Next Article in Journal
Improved Process for the Synthesis of 3-(3-Trifluoromethylphenyl)propanal for More Sustainable Production of Cinacalcet HCl
Next Article in Special Issue
Tunable Supramolecular Ag+-Host Interactions in Pillar[n]arene[m]quinones and Ensuing Specific Binding to 1-Alkynes
Previous Article in Journal
Secondary Metabolomic Analysis and In Vitro Bioactivity Evaluation of Stems Provide a Comprehensive Comparison between Dendrobium chrysotoxum and Dendrobium thyrsiflorum
Previous Article in Special Issue
Meso-Formyl, Vinyl, and Ethynyl Porphyrins—Multipotent Synthons for Obtaining a Diverse Array of Functional Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bismacrocycle: Structures and Applications

1
Hubei Key Laboratory of Pollutant Analysis and Reuse Technology, College of Chemistry and Chemical Engineering, Hubei Normal University, Huangshi 435002, China
2
College of Chemistry, Beijing Normal University, Beijing 100875, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(16), 6043; https://doi.org/10.3390/molecules28166043
Submission received: 7 July 2023 / Revised: 3 August 2023 / Accepted: 10 August 2023 / Published: 13 August 2023
(This article belongs to the Special Issue Macrocyclic Compounds: Derivatives and Applications)

Abstract

:
In the past half-century, macrocycles with different structures and functions, have played a critical role in supramolecular chemistry. Two macrocyclic moieties can be linked to form bismacrocycle molecules. Compared with monomacrocycle, the unique structures of bismacrocycles led to their specific recognition and assembly properties, also a wide range of applications, including molecular recognition, supramolecular self-assembly, advanced optical material construction, etc. In this review, we focus on the structure of bismacrocycle and their applications. Our goal is to summarize and outline the possible future development directions of bismacrocycle research.

Graphical Abstract

1. Introduction

In the past six decades, macrocycles with different structures and properties have emerged. These macrocycles include crown ethers [1,2,3], cyclodextrins [4,5,6], calixarenes [7,8], cucurbiturates [9,10], and carbon-rich macrocycles [11,12,13], etc. They have continuously promoted the development of supramolecular chemistry. As important hosts, macrocycles have been widely used in molecular recognition [14,15,16], assembly [17,18,19], molecular machine construction [20,21,22,23], novel material development [24,25,26,27,28], drug exploration [29,30,31], etc. To date, the design and synthesis of macrocycles with special structures and properties is one of the core driving forces of supramolecular chemistry. The combination of multiple cavities and active sites is a promising strategy. Herein, one macrocycle with two cavities is defined as a bismacrocycle. Currently, macrocycle-related studies have been extensive, and their supramolecular chemical properties have also been widely investigated. As the combination result of two macrocycles, a bismacrocycle can effectively expand their properties with high reliability and predictability on the basis of the reports involving its macrocyclic moieties (Scheme 1). However, even some review papers explore related study of specific bismacrocycles [32,33], the summary of bismacrocycles as the subject is lacking [34]. Herein, recent progress of bismacrocycles study is reviewed. It provides an important reference for bismacrocycle-related research (e.g., supramolecular assembly, supramolecular polymer construction, etc.). Specifically, the latest developments in bicyclic compounds since 2017 were summarized for further driving their research development in different fields.

2. Oxabismacrocycle

2.1. Bis(crown ether)

Smid group synthesized the first bis(crown ether) in 1975 [35], which exhibited stronger interactions with cations (e.g., K+, NH4+) than the corresponding mono(crown ether). It initiated the research on the host with double macrocyclic moieties (i.e., bismacrocycle) in the following decades. To date, a large number of bismacrocycles were generated and widely used in molecular recognition [36] and supramolecular polymer studies [37,38].
In 2017, Liu group synthesized a bis(crown ether) 1 (Figure 1) [39]. The 1 contained photosensitive 9,10-diphenylanthracene block and terminal pyridinyl groups. It can coordinate with lanthanide metal ions to construct assembly 2. Under ultraviolet light, 2 can undergo photocatalytic oxidation to form 3 with excellent luminescence performance. Heating induced 3 returning to 2. The light/thermal regulation between 2 and 3 indicates its good potential to achieve molecular machines and logic gate systems.
The novel bis(crown ether) 4 (Figure 2) [40] was synthesized by Xing’s group. As a supramolecular mechanical cross-linker, self-assembly 5 was generated with host-guest interaction between 4 and polymer 6. It showed significantly higher viscoelastic properties than 7. It exhibited thermal, pH, and chemical response-modulated gel-sol properties, also outstanding viscoelasticity.
In 2019, Liu and colleagues synthesized bis(crown ether) 8 containing azobenzene bridge (Figure 3) [41]. Trans-8 and cholesterol derivative 9 form a snowflake-like supramolecular bell-shaped helix structure 10. Conversion from trans- to cis-8 in 10 under 365 nm UV light created 11. The 11 has no helical structure and shows concomitant loss of CD signal. Especially, 11 can be transformed as 10 again under illumination (wavelength longer than 420 nm). The morphology and chiral property modulation of supramolecular assembly structures with light can provide en route to subsequent light-driven manual switching and information storage studies.
Meanwhile, an anthracene-bridged bis(crown ether) 12 (Figure 4) [42] was reported by the Liu group. A novel photochromic pseudo[3]rotaxane 13 was formed between 12 and (R/S)-2,2′-binaphthyl secondary ammonium salt guest (R/S)-14. The intrinsic chiral transfer from (R/S)-14 to 12 is accompanied by fluorescence resonance energy transfer (FRET). Photo-oxidation of anthracene on 12 can further modulate circular dichroism (ICD) and FRET of (R/S)-13 treated with 365 nm UV light or heating. The multi-stimuli reactive chiral transfer materials can be used to developing chiral functional materials.
In the same year, Stang and colleagues synthesized a bis(crown ether) 15 (Figure 5) [43] containing a pyridine linking unit. The 15 and platinum formed the complex 16 (Figure 5a). Hydrogen bonding interactions between 16 and water induced supramolecular polymer 17 with excellent adhesion to hydrophilic surfaces (e.g., glass).
In 2020, Yan and colleagues obtained bis(crown ether) 18 containing two dialkylammonium salt units. Its self-cross-linking supramolecular polymer network (SPN) formed on the basis of the interaction between its crown ether and dialkylammonium units (Figure 6) [44]. The gel-sol transition of SPN can be modulated with temperature or pH. This study is instructive for developing gel materials with non-covalent interactions.

2.2. Biscyclodextrin

Cyclodextrins and their derivatives are widely used in supramolecular self-assembly, supramolecular polymer, and biopharmaceuticals. They are cheap, easily availabile, and non-toxic [4,5,6]. Biscyclodextrins can be prepared via simply linking two cyclodextrin units with linking fragment(s). They can be applied as the building blocks for further self-assemblies [45,46,47,48].
In 2017, Liu group synthesized a biscyclodextrin 19 (Figure 7a) [49] bridged with dithiophene. The 19 is converted to 20 under UV light, and 20 returned to 19 after visible light irradiation. The 19 form 22 with the porphyrin tricarboxylate guest 21, then 22 further self-assembles to form 23. The 23 aggregation generated the binary supermolecular nanoassembly 24. The 23 co-assembles with amphipathic near-infrared (NIR) cyanine fluorochrome 25 to construct 26, and then further aggregated to obtain a larger ternary supermolecular nanoassembly 27. The 25 significantly induced 27 fluorescence enhancement (Figure 7b). Nanoparticles 24 or 27 emitted red light at a maximum wavelength at 647 nm or 680 nm under 417 nm excitation, respectively. Both were fluorescence quenched after treating with 254 nm UV light. Subsequently, the luminescence can be restored under >450 nm light (Figure 7c,d). The light-modifiable multivariate nano-assembled structures provides a new strategy to design and develop photoconvertible photoluminescent materials.
The azobenzene unit was used as a bridging unit to generate bis(β-cyclodextrin) 28 (Figure 8) [50]. Trans-28 and adamantanyl-modified diphenylalanine (30) form bilayer 2D nanosheets 29. Differently, cis-28 and 30 forms 1D nanotubes 31. The 29 and 31 can be converted to each other via UV/visible illumination or heating. This study provides a new way to achieve light control assembly morphology.
The bipyridinyl unit was used as linker to obtain bis(β-cyclodextrin) 32 (Figure 9) [51]. The 32 formed a 3:1 complex with Ru(II), which can bind with adamantane-modified anthracene (33) in aqueous solution to form 34. The 34 can accumulate in the nucleus of cancer cells and induce reactive oxygen species production under visible light irradiation. The net result was effective anticancer activity. This work provided a new strategy for light-driven bright agents for cancer treatment.

2.3. Biscalixarene

Calixarene is a class of artificial macrocycles with high application value in chemical, biological, material, environmental, and other multidisciplinary fields due to their low cost, efficient synthesis, and unique properties [7,8]. From the end of the last century, biscalixarenes have been developed for supramolecular assembly and supramolecular polymer construction [52].
In 2018, Ma and colleagues reported a novel water-soluble vibration-induced emission (VIE) molecule 35 (Figure 10) [53]. Two quaternary ammonium groups on 35 can insert into the cavity of 36. Then supramolecular aggregates 37 was created. Free 35 emits orange-red light, while fixed 35 emits blue light. Fluorescence emission from orange-red to white to blue is achieved via controlling the ratio between 36 and 35. Acetylcholine (ACH) was added in 37 solutions to displace 35 and thus restore orange fluorescence. Such strategies using reversible supramolecular self-assembly to control the emission of VIE molecules provide new ideas to develop tunable luminescent materials.
Linking two calix[5]arene with chiral binaphthalene provides chiral biscalix[5]arene 38 and 39. Both interact with C60 on C60-appended poly-(phenylacetylene) 40 to affect the helix of the polymer (Figure 11) [54]. It is a new approach to control the helical conformation of polyacetylene via host-guest interactions.
(E)-azobenzene or (E)-stilbene as bridging units link two calix[4]arenes to construct bis(calix[4]arenes) (E,E)-41 or (E,E)-42 (Figure 12) [55]. (E,E)-41 and its derivatives can be converted to (Z,Z)- forms with a shortened length about 4 Å under 365 nm UV illumination. (Z,Z)- forms subsequently returned back to (E,E)-41 via heating at 50 °C. Tetraester substituted (E,E)-42 can be photoisomerized as (Z,Z)-42 in the presence of photosensitizer 1,2-benzanthracene. Due to its good stability, it has been separated and cultivated to obtain a single crystal structure. (E,E)-42 generates [2 + 2] cycloaddition under 365 nm light in CH2Cl2 to obtain 43 and 44. This research provides the material basis to generate complex light-controlled supramolecular assemblies.

2.4. Bispillararene

In 2020, Cong group introduced tetraphenylethylene as linker to obtain novel fluorescent bispillararene 45 (Figure 13) [56]. The fluorescence of crystalline 45 changes from light blue to yellow after grinding, and then it can be returned to blue treated with p-xylene vapor. Further information encryption patterns can be prepared.
In 2021, Liu group reported a novel bispillararene 46 (Figure 14) [57]. The 46 and photosensitive guest 47 form a 1:1 complex 48 with [1 + 1] or [2 + 2] form. Adding Nile red (NiR) to 48 aqueous solutions can obtain three component self-assembled 49. Increasing NiR ratio gradually changed the fluorescence of 49 from green to white, and then to orange red under 365 nm light. Under 254 nm illumination, the fluorescence of 49 quenched. A total of 450 nm visible light irradiation can restore 49 luminescence (Figure 14). Herein, the reversible morphology of 47 can be controlled via irradiating 49 with different wavelength light. Then the resonance energy transfer (RET) between 46 and NiR can be regulated to modulate the luminescent performance of 49. This strategy can be further applied to the preparation of photoresponsive supramolecular intelligent materials.
In 2023, Hu group linked two pillar[5]arenes through C-C double-bonds to obtain bispillar[5]arene 50 containing a tetraphenylethylene (TPE) unit (Figure 15) [58]. Host-guest studies showed that 50 can form a ring-piercing structure with adiponitrile 51 or sebaconitrile 52 in a 1:2 ratio, and then 50 and 52 formed a linear supramolecular polyme. Furthermore, 50 forms a supramolecular layered polymer 54 with 53, and 54 can be used as a photocatalyst to catalyse the dehalogenation reaction.

2.5. Bishelicarene

Chen and colleagues reported a pair of enantiomer bishelic[6]arenes P/M-55 (Figure 16) [59] in 2022. It contained two chiral helic[6]arene units. P/M-55 self-assembles with guest 56, a tetraphenylethylene derivative containing two quaternary ammonium units. Finally, a chiral supramolecular polymer 57 formed. The anthracene contained guest 58 combined with 57 to further obtain emission-enhanced supramolecular gel 59. The circular polarization luminescence (CPL) performance of 59 is induced by the chiral transfer from P/M-55 to the non-chiral guest 58. Furthermore, regulating the ratio of guests 56 and 58 in 59 can emit white light. This work provides a new strategy to construct CPL active supramolecular gel through chiral bismacrocycles.

2.6. Biscucurbituril

Zhang group synthesized the biscucurbit[7]uril 60 in rotaxane in 2017 (Figure 17) [60]. Guest 61 containing N,N-dimethyladamantylammonium (DMAD) on both end groups was added to bind 60. Supramolecular polymer 62 was rapidly formed with a high polymerization degree. To control the polymerization degree, dicarboxylate guest 63 was used to interact with 60 to form complex 64. Then 61 was subsequently added in the system to replace 63 due to its stronger affinity with 60. However, the slow dissociation rate of 63 induced the polymerization degree of 62 can be controlled. When the environment pH was changed, the supramolecular polymerization process was suspended when the pD value increased from 9.7 to 10.8, while the supramolecular polymer remained stable for several days. In this work, pH modulated the dissociation rate of the pre-saturated complex 64, and further regulated the polymerization degree of 62.

2.7. Bisheteracalixarenes

Wang and colleagues used bridging units with different angles and stiffnesses to connect two oxacalix[2]arene[2]triazines to obtain bisheteracalixarenes (65–69) (Figure 18a) [61]. The crystal structure of 65–67 and 69 proved that the rigid bridged units can control the orientation of the two large ring cavities, and further modified the stoichiometry of the complexation between each bismacrocycle and chloride aion. Further supramolecular oligomers between 67 and binary naphthalene-1,5-disulfonate anion (70) was generated (Figure 18b). The coherent self-assembled particles formed via mixing 67 and 1 molar equiv. of 70. This study was carried out by rational design of electron-deficient bismacrocycle building units to produce the challenging anion-π interaction-directed self-assembly.

2.8. Other Oxabismacrocycle

In 2023, Yam group synthesized two multifunctional bis-CPPs 71 and 72 via integrating pillar[5]arene into the [n]CPP backbone precisely (Figure 19) [62]. Compared with [n]CPP, the photoluminescence quantum yield (ΦF) values of 71 and 72 substantially increase. Furthermore, 71 shows good circularly polarized luminescence (glum ≈ 0.02).

3. Azabismacrocycle

3.1. Pyridinium Bismacrocycle

Pyridinium macrocycles (e.g., “blue box” and its derivatives) have been used for a wide range of applications in supramolecular structures, host, and guest chemistry, catalysis, extraction and sequestration, and molecular electronics due to their unique structures and properties [63]. In addition, pyridinium bismacrocycles have also shown promising applications in supramolecular self-assembly, anion recognition, and bioimaging.
In 2019, Cao group synthesized the aggregation-induced emission (AIE) pyridinium bismacrocycles 73 containing three tetraphenylethene (TPE) units. (Figure 20a) [64] 73 emits orange light (cantered at 595 nm) in acetonitrile with ΦF as 19.7%, while it emits strong yellow light (cantered at 580 nm) in water with a ΦF as 97.7%. The fluorescence intensity of 73 changed very little with the percentage of water in acetonitrile lower than 80%, and increased rapidly with the percentage above 80%. The 73 forms nanosphere supramolecular assemblies 74 with diameters ranging from 25 to 77 nm in water, with a maximum emission wavelength of 580 nm (excitation wavelength 410 nm). The addition of NiR to the aqueous solution of 74 leads to the formation of further spherical supramolecular assemblies 75 with an increasing diameter as 35 to 83 nm through highly ordered co-assembly. Due to FRET (ΦET = 77.5%) between 73 and NiR with a high antenna effect (14.3), the maximum emission wavelength of 75 is redshifted to 650 nm (excitation wavelength as 410 nm). This AIE fluorescent nanomaterial has potential applications in cancer cell imaging and diagnosis/photodynamic therapy. In 2021, this group further introduced four different substituents (i.e., NO2, Br, OCH3, or OH) on 73 (Figure 20b) [65], resulting in bismacrocycle 76–79 with AIE properties. Electron-absorbing groups on 76 or 77 can prohibit the intramolecular PET process between TPE donor and acceptor so that induced enhanced fluorescence. Conversely, 78 and 79 containing electron-donating groups cannot prohibit intramolecular PET process and then cause fluorescence bursts. The 76–79 can also self-assemble into nanospheres in MeCN or H2O (1% MeCN). ATP can be encapsulated in the cavities or gaps of the nanospheres formed with 73, 76–79. ATP binding leads to a significant fluorescence reduction in 76, which can be used to detect ATP.
Cong group use a similar strategy to create a water-soluble AIE pyridinium bismacrocycles 80 (Figure 21) [66]. The 80 showed a constant weak yellow fluorescence in aqueous solutions up to 80% THF, while above 90% THF, the fluorescence increased significantly with a blue shift (Δλmax = 46 nm) and the solution became turbid due to the aggregation of 80. Perfluorooctane sulfonate (81) binds 80 in aggregation form following with increasing red shift fluorescence intensity (Δλmax = 22 nm). The 80 exhibits excellent specific recognition of 81 in a variety of anions or cations. This property was further used to visualize and quantify PFOS via smartphone with a detection limit of 47.3 ± 2.0 nmol/L (25.4 ± 1.1 µg/L).
In 2022, Zhao and colleagues synthesized 82 (Figure 22a) [67], a water-soluble pyridinium bismacrocycles containing a perylene diimides (PDIs) core. The 82 was encapsulated by double cationic molecular straps on both sides of PDI so that prevents PDI aggregation induced fluorescence quench even at high concentration (e.g., 2 mM) (Figure 22a). Cell fluorescence imaging studies showed that 82 co-localized with the lysosomal labelling dye LysoTracker Red in RAW 264.7 cells with an overlay coefficient as 0.98 and cell viability greater than 95% at all concentration tests (1–200 μM) (Figure 22b). PDI radical species from 82 reduction remained stable at room temperature and heat (60 °C), exhibited photothermal performance without significant decay for 20 cycles under 808 nm radiation. PDI in 82 can be converted in situ to PDI radicals with hydrogenases on the surface of the anaerobic bacterium E. coli, and can rise from 25 °C to 68 °C in 15 min under 808 nm laser radiation (Figure 22c). This work provides a new strategy to design and synthesis water-soluble non-aggregated organic dyes.
A similar strategy can be applied to design more bismacrocycle. In 2023, Wei and colleagues generated a water-soluble pyridinium bismacrocycles 83 (Figure 23) [68] containing a naphthalene diimide core. The 83 acts as an electron-deficient host that binds strongly to electron-rich guests such as water-insoluble 2,7-diaminofluorene (84), fluorene (85), 2-aminofluorene (86), tetrathiafulvalene (87), and water-soluble oligoethylene glycol chain substituted 2,7-diaminofluorene (88) in a 1:2 host-guest stoichiometry ratio. The maximum absorption wavelengths of guest283 were found to be 482, 712, 860, and 1063 nm for the case involving 85, 86, 84, and 87, respectively. The results suggested that guests with stronger electron donating ability can induce greater charge transfer (CT) absorption redshifts. Upon 1064 nm laser irradiation, aqueous solutions of 84283, 87283, or 88283 exhibited significant warming with thermal conversion efficiency value as 37.6%, 39.9%, and 47.4%, respectively. Upon 1064 nm laser irradiation, the non-toxic 88283 completely killed HeLa cells, and E. coli and S. aureus, achieving efficient NIR-II photothermal conversion for cancer cell and bacterial ablation (Figure 20 top left). This study provides new avenues to design and apply biocompatible NIR-II light absorbers with well-defined structures.
Recently, Cao and colleagues reported three TPE-containing pyridinium bismacrocycles 89–91 via a one-step SN2 reaction (Figure 24) [69]. The 89–91 binds nicotinamide adenine dinucleotides (NAD/NADH) in a 1:2 ratio in aqueous solution, with NAD in the open form and NADH in the folded form. This study provides an idea to regulate the conformation of NADH in organisms through host-guest interactions.

3.2. Biscalix[4]pyrroles

Calix[4]pyrrole is widely used to bind anions and ion pairs [70]. One or more bridging “walls” connecting two calix[4]pyrrole provide biscalix[4]pyrroles that are widely used in ion recognition, sensing and logic gate structures [33].
In 2017, Ballester group created [2]rotaxane based on biscalix[4]pyrrole 92 and 3,5-bis(amido)pyridine-N-oxide derivatives [71]. Additional anion can modulate the decomposition and combination of multi-component self-assembled structures (Figure 25a). In the same year, this group introduced chiral 1,2-substituted aliphatic diamines to the bridging unit to obtain the chiral biscalix[4]pyrrole 93 (Figure 25b) [72]. The 93 can form supramolecular capsules with double N-oxides with unprecedentedly efficient chiral transfer. Meanwhile, Sessler group generated biscalix[4]pyrrole 94 with a large-cavity (Figure 25c) [73]. The 94 can capture two monovalent H2PO4, two divalent SO42− and two trivalent HP2O73− anions simultaneously. In 2020, He and colleagues synthesized biscalix[4]pyrrole 95 that specifically recognizes F (Figure 25d) [74]. In 2022, Kim and colleagues synthesized biscalix[4]pyrrole 96 that binds F and acetate anions in solution with a 1:2 acceptor/anion ratio, and forms 1:1 complexes with oxygenated anions such as C2O42−, H2PO4, SO42− and HP2O73− (Figure 25e) [75].

3.3. Imidazolium Bismacrocycle

In 2020, Cao group synthesized imidazolium bismacrocycles 97 and 98 (Figure 26) [76] with TPE core induced AIE properties. The 97 and 98 show good emission in a wide range of solvents. The 97 changed its emission colour from blue to green (Δλ = 12 nm) after grinding, and returned to blue after fumigation with water vapor. Due to the complete flattening of the benzene ring in the bicyclic structure at high pressure, a large red shift in the emission wavelength (Δλ = ~72 nm) of 97 was observed. Due to the restricted rotation of the anthracene group in 98, its emission peak does not vary with pressure. An acetonitrile solution of 98 underwent an oxidation reaction under 365 nm light and the fluorescence gradually changed from green to blue within a few minutes. The synthesis of such novel molecules for mechanochromic and photochromic luminescence can provide novel smart luminescent materials.

3.4. Azabiscycloparaphenylene

Stępień and colleagues obtained the radially conjugated azabiscycloparaphenylene (azabis-CPP) 99 in 2019 (Figure 27) [77]. The 9,9′-bicarbazole core of 99 acts as a stereospecific element giving the entire molecule an “8” twisted structure. The electronic circular dichroism (ECD) spectra of each pure 99 enantiomer contains two major Cotton effects with opposite signs. The curvature control method proposed in this work can reduce the electronic band gap while maintaining a large conjugate length in the nano-hoop system.
In 2021, Sun group obtained the azabis-CPP 100–102 (Figure 28) [78] with the cyclocondensation between two o-diamine-substituted CPP derivative moieties and a 4,5,9,10-tetraketopyrene. The structure of 100–102 rapidly interconverts between cis- and trans- conformations. Analysis of the NMR hydrogen spectra of the well-solubilized 102 at different temperatures revealed >99% trans structure at temperatures below 183 K. The maximum emission wavelength of 102 in dichloromethane was 616 nm, the brightest fluorophore of the CPP derivatives with λem > 600 nm, and its high quantum yield of 80% was one of the highest values for CPP derivatives. The 102 can bind C60 with a 1:2 ratio in solution.

4. Biscycloparaphenylene (bis-CPP)

In 2019, Cong group synthesized bis-CPP 103 (Figure 29) [79] bridged by benzene rings [80]. The main steps involve the inversion of the dianthracene retro-[4 + 4] cycloreversion, and the ring expansion in the 64-membered macrocycle by transannular aryne [4 + 2] cycloaddition. The crystal structure shows that the two CPPs in 103 linked by the pentiptycene core are ellipsoidal in shape. The 103 precursor was treated with HPLC chiral seperation and further reduction to obtain its enantiomer with an average luminescence dissymmetry factor (glum) as 3.4 × 10−3.
In 2020, Jasti and colleagues produced a series of bis-CPPs with 9,9′-spirobifluorene as the core. These bismacrocycles included fluorescent bis-CPPs (104–106) with different sizes, and non-luminescent Azabis-CPP (107) (Figure 30a) [81]. The porosities of 104–106 were detected as 33.9 m2/g, 703.0 m2/g, and 11.8 m2/g, respectively. X-ray single crystal diffraction experiments showed that 104–105 were loosely supramolecularly arranged, whereas 107 were relatively well-ordered (Figure 30d). Double-cavity CPP had a higher affinity for VOCs (up to 480%) than the single-cavity [9]CPP (Figure 30c,d).
In 2021, Du group generated a highly strained bis-[10]CPP (108) (Figure 31a) [82]. The double-cavity twisted structure of 108 was confirmed by scanning tunnelling microscopy. Theoretical calculations show strain energy up to 110.59 kcal mol−1, while the central benzene on 108 has a torsion angle as 10.05° and a maximum interphenylene torsion angle as 46.07°. The 108 has maximum absorption wavelength at 352 nm, and maximum emission peak at 523 nm with fluorescence quantum yield as 5% and lifetime as τ1 = 4.23 ns, τ2 = 8.46 ns, and τ3 = 16.93 ns. UV-Vis titration and Job Plot analysis provided the peanut-like binding constants between 108 and 109 as K1 = (7.46 ± 0.33) × 105 M−1 and K2 = (5.85 ± 0.25) × 104 M−1 with a binding ratio as 1:2. In 2022, Du and colleagues synthesized bis-[8]CPP 110 containing a distorted benzene core similar as 108 (Figure 31(b1)) [83]. The 110 fluoresces displayed a maximum emission wavelength at 475 nm under an excitation wavelength as 380 nm in THF, with. The fluorescence quantum yield was ~3% and a fluorescence lifetime was 4.23 ns. Increasing water proportion in THF beyond 60%, 110 showed a new emission band at ~577 nm with gradually decreasing emission at 475 nm (Figure 31(b2)). The results indicate that 110 is characterized by both the aggregation-caused quenching (ACQ) and AIE effects, and can induce tunable emission from cyan to red, including near-white light emission (Figure 31(b3,b4)). The temperature-dependent CD spectra demonstrate the strong stability of both 110 enantiomeric isomers. Moreover, the AIE effect of 110 enhances its CPL properties. This molecule has potential applications as white light emitters, AIE sensors, and chiral luminescent materials.
In 2021, Juríček group synthesized two fluorescent bis-CPPs, 111 [84] and 112 [85] (Figure 32a), containing peropyrene cores. The X-ray diffraction (XRD) analysis of single crystals of 111 and 112 demonstrated that they are fully conjugated framework structures with C2 symmetry. The 111 cannot interact with C60 or C70 due to spatial resistance. In contrast, 112 can form a 1:1 complex with C60 (Figure 32b).
In 2022, Cong and colleagues synthesized 113 (Figure 33) [86], a fully conjugated bis-CPP containing a flexible cyclooctathiophene core. The crystals of 113 and C60 or C70 were obtained by slow volatilization of o-dichlorobenzene in excess of C60 or C70, respectively. The assembly structure of 113 with C60 or C70 in a 1:2 ratio to form a peanut-like topology was isolated and verified.

5. Conclusions

In summary, we highlighted recent progress of the bismacrocycle study. The combination of double cavities brings unique properties. To date, bismacrocycles were used in the construction of luminescent materials, supramolecular self-assembly, and supramolecular functional polymers (e.g., gels). The current strategies to generated bismacrocycles included using functionalized linking unit to join two macrocycles (e.g., crown ethers, cyclodextrins, calixarene, cucurbiturates, etc.). In these cases, since both cavities are known, their guest recognition properties are easily accessible for subsequent functionalisation studies. Especially, the simple introduction of functionalized bridging units (e.g., anthracene, azobenzene, dithiophene, etc.) facilitates the property modification of the resulted bismacrocycles and further supramolecular structures. Its simple synthesis steps and low cost are more conducive to pushing the relevant research results to practical applications. Bismacrocycles also can be generated as a whole. This strategy focuses on the introduction of functional structures (e.g., TPE, PDI, etc.) into the bismacrocycles (Table 1). The advantage corresponding to the strategy is some properties (e.g., AIE) can be expected shown in the final product, and mainly used to further supramolecular chemistry exploration.
The design and synthesis of bismacrocycle compounds with outstanding structural novelty and performance is a major challenge in related research. The combination between the art of current bismacrocyclic research and computational methods [87,88,89,90] lead the design and study of specific functionally oriented bismacrocycles. As one of the fast-developing research frontiers, it is expected that bismacrocycle-related chemistry will be further penetrated many fields, including advanced optical materials, disease treatment, and molecular machines, etc. Bismacrocycle study will provide more excellent solutions to achieve the precise construction and properties of complex systems.

Author Contributions

Conceptualization, X.-L.C. and H.-Y.G.; software and drawing, X.-L.C., S.-Q.Y. and X.-H.H.; writing—original draft preparation, X.-L.C., S.-Q.Y. and X.-H.H.; writing—review and editing, X.-L.C. and H.-Y.G.; funding acquisition,X.-L.C. and H.-Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

H.-Y.G. is grateful to the National Natural Science Foundation of China (92156009 and 21971022) for financial support. X.-L.C. is grateful to the Science and Technology Research Project of Hubei Provincial Department of Education (Q20222506).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Qi, Z.; Qin, Y.; Wang, J.; Zhao, M.; Yu, Z.; Xu, Q.; Nie, H.; Yan, Q.; Ge, Y. The aqueous supramolecular chemistry of crown ethers. Front. Chem. 2023, 11, 1119240. [Google Scholar] [CrossRef] [PubMed]
  2. Gokel, M.R.; McKeever, M.; Meisel, J.W.; Negin, S.; Patel, M.B.; Yin, S.; Gokel, G.W. Crown ethers having side arms: A diverse and versatile supramolecular chemistry. J. Coord. Chem. 2021, 74, 14–39. [Google Scholar] [CrossRef]
  3. Li, J.; Yim, D.; Jang, W.-D.; Yoon, J. Recent progress in the design and applications of fluorescence probes containing crown ethers. Chem. Soc. Rev. 2017, 46, 2437–2458. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, W.; Ye, B.; Yu, G.; Huang, F.; Mao, Z.; Ding, Y.; Wang, W. Recent Development of Supramolecular Cancer Theranostics Based on Cyclodextrins: A Review. Molecules 2023, 28, 3441. [Google Scholar] [CrossRef] [PubMed]
  5. Healy, B.; Yu, T.; da Silva Alves, D.C.; Okeke, C.; Breslin, C.B. Cyclodextrins as Supramolecular Recognition Systems: Applications in the Fabrication of Electrochemical Sensors. Materials 2021, 14, 1668. [Google Scholar] [CrossRef]
  6. Crini, G. Review: A history of cyclodextrins. Chem. Rev. 2014, 114, 10940–10975. [Google Scholar]
  7. Shurpik, D.N.; Padnya, P.L.; Stoikov, I.I.; Cragg, P.J. Antimicrobial Activity of Calixarenes and Related Macrocycles. Molecules 2020, 25, 5145. [Google Scholar] [CrossRef]
  8. Pan, Y.-C.; Hu, X.-Y.; Guo, D.-S. Biomedical Applications of Calixarenes: State of the Art and Perspectives. Angew. Chem. Int. Ed. 2021, 60, 2768–2794. [Google Scholar] [CrossRef]
  9. Wang, Z.; Sun, C.; Yang, K.; Chen, X.; Wang, R. Cucurbituril-Based Supramolecular Polymers for Biomedical Applications. Angew. Chem. Int. Ed. 2022, 61, e2022067. [Google Scholar]
  10. Liu, Y.-H.; Zhang, Y.-M.; Yu, H.-J.; Liu, Y. Cucurbituril-Based Biomacromolecular Assemblies. Angew. Chem. Int. Ed. 2021, 60, 3870–3880. [Google Scholar] [CrossRef]
  11. Shen, Y.-J.; Zhu, K.-L.; Liang, J.-Q.; Sun, X.; Gong, H.-Y. Carbon-rich macrocycles and carbon nanoribbons as unique optical materials. J. Mater. Chem. C 2023, 11, 4267–4287. [Google Scholar] [CrossRef]
  12. Guo, G.-H.; Qiu, Y.; Wang, M.-X.; Stoddart, J.F. Aromatic hydrocarbon belts. Nat. Chem. 2021, 13, 402–419. [Google Scholar] [CrossRef]
  13. Leonhardt, E.J.; Jasti, R. Emerging applications of carbon nanohoops. Nat. Rev. Chem. 2019, 3, 672–686. [Google Scholar] [CrossRef]
  14. Crowley, P.B. Protein−Calixarene Complexation: From Recognition to Assembly. Acc. Chem. Res. 2022, 55, 2019–2032. [Google Scholar] [CrossRef]
  15. Barrow, S.J.; Kasera, S.; Rowland, M.J.; del Barrio, J.; Scherman, O.A. Cucurbituril-Based Molecular Recognition. Chem. Rev. 2015, 115, 12320–12406. [Google Scholar] [CrossRef] [Green Version]
  16. Yu, G.; Jie, K.; Huang, F. Supramolecular Amphiphiles Based on Host–Guest Molecular Recognition Motifs. Chem. Rev. 2015, 115, 7240–7303. [Google Scholar] [CrossRef]
  17. Wu, J.-R.; Wu, G.; Li, D.; Yang, Y.-W. Macrocycle-Based Crystalline Supramolecular Assemblies Built with Intermolecular Charge-Transfer Interactions. Angew. Chem. Int. Ed. 2023, 135, e202218142. [Google Scholar] [CrossRef]
  18. Nie, H.; Wei, Z.; Ni, X.-L.; Liu, Y. Assembly and Applications of Macrocyclic-Confinement-Derived Supramolecular Organic Luminescent Emissions from Cucurbiturils. Chem. Rev. 2022, 122, 9032–9077. [Google Scholar] [CrossRef]
  19. 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]
  20. Erbas-Cakmak, S.; Leigh, D.A.; McTernan, C.T.; Nussbaumer, A.L. Artifificial Molecular Machines. Chem. Rev. 2015, 115, 10081–10206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Dattler, D.; Fuks, G.; Heiser, J.; Moulin, E.; Perrot, A.; Yao, X.; Giuseppone, N. Design of Collective Motions from Synthetic Molecular Switches, Rotors, and Motors. Chem. Rev. 2020, 120, 310–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Moulin, E.; Faour, L.; Carmona-Vargas, C.C.; Giuseppone, N. From Molecular Machines to Stimuli-Responsive Materials. Adv. Mater. 2020, 32, 1906036. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, Y.; Ovalle, M.; Seale, J.S.W.; Lee, C.K.; Kim, D.J.; Astumian, R.D.; Stoddart, J.F. Molecular Pumps and Motors. J. Am. Chem. Soc. 2021, 143, 5569–5591. [Google Scholar] [CrossRef] [PubMed]
  24. Li, E.; Zhu, W.; Fang, S.; Jie, K.; Huang, F. Reimplementing Guest Shape Sorting of Nonporous Adaptive Crystals via Substituent-Size-Dependent Solid-Vapor Postsynthetic Modification. Angew. Chem. Int. Ed. 2022, 61, e202211780. [Google Scholar]
  25. Liang, Y.; Li, E.; Wang, K.; Guan, Z.-J.; He, H.-H.; Zhang, L.; Zhou, H.-C.; Huang, F.; Fang, Y. Organo-macrocycle-containing hierarchical metal–organic frameworks and cages: Design, structures, and applications. Chem. Soc. Rev. 2022, 51, 8378–8405. [Google Scholar] [CrossRef]
  26. Ji, X.; Ahmed, M.; Long, L.; Khashab, N.M.; Huang, F.; Sessler, J.L. Adhesive supramolecular polymeric materials constructed from macrocycle-based host–guest interactions. Chem. Soc. Rev. 2019, 48, 2682–2697. [Google Scholar] [CrossRef] [PubMed]
  27. Zhou, Y.; Jie, K.; Zhao, R.; Huang, F. Supramolecular-Macrocycle-Based Crystalline Organic Materials. Adv. Mater. 2019, 32, 1904824. [Google Scholar] [CrossRef]
  28. Zhang, H.; Zou, R.; Zhao, Y. Macrocycle-based metal-organic frameworks. Coord. Chem Rev. 2015, 292, 74–90. [Google Scholar] [CrossRef]
  29. Pan, Z.; Zhao, X.; Li, X.; Zhang, Z.; Liu, Y. Recent Advances in Supramolecular-Macrocycle-Based Nanomaterials in Cancer Treatment. Molecules 2023, 28, 1241. [Google Scholar] [CrossRef]
  30. Shan, P.H.; Hu, J.H.; Liu, M.; Tao, Z.; Xiao, X.; Redshaw, C. Progress in host-guest macrocycle/pesticide research: Recognition, detection, release and application. Coord. Chem Rev. 2022, 467, 214580. [Google Scholar] [CrossRef]
  31. Zhou, J.; Yu, G.; Huang, F. Supramolecular chemotherapy based on host–guest molecular recognition: A novel strategy in the battle against cancer with a bright future. Chem. Soc. Rev. 2017, 46, 7021–7053. [Google Scholar] [CrossRef]
  32. Guo, D.-S.; Liu, Y. Calixarene-based supramolecular polymerization in solution. Chem. Soc. Rev. 2012, 41, 5907–5921. [Google Scholar] [CrossRef] [PubMed]
  33. Lai, Z.; Zhao, T.; Sessler, J.L.; He, Q. Bis–Calix[4]pyrroles: Preparation, structure, complexation properties and beyond. Coord. Chem. Rev. 2020, 425, 213528. [Google Scholar] [CrossRef]
  34. Peng, H.-Q.; Zhu, W.; Guo, W.-J.; Li, Q.; Ma, S.; Bucher, C.; Liu, B.; Ji, X.; Huang, F.; Sessler, J.L. Supramolecular polymers: Recent advances based on the types of underlying interactions. Prog. Polym. Sci. 2023, 137, 101635. [Google Scholar] [CrossRef]
  35. Bourgoin, M.; Wong, K.H.; Hui, J.Y.; Smid, J. Interactions of macrobicyclic polyethers with ions and ion pairs of picrate salts. J. Am. Chem. Soc. 1975, 97, 3462–3467. [Google Scholar] [CrossRef]
  36. An, H.; Bradshaw, J.S.; Izatt, R.M.; Yan, Z. Bis- and Oligo(benzocrown ether)s. Chem. Rev. 1994, 94, 939–991. [Google Scholar] [CrossRef]
  37. Li, S.; Lu, H.-Y.; Shen, Y.; Chen, C.-F. A Stimulus-Response and Self-Healing Supramolecular Polymer Gel Based on Host–Guest Interactions. Macromol. Chem. Phys. 2013, 214, 1596–1601. [Google Scholar] [CrossRef]
  38. Yan, X.; Xu, D.; Chen, J.; Zhang, M.; Hu, B.; Yu, Y.; Huang, F. A self-healing supramolecular polymer gel with stimuli-responsiveness constructed by crown ether based molecular recognition. Polym. Chem. 2013, 4, 3312–3322. [Google Scholar] [CrossRef]
  39. Zhou, Y.; Zhang, H.-Y.; Zhang, Z.-Y.; Liu, Y. Tunable Luminescent Lanthanide Supramolecular Assembly Based on Photoreaction of Anthracene. J. Am. Chem. Soc. 2017, 139, 7168–7171. [Google Scholar] [CrossRef]
  40. Wang, W.; Xing, H. A novel supramolecular polymer network based on a catenane-type crosslinker. Polym. Chem. 2018, 9, 2087–2091. [Google Scholar] [CrossRef]
  41. Wang, H.-J.; Zhang, H.-Y.; Wu, H.; Dai, X.-Y.; Li, P.-Y.; Liu, Y. Photocontrolled morphological conversion and chiral transfer of a snowflake-like supramolecular assembly based on azobenzene-bridged bis(dibenzo-24-crown-8) and a cholesterol derivative. Chem. Commun. 2019, 55, 4499–4502. [Google Scholar] [CrossRef] [PubMed]
  42. Fu, H.-G.; Zhang, H.-Y.; Zhang, H.-Y.; Liu, Y. Photo-controlled chirality transfer and FRET effects based on pseudo[3]rotaxane. Chem. Commun. 2019, 55, 13462–13465. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Q.; Li, T.; Duan, A.; Dong, S.; Zhao, W.; Stang, P.J. Formation of a Supramolecular Polymeric Adhesive via Water–Participant Hydrogen Bond Formation. J. Am. Chem. Soc. 2019, 141, 8058–8063. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, L.; Cheng, L.; Li, G.; Liu, K.; Zhang, Z.; Li, P.; Dong, S.; Yu, W.; Huang, F.; Yan, X. A Self-Cross-Linking Supramolecular Polymer Network Enabled by Crown-Ether-Based Molecular Recognition. J. Am. Chem. Soc. 2020, 142, 2051–2058. [Google Scholar] [CrossRef]
  45. Kuad, P.; Miyawaki, A.; Takashima, Y.; Yamaguchi, H.; Harada, A. External Stimulus-Responsive Supramolecular Structures Formed by a Stilbene Cyclodextrin Dimer. J. Am. Chem. Soc. 2007, 129, 12630–12631. [Google Scholar] [CrossRef] [PubMed]
  46. Dong, R.; Su, Y.; Yu, S.; Zhou, Y.; Lu, Y.; Zhu, X. A redox-responsive cationic supramolecular polymer constructed from small molecules as a promising gene vector. Chem. Commun. 2013, 49, 9845–9847. [Google Scholar] [CrossRef] [PubMed]
  47. Li, Z.-Q.; Zhang, Y.-M.; Chen, Y.; Liu, Y. A Supramolecular Tubular Nanoreactor. Chem. Eur. J. 2014, 20, 8566–8570. [Google Scholar] [CrossRef]
  48. Liu, Y.; Wang, K.R.; Guo, D.-S.; Jiang, B.-P. Supramolecular Assembly of Perylene Bisimide with β-Cyclodextrin Grafts as a Solid-State Fluorescence Sensor for Vapor Detection. Adv. Funct. Mater. 2009, 19, 2230–2235. [Google Scholar] [CrossRef]
  49. Liu, G.; Zhang, Y.-M.; Xu, X.; Zhang, L.; Yu, Q.; Zhao, Q.; Zhao, C.; Liu, Y. Controllable Photoluminescence Behaviors of Amphiphilic Porphyrin Supramolecular Assembly Mediated by Cyclodextrins. Adv. Opt. Mater. 2017, 5, 1700770. [Google Scholar] [CrossRef]
  50. Sun, H.-L.; Chen, Y.; Han, X.; Liu, Y. Tunable Supramolecular Assembly and Photoswitchable Conversion of Cyclodextrin/Diphenylalanine-Based 1D and 2D Nanostructures. Angew. Chem. Int. Ed. 2017, 56, 7062–7065. [Google Scholar] [CrossRef]
  51. Cheng, N.; Chen, Y.; Yu, J.; Li, J.-J.; Liu, Y. Enhanced DNA Binding and Photocleavage Abilities of β-Cyclodextrin Appended Ru(II) Complex through Supramolecular Strategy. Bioconjugate Chem. 2018, 29, 1829–1833. [Google Scholar] [CrossRef] [PubMed]
  52. Rebek, J. Host–guest chemistry of calixarene capsules. Chem. Commun. 2000, 637–643. [Google Scholar] [CrossRef]
  53. Wang, J.; Yao, X.; Liu, Y.; Zhou, H.; Chen, W.; Sun, G.; Su, J.; Ma, X.; Tian, H. Tunable Photoluminescence Including White-Light Emission Based on Noncovalent Interaction-Locked N,N′-Disubstituted Dihydrodibenzo[a,c]phenazines. Adv. Opt. Mater. 2018, 6, 1800074. [Google Scholar] [CrossRef]
  54. Hirao, T.; Iwabe, Y.; Hisano, N.; Haino, T. Helicity of a polyacetylene directed by molecular recognition of biscalixarene and fullerene. Chem. Commun. 2020, 56, 6672–6675. [Google Scholar] [CrossRef] [PubMed]
  55. Lentin, I.; Gorbunov, A.; Bezzubov, S.; Nosova, V.; Cheshkov, D.; Kovalev, V.; Vatsouro, I. Shrinkable/stretchable bis(calix[4]arenes) comprising photoreactive azobenzene or stilbene linkers. Org. Chem. Front. 2023, 10, 1470–1484. [Google Scholar] [CrossRef]
  56. Lei, S.-N.; Xiao, H.; Zeng, Y.; Tung, C.-H.; Wu, L.-Z.; Cong, H. BowtieArene: A Dual Macrocycle Exhibiting Stimuli-Responsive Fluorescence. Angew. Chem. Int. Ed. 2020, 59, 10059–10065. [Google Scholar] [CrossRef]
  57. Liu, G.; Zhang, H.; Xu, X.; Zhou, Q.; Dai, X.; Fan, L.; Mao, P.; Liu, Y. Supramolecular photoswitch with white-light emission based on bridged bis(pillar[5]arene)s. Mater. Today Chem. 2021, 22, 100628. [Google Scholar] [CrossRef]
  58. Wang, K.; Zhang, R.; Song, Z.; Zhang, K.; Tian, X.; Pangannaya, S.; Zuo, M.; Hu, X.-Y. Dimeric Pillar[5]arene as a Novel Fluorescent Host for Controllable Fabrication of Supramolecular Assemblies and Their Photocatalytic Applications. Adv. Sci. 2023, 10, 2206897. [Google Scholar] [CrossRef]
  59. Guo, Y.; Han, Y.; Du, X.-S.; Chen, C.-F. Chiral Bishelic[6]arene-Based Supramolecular Gels with Circularly Polarized Luminescence Property. ACS Appl. Polym. Mater. 2022, 4, 3473–3481. [Google Scholar] [CrossRef]
  60. Chen, H.; Huang, Z.; Wu, H.; Xu, J.-F.; Zhang, X. Supramolecular Polymerization Controlled through Kinetic Trapping. Angew. Chem. Int. Ed. 2017, 56, 16575–16578. [Google Scholar] [CrossRef]
  61. Tuo, D.-H.; Liu, W.; Wang, X.-Y.; Wang, X.-D.; Ao, Y.-F.; Wang, Q.-Q.; Li, Z.-Y.; Wang, D.-X. Toward Anion−π Interactions Directed Self-Assembly with Predesigned Dual Macrocyclic Receptors and Dianions. J. Am. Chem. Soc. 2019, 141, 1118–1125. [Google Scholar] [CrossRef]
  62. Xu, Y.; Steudel, F.; Leung, M.-Y.; Xia, B.; von Delius, M.; Yam, V.W.-W. [n]Cycloparaphenylene-Pillar[5]arene Bismacrocycles: Their Circularly Polarized Luminescence and Multiple Guest Recognition Properties. Angew. Chem. Int. Ed. 2023, 62, e202302978. [Google Scholar] [CrossRef] [PubMed]
  63. Chen, X.-Y.; Chen, H.; Stoddart, J.F. The story of the little blue box: A tribute to Siegfried Hünig. Angew. Chem. Int. Ed. 2023, 62, e202211387. [Google Scholar]
  64. Li, Y.; Dong, Y.; Cheng, L.; Qin, C.; Nian, H.; Zhang, H.; Yu, Y.; Cao, L. Aggregation-Induced Emission and Light-Harvesting Function of Tetraphenylethene-Based Tetracationic Dicyclophane. J. Am. Chem. Soc. 2019, 141, 8412–8415. [Google Scholar] [CrossRef]
  65. Qin, C.; Li, Y.; Li, Q.; Yan, C.; Cao, L. Aggregation-induced emission and self-assembly of functional tetraphenylethene-based tetracationic dicyclophanes for selective detection of ATP in water. Chin. Chem. Lett. 2021, 32, 3531–3534. [Google Scholar] [CrossRef]
  66. Lei, S.-N.; Cong, H. Fluorescence detection of perfluorooctane sulfonate in water employing a tetraphenylethylene-derived dual macrocycle BowtieCyclophane. Chin. Chem. Lett. 2022, 33, 1493–1496. [Google Scholar] [CrossRef]
  67. Yang, F.; Li, R.; Wei, W.; Ding, X.; Xu, Z.; Wang, P.; Wang, G.; Xu, Y.; Fu, H.; Zhao, Y. Water-Soluble Doubly-Strapped Isolated Perylene Diimide Chromophore. Angew. Chem. Int. Ed. 2022, 61, e202202491. [Google Scholar]
  68. Li, R.; Yang, F.; Zhang, L.; Li, M.; Wang, G.; Wang, W.; Xu, Y.; Wei, W. Manipulating Host-Guest Charge Transfer of a Water-Soluble Double-Cavity Cyclophane for NIR-II Photothermal Therapy. Angew. Chem. Int. Ed. 2023, 62, e202301267. [Google Scholar] [CrossRef]
  69. Wang, L.; Guo, Z.; Cheng, L.; Nian, H.; Yao, N.; Zhao, Y.; Liu, K.; Cao, L. Tetraphenylethene-linked octacationic dicyclophanes with enhanced recognition of NADH over NAD+ in water. Dyes Pigments 2023, 216, 111364. [Google Scholar] [CrossRef]
  70. Kim, D.S.; Sessler, J.F. Calix[4]pyrroles: Versatile molecular containers with ion transport, recognition, and molecular switching functions. Chem. Soc. Rev. 2015, 44, 532–546. [Google Scholar] [CrossRef]
  71. Romero, J.R.; Aragay, G.; Ballester, P. Ion-pair recognition by a neutral [2]rotaxane based on a bis-calix[4]pyrrole cyclic component. Chem. Sci. 2017, 8, 491–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Galán, A.; Escudero-Adán, E.C.; Ballester, P. Template-directed self-assembly of dynamic covalent capsules with polar interiors. Chem. Sci. 2017, 8, 7746–7750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. He, Q.; Kelliher, M.; Bähring, S.; Lynch, V.M.; Sessler, J.L. A Bis-calix[4]pyrrole Enzyme Mimic That Constrains Two Oxoanions in Close Proximity. J. Am. Chem. Soc. 2017, 139, 7140–7143. [Google Scholar] [CrossRef] [PubMed]
  74. Xiong, S.; Chen, F.; Zhao, T.; Li, A.; Xu, G.; Sessler, J.L.; He, Q. Selective Inclusion of Fluoride within the Cavity of a Two-Wall Bis-calix[4]pyrrole. Org. Lett. 2020, 22, 4451–4455. [Google Scholar] [CrossRef] [PubMed]
  75. Oh, J.H.; Hay, B.P.; Lynch, V.M.; Li, H.; Sessler, J.L.; Kim, S.K. Calix[4]pyrrole-Based Molecular Capsule: Dihydrogen Phosphate Promoted 1:2 Fluoride Anion Complexation. J. Am. Chem. Soc. 2022, 144, 16996–17009. [Google Scholar] [CrossRef]
  76. Nian, H.; Li, A.; Li, Y.; Cheng, L.; Wang, L.; Xu, W.; Cao, L. Tetraphenylethene-based tetracationic dicyclophanes: Synthesis, mechanochromic luminescence, and photochemical reactions. Chem. Commun. 2020, 56, 3195–3198. [Google Scholar] [CrossRef]
  77. Senthilkumar, K.; Kondratowicz, M.; Lis, T.; Chmielewski, P.J.; Cybińska, J.; Zafra, J.L.; Casado, J.; Vives, T.; Crassous, J.; Favereau, L.; et al. Lemniscular [16]Cycloparaphenylene: A Radially Conjugated Figure-Eight Aromatic Molecule. J. Am. Chem. Soc. 2019, 141, 7421–7427. [Google Scholar] [CrossRef]
  78. Li, K.; Xu, Z.; Deng, H.; Zhou, Z.; Dang, Y.; Sun, Z. Dimeric Cycloparaphenylenes with a Rigid Aromatic Linker. Angew. Chem. Int. Ed. 2021, 60, 7649–7653. [Google Scholar] [CrossRef]
  79. Xu, W.; Yang, X.-D.; Fan, X.-B.; Wang, X.; Tung, C.-H.; Wu, L.-Z.; Cong, H. Synthesis and Characterization of a Pentiptycene-Derived Dual Oligoparaphenylene Nanohoop. Angew. Chem. Int. Ed. 2019, 58, 3943–3947. [Google Scholar] [CrossRef]
  80. Huang, Z.-A.; Chen, C.; Yang, X.-D.; Fan, X.-B.; Zhou, W.; Tung, C.-H.; Wu, L.-Z.; Cong, H. Synthesis of Oligoparaphenylene-Derived Nanohoops Employing an Anthracene Photodimerization–Cycloreversion Strategy. J. Am. Chem. Soc. 2016, 138, 11144–11147. [Google Scholar] [CrossRef]
  81. Schaub, T.A.; Prantl, E.A.; Kohn, J.; Bursch, M.; Marshall, C.R.; Leonhardt, E.J.; Lovell, T.C.; Zakharov, L.N.; Brozek, C.K.; Waldvogel, S.R.; et al. Exploration of the Solid-State Sorption Properties of Shape-Persistent Macrocyclic Nanocarbons as Bulk Materials and Small Aggregates. J. Am. Chem. Soc. 2020, 142, 8763–8775. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, X.; Shi, H.; Zhuang, G.; Wang, S.; Wang, J.; Yang, S.; Shao, X.; Du, P. A Highly Strained All-Phenylene Conjoined Bismacrocycle. Angew. Chem. Int. Ed. 2021, 60, 17368–17372. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, X.; Liu, H.; Zhuang, G.; Yang, S.; Du, P. An unexpected dual-emissive luminogen with tunable aggregation-induced emission and enhanced chiroptical property. Nat. Commun. 2022, 13, 3543. [Google Scholar] [CrossRef] [PubMed]
  84. Yang, Y.; Blacque, O.; Sato, S.; Juríček, M. Cycloparaphenylene–Phenalenyl Radical and Its Dimeric Double Nanohoop. Angew. Chem. Int. Ed. 2021, 60, 13529–13535. [Google Scholar] [CrossRef]
  85. Yang, Y.; Huangfu, S.; Sato, S.; Juríček, M. Cycloparaphenylene Double Nanohoop: Structure, Lamellar Packing, and Encapsulation of C60 in the Solid State. Org. Lett. 2021, 23, 7943–7948. [Google Scholar] [CrossRef]
  86. Cong, H. A Conjugated Figure-of-Eight Oligoparaphenylene Nanohoop with Adaptive Cavities Derived from Cyclooctatetrathiophene Core. Angew. Chem. Int. Ed. 2022, 61, e202113334. [Google Scholar]
  87. Xu, H.; Guan, D.; Ma, L. The bio-inspired heterogeneous single-cluster catalyst Ni100–Fe4S4 for enhanced electrochemical CO2 reduction to CH4. Nanoscale 2023, 15, 2756–2766. [Google Scholar] [CrossRef]
  88. Xiao, W.; Kiran, G.K.; Yoo, K.; Kim, J.-H.; Xu, H. The Dual-Site Adsorption and High Redox Activity Enabled by Hybrid Organic-Inorganic Vanadyl Ethylene Glycolate for High-Rate and Long-Durability Lithium–Sulfur Batteries. Small 2023, 19, 2206750. [Google Scholar] [CrossRef]
  89. Xu, H.; Guan, D. Exceptional Anisotropic Noncovalent Interactions in Ultrathin Nanorods: The Terminal σ-Hole. ACS Appl. Mater. Interfaces 2022, 14, 51190–51199. [Google Scholar] [CrossRef]
  90. Lv, Z.; Xu, H.; Xu, W.; Peng, B.; Zhao, C.; Xie, M.; Lv, X.; Gao, Y.; Hu, K.; Fang, Y.; et al. Quasi-Topological Intercalation Mechanism of Bi0.67NbS2 Enabling 100 C Fast-Charging for Sodium-Ion Batteries. Adv. Energy Mater. 2023, 13, 2300790. [Google Scholar] [CrossRef]
Scheme 1. Examples of bismacrocycles and related classification.
Scheme 1. Examples of bismacrocycles and related classification.
Molecules 28 06043 sch001
Figure 1. The 1 forms 2 with lanthanide cations, and 1 after UV illumination can further generate 3 with lanthanide cations, and 2 transforms to 3 after UV illumination, and then changes to 2 after heating. Reprinted with permission from ref. [39], copyright 2017 American Chemical Society.
Figure 1. The 1 forms 2 with lanthanide cations, and 1 after UV illumination can further generate 3 with lanthanide cations, and 2 transforms to 3 after UV illumination, and then changes to 2 after heating. Reprinted with permission from ref. [39], copyright 2017 American Chemical Society.
Molecules 28 06043 g001
Figure 2. Supramolecular assembly 5 formed between 4 and 6 through host-guest interaction and 7. Reprinted with permission from ref. [40], copyright 2018 Royal Society of Chemistry.
Figure 2. Supramolecular assembly 5 formed between 4 and 6 through host-guest interaction and 7. Reprinted with permission from ref. [40], copyright 2018 Royal Society of Chemistry.
Molecules 28 06043 g002
Figure 3. Schematic diagrams of light-controlled interchangeable supramolecular assembly 10 or 11 constructed between trans- or cis-8 and 9. Reprinted with permission from ref. [41], copyright 2019 Royal Society of Chemistry.
Figure 3. Schematic diagrams of light-controlled interchangeable supramolecular assembly 10 or 11 constructed between trans- or cis-8 and 9. Reprinted with permission from ref. [41], copyright 2019 Royal Society of Chemistry.
Molecules 28 06043 g003
Figure 4. Representation of the stimuli-responsive chirality transfer and FRET between chiral donor 14 and achiral acceptor 12 in 13. Reprinted with permission from ref. [42], copyright 2019 Royal Society of Chemistry.
Figure 4. Representation of the stimuli-responsive chirality transfer and FRET between chiral donor 14 and achiral acceptor 12 in 13. Reprinted with permission from ref. [42], copyright 2019 Royal Society of Chemistry.
Molecules 28 06043 g004
Figure 5. (a) Preparation of 16 from 15; (b) the stronger ductility as well as adhesion of 17. Reprinted with permission from ref. [43], copyright 2019 American Chemical Society.
Figure 5. (a) Preparation of 16 from 15; (b) the stronger ductility as well as adhesion of 17. Reprinted with permission from ref. [43], copyright 2019 American Chemical Society.
Molecules 28 06043 g005
Figure 6. Schematic representation of self-cross-linking supramolecular polymer network formation with 18. Reprinted with permission from ref. [44], copyright 2020 American Chemical Society.
Figure 6. Schematic representation of self-cross-linking supramolecular polymer network formation with 18. Reprinted with permission from ref. [44], copyright 2020 American Chemical Society.
Molecules 28 06043 g006
Figure 7. Chemical structure of 19, 20, 21, and 25 (a); schematic illustration of spherical nanoparticles (24, 27) formed with 23 and 26 assemblies (b), also the photocontrolled energy transfer process in binary assembly 24 (c) and ternary assembly 27 (d). Reprinted with permission from ref. [49], copyright 2017 Wiley-VCH.
Figure 7. Chemical structure of 19, 20, 21, and 25 (a); schematic illustration of spherical nanoparticles (24, 27) formed with 23 and 26 assemblies (b), also the photocontrolled energy transfer process in binary assembly 24 (c) and ternary assembly 27 (d). Reprinted with permission from ref. [49], copyright 2017 Wiley-VCH.
Molecules 28 06043 g007
Figure 8. Chemical structure of 28, 30, and schematic illustration of their assembly. Reprinted with permission from ref. [50], copyright 2017 Wiley-VCH.
Figure 8. Chemical structure of 28, 30, and schematic illustration of their assembly. Reprinted with permission from ref. [50], copyright 2017 Wiley-VCH.
Molecules 28 06043 g008
Figure 9. Chemical structure of 32, 33, and schematic representation of intercalation and photocleavage of DNA using 34. Reprinted with permission from ref. [51], copyright 2018 American Chemical Society.
Figure 9. Chemical structure of 32, 33, and schematic representation of intercalation and photocleavage of DNA using 34. Reprinted with permission from ref. [51], copyright 2018 American Chemical Society.
Molecules 28 06043 g009
Figure 10. Schematic representation of the preparation from host-guest interactions between 36 and the bifluorescent-emitting VIE guest 35 to generate supramolecular aggregates 37. Reprinted with permission from ref. [53], copyright 2018 Wiley-VCH.
Figure 10. Schematic representation of the preparation from host-guest interactions between 36 and the bifluorescent-emitting VIE guest 35 to generate supramolecular aggregates 37. Reprinted with permission from ref. [53], copyright 2018 Wiley-VCH.
Molecules 28 06043 g010
Figure 11. Schematic representation of the helix interconversion of C60-appended polyacetylene (40) directed by molecular recognition of 38 and 39. Reprinted with permission from ref. [54], copyright 2020 Royal Society of Chemistry.
Figure 11. Schematic representation of the helix interconversion of C60-appended polyacetylene (40) directed by molecular recognition of 38 and 39. Reprinted with permission from ref. [54], copyright 2020 Royal Society of Chemistry.
Molecules 28 06043 g011
Figure 12. Schematic diagram of the light synthesis of 43 and 44 from (E,E)-42 with the light-regulated length of (E,E)-41 and its derivatives. Reprinted with permission from ref. [55], copyright 2023 Royal Society of Chemistry.
Figure 12. Schematic diagram of the light synthesis of 43 and 44 from (E,E)-42 with the light-regulated length of (E,E)-41 and its derivatives. Reprinted with permission from ref. [55], copyright 2023 Royal Society of Chemistry.
Molecules 28 06043 g012
Figure 13. The structure of 45 and its solid-phase fluorescence properties. Reprinted with permission from ref. [56], copyright 2020 Wiley-VCH.
Figure 13. The structure of 45 and its solid-phase fluorescence properties. Reprinted with permission from ref. [56], copyright 2020 Wiley-VCH.
Molecules 28 06043 g013
Figure 14. Schematic illustration of the controllable light-harvesting nanosystem based on photo-modulation of the energy transfer pathway. Reprinted with permission from ref. [57], copyright 2021 Elsevier.
Figure 14. Schematic illustration of the controllable light-harvesting nanosystem based on photo-modulation of the energy transfer pathway. Reprinted with permission from ref. [57], copyright 2021 Elsevier.
Molecules 28 06043 g014
Figure 15. Schematic representation of the controlled assembly and photocatalytic process between 50 and guest 51–53. Reprinted with permission from ref. [58], copyright 2023 Wiley-VCH.
Figure 15. Schematic representation of the controlled assembly and photocatalytic process between 50 and guest 51–53. Reprinted with permission from ref. [58], copyright 2023 Wiley-VCH.
Molecules 28 06043 g015
Figure 16. Chemical structure of 55, 56, and schematic illustration of the supramolecular polymerization of 57 and 59 (inset: photo of the supramolecular gels 57 and 59 under 365 nm UV light). Reprinted with permission from ref. [59], copyright 2022 American Chemical Society.
Figure 16. Chemical structure of 55, 56, and schematic illustration of the supramolecular polymerization of 57 and 59 (inset: photo of the supramolecular gels 57 and 59 under 365 nm UV light). Reprinted with permission from ref. [59], copyright 2022 American Chemical Society.
Molecules 28 06043 g016
Figure 17. Chemical structure of 60, 61, 63, and schematic representation of the polymerisation degree adjustment between 60 and 61 via the introduction of 63, a pH-responsive competitive guest. Reprinted with permission from ref. [60], copyright 2017 Wiley-VCH.
Figure 17. Chemical structure of 60, 61, 63, and schematic representation of the polymerisation degree adjustment between 60 and 61 via the introduction of 63, a pH-responsive competitive guest. Reprinted with permission from ref. [60], copyright 2017 Wiley-VCH.
Molecules 28 06043 g017
Figure 18. (a) Structure of Bisheteracalixarenes (65–69) and 70; (b) schematic representation of 67 self-assembling with 70 to form coherent particles via anion-π interactions. Reprinted with permission from ref. [61], copyright 2019 American Chemical Society.
Figure 18. (a) Structure of Bisheteracalixarenes (65–69) and 70; (b) schematic representation of 67 self-assembling with 70 to form coherent particles via anion-π interactions. Reprinted with permission from ref. [61], copyright 2019 American Chemical Society.
Molecules 28 06043 g018
Figure 19. The structure of 71 and its luminescent properties. Reprinted with permission from ref. [62], copyright 2023 Wiley-VCH.
Figure 19. The structure of 71 and its luminescent properties. Reprinted with permission from ref. [62], copyright 2023 Wiley-VCH.
Molecules 28 06043 g019
Figure 20. Synthesis of 73, 76–79 and schematic diagram of the preparation of AIE fluorescent nanomaterial 74 and 75 (a), schematic diagram of the preparation of AIE fluorescent nanomaterial (b). Reprinted with permission from refs. [64,65], copyright 2019 American Chemical Society and 2021 Elsevier.
Figure 20. Synthesis of 73, 76–79 and schematic diagram of the preparation of AIE fluorescent nanomaterial 74 and 75 (a), schematic diagram of the preparation of AIE fluorescent nanomaterial (b). Reprinted with permission from refs. [64,65], copyright 2019 American Chemical Society and 2021 Elsevier.
Molecules 28 06043 g020
Figure 21. Structure of 80 and schematic representation of 80 as a fluorescent sensor to detect 81. Reprinted with permission from ref. [66], copyright 2022 Elsevier.
Figure 21. Structure of 80 and schematic representation of 80 as a fluorescent sensor to detect 81. Reprinted with permission from ref. [66], copyright 2022 Elsevier.
Molecules 28 06043 g021
Figure 22. (a) Structure of 82; (b) colocalization images of 82 with LysoTracker Red in RAW 264.7 cells; (c) photothermal images of 82 (0.1 mM) in the presence of E. coli under irradiation at 1 W cm−2 for different times. (b,c) Reprinted with permission from ref. [67], copyright 2022 Wiley-VCH.
Figure 22. (a) Structure of 82; (b) colocalization images of 82 with LysoTracker Red in RAW 264.7 cells; (c) photothermal images of 82 (0.1 mM) in the presence of E. coli under irradiation at 1 W cm−2 for different times. (b,c) Reprinted with permission from ref. [67], copyright 2022 Wiley-VCH.
Molecules 28 06043 g022
Figure 23. Confocal fluorescence imaging of the photothermal ablation induced by 88283 upon irradiation at 1064 nm (top left); UV/Vis-NIR spectra of 84283, 85283, 86283, and 88283. Reprinted with permission from ref. [68], copyright 2023 Wiley-VCH.
Figure 23. Confocal fluorescence imaging of the photothermal ablation induced by 88283 upon irradiation at 1064 nm (top left); UV/Vis-NIR spectra of 84283, 85283, 86283, and 88283. Reprinted with permission from ref. [68], copyright 2023 Wiley-VCH.
Molecules 28 06043 g023
Figure 24. Schematic synthesis of three TPE-containing pyridinium bismacrocycles 89–91. Reprinted with permission from ref. [69], copyright 2023 Elsevier.
Figure 24. Schematic synthesis of three TPE-containing pyridinium bismacrocycles 89–91. Reprinted with permission from ref. [69], copyright 2023 Elsevier.
Molecules 28 06043 g024
Figure 25. (a) Construction of anionic regulated multi-component self-assembly structures based on biscalix[4]pyrrole 92; (b) The structure of 93 and its assembly with double N-oxides; (c) The structure of 94 and its assembly structure with two SO42−; (d) 95 specific recognition of F; (e) 96 forms a 1:2 complex with F through different pathways. (ae) Reprinted with permission from ref. [71,72,73,74,75], copyright 2017 Royal Society of Chemistry and 2017, 2020, 2022 American Chemical Society.
Figure 25. (a) Construction of anionic regulated multi-component self-assembly structures based on biscalix[4]pyrrole 92; (b) The structure of 93 and its assembly with double N-oxides; (c) The structure of 94 and its assembly structure with two SO42−; (d) 95 specific recognition of F; (e) 96 forms a 1:2 complex with F through different pathways. (ae) Reprinted with permission from ref. [71,72,73,74,75], copyright 2017 Royal Society of Chemistry and 2017, 2020, 2022 American Chemical Society.
Molecules 28 06043 g025
Figure 26. Synthesis of 97–98 and related fluorescence properties. Reprinted with permission from ref. [76], copyright 2020 Royal Society of Chemistry.
Figure 26. Synthesis of 97–98 and related fluorescence properties. Reprinted with permission from ref. [76], copyright 2020 Royal Society of Chemistry.
Molecules 28 06043 g026
Figure 27. The structure of the 99 and a schematic diagram of its π-electronic system. Reprinted with permission from ref. [77], copyright 2019 American Chemical Society.
Figure 27. The structure of the 99 and a schematic diagram of its π-electronic system. Reprinted with permission from ref. [77], copyright 2019 American Chemical Society.
Molecules 28 06043 g027
Figure 28. The synthetic strategy of 100–102 and its structure conversion. Reprinted with permission from ref. [78], copyright 2021 Wiley-VCH.
Figure 28. The synthetic strategy of 100–102 and its structure conversion. Reprinted with permission from ref. [78], copyright 2021 Wiley-VCH.
Molecules 28 06043 g028
Figure 29. Synthesis of 103 and its single crystal structure. Reprinted with permission from ref. [79], copyright 2019 Wiley-VCH.
Figure 29. Synthesis of 103 and its single crystal structure. Reprinted with permission from ref. [79], copyright 2019 Wiley-VCH.
Molecules 28 06043 g029
Figure 30. (a) The structure of 104–107; (b) The analogue structure of 104; (c) QCM airborne analyte sensing study of the title compounds as affinity materials relative to a passivated surface (denoted as “no affinity material”); (d) X-ray crystal structures of [9]CPP and 105. (bd) Reprinted with permission from ref. [81], copyright 2020 American Chemical Society.
Figure 30. (a) The structure of 104–107; (b) The analogue structure of 104; (c) QCM airborne analyte sensing study of the title compounds as affinity materials relative to a passivated surface (denoted as “no affinity material”); (d) X-ray crystal structures of [9]CPP and 105. (bd) Reprinted with permission from ref. [81], copyright 2020 American Chemical Society.
Molecules 28 06043 g030
Figure 31. (a) The structure of 108 and the peanut-like complexation between 108 and 109 in a 1:2 ratio; (b) (b1) Structure and properties of 110; (b2) Fluorescence spectra of 110 in the solvent with different THF/H2O ratios; (b3) CIE 1931 chromaticity diagram of 110 in THF/H2O mixtures; (b4) r Emission colour changes of 110 from cyan to red in aqueous THF with fw = 0–99 vol% under 365 nm UV light. Reprinted with permission from ref. [82,83], copyright 2021 Wiley-VCH and 2022 Nature Publishing Group.
Figure 31. (a) The structure of 108 and the peanut-like complexation between 108 and 109 in a 1:2 ratio; (b) (b1) Structure and properties of 110; (b2) Fluorescence spectra of 110 in the solvent with different THF/H2O ratios; (b3) CIE 1931 chromaticity diagram of 110 in THF/H2O mixtures; (b4) r Emission colour changes of 110 from cyan to red in aqueous THF with fw = 0–99 vol% under 365 nm UV light. Reprinted with permission from ref. [82,83], copyright 2021 Wiley-VCH and 2022 Nature Publishing Group.
Molecules 28 06043 g031
Figure 32. (a) The structure of 111 and 112 (b) Solid-state structures of a complex of 112 and C60 in a 1:1 ratio. Reprinted with permission from ref. [85], copyright 2021 American Chemical Society.
Figure 32. (a) The structure of 111 and 112 (b) Solid-state structures of a complex of 112 and C60 in a 1:1 ratio. Reprinted with permission from ref. [85], copyright 2021 American Chemical Society.
Molecules 28 06043 g032
Figure 33. Crystal structure of a complex of 113 formed in a 1:2 ratio with C60 or C70. Reprinted with permission from ref. [86], copyright 2022 Wiley-VCH.
Figure 33. Crystal structure of a complex of 113 formed in a 1:2 ratio with C60 or C70. Reprinted with permission from ref. [86], copyright 2022 Wiley-VCH.
Molecules 28 06043 g033
Table 1. Summary of bismacrocycle shown above.
Table 1. Summary of bismacrocycle shown above.
BismacrocycleSub-ClassificationRepresentative StructureApplications
OxabismacrocycleBis(crown ether)Molecules 28 06043 i001Self-assembly,
Supramolecular polymer,
Luminescent material,
chiral luminescent material.
BiscyclodextrinMolecules 28 06043 i002Self-assembly,
Supramolecular polymer,
Luminescent material,
chiral luminescent material,
anticancer active materials.
BiscalixareneMolecules 28 06043 i003Reversible self-assembly,
supramolecular polymer.
BispillarareneMolecules 28 06043 i004(Stimulus-responsive) luminescent material
BishelicareneMolecules 28 06043 i005CPL-active supramolecular gels
BiscucurbiturilMolecules 28 06043 i006Controlled supramolecular
polymer
BisheteracalixarenesMolecules 28 06043 i007anion-π interaction-directed
self-assembly
Other oxabismacrocycleMolecules 28 06043 i008Chiral luminescent material
AzabismacrocyclePyridinium bismacrocycleMolecules 28 06043 i009Highly efficient luminescent material, multicomponent assemblies, AIE fluorescent nanomaterial, pollutant detection.
Biscalix[4]pyrrolesMolecules 28 06043 i010Two/multi-component
self-assembly, anion recognition.
Imidazolium bismacrocycleMolecules 28 06043 i011Mechanochromic and photochromic luminescence compounds.
AzabiscycloparaphenyleneMolecules 28 06043 i012Complex chiral compounds,
conformationally interchangeable fluorescent compounds.
BiscycloparaphenyleneBiscycloparaphenyleneMolecules 28 06043 i013Complex chiral (luminescent) compounds, VOC adsorbent materials, self-assembly,
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

Chen, X.-L.; Yu, S.-Q.; Huang, X.-H.; Gong, H.-Y. Bismacrocycle: Structures and Applications. Molecules 2023, 28, 6043. https://doi.org/10.3390/molecules28166043

AMA Style

Chen X-L, Yu S-Q, Huang X-H, Gong H-Y. Bismacrocycle: Structures and Applications. Molecules. 2023; 28(16):6043. https://doi.org/10.3390/molecules28166043

Chicago/Turabian Style

Chen, Xu-Lang, Si-Qian Yu, Xiao-Huan Huang, and Han-Yuan Gong. 2023. "Bismacrocycle: Structures and Applications" Molecules 28, no. 16: 6043. https://doi.org/10.3390/molecules28166043

Article Metrics

Back to TopTop