 co-crystals; (b) S-triphenyltriazine-derived calix[3]acridine 91; the PL spectra of 91 before and after nitrogen purge and the transient PL attenuation curve of 91 in toluene under saturated nitrogen were studied. Copyright © 2022 Wiley-VCH GmbH.
Molecules 30 03646 g021
Figure 22. (a) The schematics of RTP emission of selectively activated G12@92 and G13@92 are shown; (b) photographs of 92, G12@92, and G13@92 solid powders under visible light and 365 nm UV light, as well as phosphorescent QR codes and solvent-responsive information storage/encryption. Copyright © 2023 Wiley-VCH GmbH.
Figure 22. (a) The schematics of RTP emission of selectively activated G12@92 and G13@92 are shown; (b) photographs of 92, G12@92, and G13@92 solid powders under visible light and 365 nm UV light, as well as phosphorescent QR codes and solvent-responsive information storage/encryption. Copyright © 2023 Wiley-VCH GmbH.
Molecules 30 03646 g022
Figure 23. (a) The construction of TADF polymer materials based on 93-P; (b) a schematic diagram of the application of the polymer material in advanced information encryption. Copyright © 2024 Chinese Chemical Society.
Figure 23. (a) The construction of TADF polymer materials based on 93-P; (b) a schematic diagram of the application of the polymer material in advanced information encryption. Copyright © 2024 Chinese Chemical Society.
Molecules 30 03646 g023
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, J.; Wu, W.; Yang, C. Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes. Molecules 2025, 30, 3646. https://doi.org/10.3390/molecules30173646

AMA Style

Hu J, Wu W, Yang C. Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes. Molecules. 2025; 30(17):3646. https://doi.org/10.3390/molecules30173646

Chicago/Turabian Style

Hu, Jianhang, Wanhua Wu, and Cheng Yang. 2025. "Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes" Molecules 30, no. 17: 3646. https://doi.org/10.3390/molecules30173646

APA Style

Hu, J., Wu, W., & Yang, C. (2025). Recent Advances in the Synthesis and Applications of Nitrogen-Containing Macrocyclic Arenes. Molecules, 30(17), 3646. https://doi.org/10.3390/molecules30173646

Article Metrics

Back to TopTop