Next Article in Journal
Tannins from Acacia mearnsii De Wild as a Sustainable Alternative for the Development of Latent Fingerprints
Previous Article in Journal
Macrocyclic Azopyrrole: Synthesis, Structure and Fluoride Recognition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Study of the Inclusion Complex Formed Between Cucurbit[8]uril and N,4-Di(pyridinyl)benzamide Derivative

1
College of Chemical and Biological Engineering, Shandong University of Science and Technology, Qingdao 266590, China
2
Guangxi Key Laboratory of Advanced Structural Materials and Carbon Neutralization, School of Materials and Environment, Guangxi Colleges and Universities Key Laboratory of Eco-Friendly Materials and Ecological Restoration, Guangxi Minzu University, Nanning 530105, China
3
Commonwealth Scientific and Industrial Research Organisation (CSIRO), Mineral Resources, P.O. Box 218, Lindfield, NSW 2070, Australia
*
Authors to whom correspondence should be addressed.
Organics 2025, 6(2), 26; https://doi.org/10.3390/org6020026
Submission received: 27 April 2025 / Revised: 7 June 2025 / Accepted: 12 June 2025 / Published: 17 June 2025

Abstract

:
The interaction between cucurbit[8]uril (Q[8]) and the guest 1-methyl-4-(4-(1-methylpyridin-1-ium-4-yl)benzamido)pyridin-1-ium (PB2+) has been thoroughly investigated. Multiple techniques were employed, including 1H NMR spectroscopy, mass spectrometry, isothermal titration calorimetry (ITC), UV–vis absorption spectrophotometry, and quantum chemistry calculations. The experimental results and calculation analysis have clearly shown that in aqueous solution, the host Q[8] preferentially encapsulates the phenylpyridinium salt moiety of the PB2+ guest within its hydrophobic cavity, forming a 1:2 inclusion complex.

Graphical Abstract

1. Introduction

Cucurbit[n]urils (Q[n]s) are a class of macrocyclic compounds composed of glycoluril units connected by methylene bridges [1,2,3]. Their unique hydrophobic cavities and polar carbonyl ports endow them with excellent host–guest recognition abilities, which have attracted extensive attention in supramolecular chemistry, drug delivery, biomedicine, and functional materials [4,5,6,7]. Based on the cavity size of Q[n], cucurbit[8]uril (Q[8]) has a relatively large cavity size (~8.8 Å). This large cavity enables Q[8] to exhibit a unique molecular recognition performance and allows specific guest molecules to dimerize in its cavity in a controlled manner [8,9,10,11,12]. Two guest molecules or large-volume functional groups, such as phenyl, naphthyl, indole, quinoline, and larger aromatic rings, can be included in the cavity of Q[8] to form homoternary inclusion complexes. These complexes have become an important platform for constructing multi-component complex supramolecular systems [13,14,15,16]. In recent years, research on the host–guest interaction based on Q[8] has been extended to ionic aromatic compounds, especially derivatives containing pyridine or benzene rings [17,18,19,20]. These molecules can form stable inclusion complexes with Q[8] through hydrophobic interactions, π-π stacking, and electrostatic effects. However, their binding modes and structural regulation mechanisms still need further exploration [21,22,23,24]. For instance, Zhang’s group constructed a variety of supramolecular self-assembly systems for photocatalytic applications through the self-assembly of linear molecules and Q[8] [25,26]. Liu’s group constructed a variety of supramolecular room-temperature phosphorescence systems based on 4-(4-bromophenyl)-pyridinium derivatives and Q[8] [27,28,29], which have shown good results in many fields such as biological imaging, molecular detection, drug delivery, and so on. Li’s group synthesized a tetrahedral molecule featuring tetraphenylmethane as its core and combined it with Q[8] to create a novel water-soluble three-dimensional supramolecular organic framework. This framework was utilized as a broad-spectrum reversal agent for heparin anticoagulation [30]. Moreover, a self-assembly process utilizing hexa-armed [Ru(bpy)3]2+ derivatives and Q[8] constructs another three-dimensional cubic framework. This framework can serve as a heterogeneous catalyst for hydrogen production and organic reactions [31]. Xiao’s group has constructed a series of Q[8]-based ternary self-assemblies and achieved a series of excellent results in the fields of fluorescence sensing and molecular recognition [32,33]. Overall, non-covalent interactions-based supramolecular functional materials have shown great potential in many fields [34,35].
The N,4-dipyridylbenzamide derivative, as a bi-cationic aromatic molecule containing phenylpyridinium salt structural units, is an ideal object for studying the ternary supramolecular self-assembly of Q[8] [11,29]. Phenylpyridinium salt is one of the building blocks for supramolecular network structures. In the field of supramolecular self-assembly, we have conducted some related research and achieved certain results. A series of supramolecular polymer systems were constructed by using the antiparallel arrangement of naphthyl and styrene pyridine dimers in the Q[8] cavity [36,37,38,39]. For instance, an irreversible covalent component was created through the photocatalytic photodimerization of carbon–carbon double bonds, which was utilized to construct the first highly water-soluble 3D supramolecular covalent organic framework. When the framework material supported a POM catalyst and a Ru2+ complex photosensitizer to form a composite catalytic system, the electron transfer of protons to H2 was significantly enhanced [40]. Consequently, we believe that cucurbit[n]uril is one of the ideal macrocyclic host molecules for constructing a supramolecular network system.
Herein, in this study, we selected Q[8] and the guest 1-methyl-4-(4-(1-methylpyridin-1-ium-4-yl)benzamido)pyridin-1-ium (PB2+) (Figure 1) as the research objects. We systematically examined the host–guest interaction between Q[8] and PB2+ in an aqueous solution, employing techniques such as 1H NMR, ITC, UV–vis, ESI–MS, and quantum chemical calculations. The experiments revealed that Q[8] selectively encapsulated the phenylpyridine salt of PB2+ within its hydrophobic cavity, forming a stable inclusion compound with a stoichiometric ratio of 1:2. Additionally, it engaged in polar interactions with the pyridine group through the carbonyl port. This research not only reveals the supramolecular binding mechanism between Q[8] and PB2+ molecules for the first time but also offers a theoretical and experimental foundation for the development of multi-component functional materials based on Q[n]s and phenylpyridine salts.

2. Materials and Methods

Cucurbit[8]uril was prepared and purified following our previously published procedure [41,42]. Add 500 g of semicarbazide and 245 g of paraformaldehyde (maintaining a weight ratio of 2:1) to a 3000 mL three-necked flask. Subsequently, introduce 600 mL of ice-cold concentrated hydrochloric acid. Using a constant temperature magnetic stirrer, maintain the temperature at 90~110 °C for 4 to 6 h. Following the reaction, add approximately 1 L of distilled water to the mixture. Upon stirring with a rod, a large amount of white solid (crude product Q[8]) is observed to form. The mixture is then left to stand overnight. The solid is subsequently purified through recrystallization (3M hydrochloric acid solution).
  • Synthesis of compound PB2+
Organics 06 00026 i001 Compound PB2+. Compound 1 (4-Aminopyridine, 0.52 g, 5.1 mmol), 2 (1.0 g, 5.0 mmol), EDCI (N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride) (0.96 g, 5.0 mmol), and HOBT (1-Hydroxybenzotriazole) (0.68 g, 5.0 mmol) were dissolved in DMF (10 mL) and stirred at 60 °C for 24 h. After the reaction was completed, it was allowed to cool to room temperature. The resulting precipitate was filtered, washed with ethanol three times, and dried in vacuum to obtain compound 3 as a white solid.
Compound 3 (0.58 g, 2.1 mmol) and methyl iodide (0.71 g, 4.0 mmol) were added to MeCN (15 mL), and the mixture was stirred at 80 °C for 24 h. After the reaction was completed, it was allowed to cool to room temperature. The precipitate was filtered, washed with a small amount of MeCN, and vacuum-dried to obtain a crude product. The solid was then dissolved in a small amount of water, and a saturated aqueous solution of ammonium hexafluorophosphate was added dropwise to the solution. The compound was heated until it was completely dissolved. The precipitate was filtered, washed with cold water, and vacuum-dried to obtain a pale white solid product. Finally, the product was dissolved in a small amount of acetonitrile. A saturated aqueous solution of tetrabutylammonium chloride (TBA-Cl) was added dropwise to this solution, and the resulting precipitate was filtered, washed with cold acetonitrile, and further dried under vacuum to obtain the compound. PB2+ (0.95 g, 62%). M.p. > 280 °C (decomp). 1H NMR (400 MHz, DMSO-d6): δ 11.10 (s, 1H), 9.14 (d, J = 4.0 Hz, 2H), 8.62 (d, J = 2.0 Hz, 2H), 8.50 (d, J = 2.0 Hz, 2H), 8.30 (dd, J = 8 Hz, 2H. J = 8 Hz, 2H), 7.92 (d, J = 8.4 Hz, 2H), 4.39 (s, 3H). 13C NMR (100 MHz, DMSO-d6): δ = 166.8, 153.3, 152.2, 146.5, 146.3, 138.0, 136.0, 129.9, 128.9, 125.2, 116.1, 47.8, 47.0. HRMS (ESI): Calcd for C19H19N3O2+ (305.1528) [M-2Cl]2+/2 = 305.1528/2 = 152.5764. Found: 152.5741. We have only analyzed the double charge mode of the molecule (Figure S6).
  • Nuclear magnetic resonance measurements.
Deuterated water was used as the solvent to prepare a 2.0 mM deuterated aqueous solution of PB2+ and a supersaturated deuterated aqueous solution of Q[8]. Different proportions of Q[8] were then added to PB2+ to create nuclear magnetic resonance tubes with varying host–guest molar ratios (Q[8]/PB2+). All NMR spectra were measured by VARIAN INOVA-400 spectrometer at 25 °C, and the resulting nuclear magnetic resonance data for the different host–guest ratios were collected and analyzed.
  • UV–vis absorption.
The UV–vis absorption spectrum of the host–guest complex was tested using Agilent 8453 spectrophotometer (1 cm quartz cells) at 25 °C. Accurately prepare a 1.0 × 10−3 mol/L aqueous solution of PB2+, transfer 0.2 mL into a 10 mL volumetric flask, and add varying amounts of Q[8] aqueous solution with a concentration of 1.0 × 10−4 mol/L to prepare a series of host–guest interaction test solutions with different molar ratios (0, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 1.0, 1.1, 1.2, 1.3, 1.4, 1.5). Adjust the volume to 10 mL, allow it to stand for half an hour, and measure its electronic absorption spectrum and fluorescence emission spectrum at room temperature. The absorption spectra were obtained at a concentration of 2.0 × 10−5 mol L−1 PB2+ and at varying Q[8] concentrations for the PB2+@Q[8] system at 25 °C.
  • Isothermal titration calorimetry (ITC) experiments.
The association constants and thermodynamic parameters for the 1:2 complexation of PB2+ with Q[8] were determined using isothermal titration calorimetry with a Malvern MicroCal VP-ITC instrument. We did not conduct repetitive experiments in the ITC lab; we only performed the experiment once. In a typical procedure, an aqueous solution (0.1 mM) of Q[8] was placed in the sample cell (170 μL), and a solution (1.0 mM) of PB2+ was added in a series of 20 injections (2 μL) at 150 s intervals, with the heat evolved recorded at 298.15 K. The data is fitted by MicroCal ITC analysis software (Origin 7 SR4 v7.0552). We utilized the Multiple Sites mode for fitting. The results of the Multiple Sites mode fitting are detailed in the supporting materials, based on experimental data (Table S1).
  • High-resolution electrospray mass spectrometer (HR ESI–MS).
HR ESI–MS was tested on a Bruker 5 MicrOTOF II. The ESI–MS experiments were conducted by adding a PB2+ solution (1.0 mM, 1.0 mL) to a Q[8] solution (0.1 mM, 10.0 mL). The resulting solution concentration was approximately 0.1 mM (Q[8]/PB2+ = 1:2).
  • Quantum chemistry calculations.
Utilize the Gaussian 09 software package to perform all calculations. With the assistance of the HyperChem Release 7.52 software package, the initial geometries of all structures are constructed. Becke’s three-parameter mixed functional, combined with the correlation functional of Lee, Yang, and Parr (B3LYP), is employed to optimize the elements C, H, N, and O at the theoretical level of 6-311G(d, p).

3. Results and Discussion

Binding behaviour in aqueous solution: 1H NMR spectroscopy. The interaction between PB2+ and Q[8] was explored through 1H NMR. It has been previously reported that the 1H NMR peak of the guest proton within the low polarizability cavity of Q[8] has experienced a field shift. Simultaneously, due to proton unmasking, its interaction with the carbonyl oxygen of Q[8] results in a low field displacement. The 1H NMR spectroscopy measurements clearly indicate that PB2+ forms a host–guest inclusion complex with the Q[8] host. Figure 2 presents the titration 1H NMR spectra, where distinct changes in the proton signals were observed compared with those of the free guest PB2+. Specifically, the H5 and H6 proton signals gradually shifted upfield by approximately 1.6 and 0.1 ppm, respectively. Additionally, the H3 and H4 proton signals shifted upfield by approximately 0.7 and 1.3 ppm, respectively. The chemical shift in protons to a higher field is due to the shielding effect of electron-rich regions in Q[8]. It strongly suggests that the N-dimethyl and benzene ring moiety of the guest PB2+ are deeply included in the cavity of Q[8] due to the formation of the complex PB2+@Q[8]. In contrast, under a 1:1 host–guest ratio in D2O, the signal for the H1 and H2 protons showed an obvious downfield shift of approximately 0.2 ppm. Moreover, the H7 and H8 proton signals for the N-methyl also exhibited a downfield shift. These experimental results indicate that the phenylpyridinium salt is included in the low-polarizability cavity of Q[8]. At the same time, the amidopyridinium salt and the N-methyl on the pyridine ring are mainly located in a deshielding environment, resulting in the formation of a 1:1 host–guest inclusion complex in aqueous solution. The formation of the inclusion complex between PB2+ and Q[8] was further confirmed by 1H-1H COSY NMR spectra (Figure S1).
UV–vis absorbance spectrophotometry. First, the changes in absorption intensity during the complexation of Q[8] with PB2+ were investigated. As shown in Figure 3a, in an aqueous solution, PB2+ exhibits an absorption maximum centred at 298 nm. When Q[8] is added, the complexation of PB2+ leads to a decrease in absorption from 0.99 to 0.80, along with a bathochromic shift from 298 to 301 nm. As the amount of Q[8] increases, the intensity of the absorption maximum peak gradually decreases, which is a clear indication of an interaction between PB2+ and Q[8]. Moreover, the shape of the absorption spectrum of Q[8] remains unchanged, suggesting a non-covalent interaction between PB2+ and Q[8]. With a further increase in Q[8], the UV–vis absorption intensity of PB2+ no longer shows obvious changes. These experimental results imply the formation of supramolecular complexes PB2+@Q[8]. Based on the continuous variation method, the Job plots clearly demonstrate that the UV–−vis spectra data of PB2+ fit well with a 1:2 stoichiometry for the host–guest inclusion complexes (Figure 3b). Furthermore, the binding stoichiometry was determined through Job plot analysis at a constant total concentration of the host and guest molecules. By plotting the changes in absorption intensity (ΔA) against the molar fraction of PB2+, a peak was observed at a molar fraction of 0.66, indicating a 1:2 stoichiometry for the Q[8]:PB2+ inclusion complex. (Figure S2).
Isothermal titration calorimetry (ITC). ITC is one of the most effective techniques for studying supramolecular interactions. It can provide not only thermodynamic parameters (ΔS, ΔH, ΔG) but also the binding constant (Ka). To gain a deeper understanding of the nature of the host–guest complexations between Q[8] and PB2+, we conducted ITC experiments, as shown in Figure 4. We only conducted one ITC experiment, and did not conduct many repeated experiments. First, from the ΔH and TΔS values presented in Table S1, it was evident that the formation of homoternary complexes was enthalpically driven. The observed negative enthalpy change (ΔH1 = −60.70 kJ∙mol−1, ΔH2 = −80.89 kJ∙mol−1 for PB2+@Q[8]) was likely due to the cooperativity of weak interactions. Secondly, the intermolecular complexation between Q[8] and PB2+ were driven by favorable enthalpy changes. Based on the corresponding experimental results, we also obtained the association constant of Ka = 4.413 × 109 M−2 for the interaction with PB2+. The high binding constant indicates a relatively strong host–guest interaction between Q[8] and PB2+, suggesting the formation of a stable homoternary complex PB2+@Q[8] in aqueous solution.
Electrospray ionization–mass spectrometry (ESI–MS). In addition, the 2:1 complex PB2+@Q[8] was further characterized by ESI–MS. The presence of a peak at m/z = 484.6730, corresponding to PB2+@Q[8] (calculated for [2(C19H19ON3Cl2)@(C48H48N32O16)-4Cl]4+/4, 484.6746) (Figure 5). This observation confirmed that the binding ratio between Q[8] and PB2+ is 1:2. This result is consistent with the findings from 1H NMR spectroscopy, ITC, and UV–vis absorption.
Quantum chemistry calculations. To gain a more in-depth proof of the host–guest complexation between Q[8] and PB2+, molecular modeling was utilized. Density functional theory (DFT) calculations were performed at the B3LYP/6-31G(d, p) level to obtain structural information on supramolecular self-assembly. Figure 6 depicts the low energy conformations of the host–guest complexes. These conformations clearly indicate that PB2+ forms stable 1:2 inclusion complexes. Specifically, the phenylpyridinium salt of PB2+ is situated within the cavity of Q[8]. Meanwhile, the amidopyridinium salt and the N-methyl group on the pyridine ring are located just outside the host’s portal. This finding is in excellent agreement with the NMR results.

4. Conclusions

In summary, we have conducted an investigation into a novel supramolecular host–guest complex formed between the N,4-di(pyridinyl)benzamide derivative (PB2+) and Q[8] in aqueous solutions. Multiple analytical techniques, including 1H NMR spectroscopy, ESI–mass spectrometry, UV–vis absorbance spectrophotometry, and ITC, were employed for this study. The results obtained from these analyses revealed that a PB2+@Q[8] (2:1) inclusion complex was formed. Specifically, the phenylpyridinium salt of the PB2+ was found to be located within the cavity of Q[8], while the N-methyl on the pyridine ring was situated just outside the host’s portal. To the best of our knowledge, the present result represents the first example of structurally characterized Q[n]s-isoniazid host–guest complexes. Based on this study, we conclude that the Q[8] host and the molecules containing phenylpyridinium units can serve as building blocks to construct a diverse range of supramolecular network structures.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/org6020026/s1, General Information; Synthesis of compound PA; Nuclear magnetic resonance measurements; UV–Vis absorption; ITC experiments; HR ESI–MS; Quantum chemistry calculations [43,44]; Figure S1: 1H-1H COSY spectra of PB2+@Q[8] in D2O at 25 °C; Figure S2: Job’s plot obtained from the absorption spectra of the mixtures of PB2+ and Q[8] ([PB2+] + [Q[8]] = 50 μM) in water at 25 °C; Figure S3: 1H NMR spectra of PB2+ in D2O at 25 °C; Figure S4: 13C NMR spectra of PB2+ in DMSO at 25 °C; Figure S5: 1H-1H COSY spectra of PB2+ in D2O at 25 °C; Figure S6: HR-MS (ESI): Calcd for PB2+ C19H19N3O2+: 305.1517 [M-2Cl]2+. Found: [M-2Cl]2+/2 = 152.5741; Figure S7: 1H NMR spectra of compound 3 in DMSO at 25 °C; Figure S8: 13C NMR spectra of compound 3 in DMSO at 25 °C; Figure S9: 1H NMR spectra of Q[8] in D2O at 25 °C. Table S1: ITC measurements of the thermodynamics of 2PB2+@Q[8] interactions in aqueous solution at 298.15 K.

Author Contributions

Z.W.: Conceptualization. M.Y.: Data curation and Conceptualization. W.Y.: Data curation. Z.G.: Writing—original draft. H.Z.: Writing—original draft. G.W.: Validation. J.S.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of Shandong Province (ZR2023MB138, ZR2021QB178, and ZR2021QB197); Guangxi Colleges and Universities Key Laboratory of Eco-friendly Materials and Ecological Restoration, Guangxi Minzu University (GXAMCN23-9); and Key Laboratory of Natural Polymer Function Material of Haikou City, Hainan Normal University (HKTRKT-202402).

Data Availability Statement

No data was used for the research described in the article.

Acknowledgments

The authors would like to thank Bo Yang at Zhengzhou University for the beneficial discussion.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Yang, H.; Yuan, B.; Zhang, X.; Scherman, O.A. Supramolecular Chemistry at Interfaces: Host-Guest Interactions for Fabricating Multifunctional Biointerfaces. Acc. Chem. Res. 2014, 47, 2106–2115. [Google Scholar] [CrossRef]
  2. Ghale, G.; Nau, W.M. Dynamically Analyte-Responsive Macrocyclic Host-Fluorophore Systems. Acc. Chem. Res. 2014, 47, 2150–2159. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, J.-X.; Chen, K.; Redshaw, C. Stimuli-responsive mechanically interlocked molecules constructed from cucurbit[n]uril homologues and derivatives. Chem. Soc. Rev. 2023, 52, 1428–1455. [Google Scholar] [CrossRef] [PubMed]
  4. Shetty, D.; Khedkar, J.K.; Park, K.M.; Kim, K. Can we beat the biotin-avidin pair: Cucurbit[7]uril-based ultrahigh affinity host-guest complexes and their applications. Chem. Soc. Rev. 2015, 44, 8747–8761. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Z.; Sun, C.; Yang, K.; Chen, X.; Wang, R. Cucurbituril-Based Supramolecular Polymers for Biomedical Applications. Angew. Chem. Int. Edit. 2022, 61, e202206763. [Google Scholar] [CrossRef]
  6. 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]
  7. Guan, W.-L.; Chen, J.-F.; Liu, J.; Shi, B.; Yao, H.; Zhang, Y.M.; Wei, T.-B.; Lin, Q. Macrocycles-assembled AIE supramolecular polymer networks. Coordin. Chem. Rev. 2024, 507, 215717. [Google Scholar] [CrossRef]
  8. Ziganshina, A.Y.; Ko, Y.H.; Jeon, W.S.; Kim, K. Stable π-dimer of a tetrathiafulvalene cation radical encapsulated in the cavity of cucurbit[8]uril. Chem. Commun. 2004, 7, 806–807. [Google Scholar] [CrossRef]
  9. Ni, X.-L.; Chen, S.; Yang, Y.; Tao, Z. Facile Cucurbit[8]uril-Based Supramolecular Approach To Fabricate Tunable Luminescent Materials in Aqueous Solution. J. Am. Chem. Soc. 2016, 138, 6177–6183. [Google Scholar] [CrossRef]
  10. Rabbani, R.; Masson, E. Probing Interactions between Hydrocarbons and Auxiliary Guests inside Cucurbit[8]uril. Org. Lett. 2017, 19, 4303–4306. [Google Scholar] [CrossRef]
  11. Wang, J.; Huang, Z.; Ma, X.; Tian, H. Visible-Light-Excited Room-Temperature Phosphorescence in Water by Cucurbit[8]uril-Mediated Supramolecular Assembly. Angew. Chem. Int. Edit. 2020, 59, 9928–9933. [Google Scholar] [CrossRef] [PubMed]
  12. Armstrong, L.; Chang, S.L.; Clements, N.; Hirani, Z.; Kimberly, L.B.; Odoi-Adams, K.; Suating, P.; Taylor, H.F.; Trauth, S.A.; Urbach, A.R. Molecular recognition of peptides and proteins by cucurbit[n]urils: Systems and applications. Chem. Soc. Rev. 2024, 53, 11519–11556. [Google Scholar] [CrossRef] [PubMed]
  13. Ma, X.-K.; Zhang, W.; Liu, Z.; Zhang, H.; Zhang, B.; Liu, Y. Supramolecular Pins with Ultralong Efficient Phosphorescence. Adv. Mater. 2021, 33, 2007476. [Google Scholar] [CrossRef]
  14. Yang, B.; Yu, S.-B.; Zhang, P.-Q.; Wang, Z.-K.; Qi, Q.-Y.; Wang, X.-Q.; Xu, X.-H.; Yang, H.-B.; Wu, Z.-Q.; Liu, Y.; et al. Self-Assembly of a Bilayer 2D Supramolecular Organic Framework in Water. Angew. Chem. Int. Edit. 2021, 60, 26268–26275. [Google Scholar] [CrossRef] [PubMed]
  15. Wu, W.; Xu, Y.; Wang, S.; Pang, Q.; Liu, S. Metal-organic rotaxane frameworks constructed from a cucurbit[8]uril-based ternary complex for the selective detection of antibiotics. Chem. Commun. 2023, 59, 5890–5893. [Google Scholar] [CrossRef]
  16. Liu, J.; Zhang, R.-Z.; Han, N.; Liu, H.; Xing, L.-B. Supramolecular organic framework nanosheets for efficient singlet oxygen generation: A multilevel energy transfer approach for photocatalytic Minisci reactions. J. Mater. Chem. A. 2025, 13, 13789–13796. [Google Scholar] [CrossRef]
  17. Li, Y.; Yan, C.; Li, Q.; Cao, L. Successive construction of cucurbit[8]uril-based covalent organic frameworks from a supramolecular organic framework through photochemical reactions in water. Sci. China Chem. 2022, 65, 1279–1285. [Google Scholar] [CrossRef]
  18. Yan, C.; Li, Q.; Miao, X.; Zhao, Y.; Li, Y.; Wang, P.; Wang, K.; Duan, H.; Zhang, L.; Cao, L. Chiral Adaptive Induction of an Achiral Cucurbit[8]uril-Based Supramolecular Organic Framework by Dipeptides in Water. Angew. Chem. Int. Edit. 2023, 62, e202308029. [Google Scholar] [CrossRef]
  19. Chen, D.-T.; Zhang, X.-W.; Xia, Q.-X.; Yang, X.-N.; Wang, C.-H.; Shan, P.-H.; Xiao, X. A novel styrylpyridine derivatives: Supramolecular assembly with cucurbit[8]uril for detection of difenzoquat, white light material and beyond. J. Mol. Liq. 2023, 385, 122414. [Google Scholar] [CrossRef]
  20. Li, Y.; Huang, Z.; Shao, A.; Wu, Z.; He, Z.; Tian, H.; Ma, X. Aqueous up-conversion organic phosphorescence and tunable dual emission in a single-molecular emitter. Chem. Sci. 2025, 16, 6290–6297. [Google Scholar] [CrossRef]
  21. Geng, Q.-X.; Wang, F.; Cong, H.; Tao, Z.; Wei, G. Recognition of silver cations by a cucurbit[8]uril-induced supramolecular crown ether. Org. Biomol. Chem. 2016, 14, 2556–2562. [Google Scholar] [CrossRef]
  22. Tabet, A.; Forster, R.A.; Parkins, C.C.; Wu, G.; Scherman, O.A. Modulating stiffness with photo-switchable supramolecular hydrogels. Polym. Chem. 2019, 10, 467–472. [Google Scholar] [CrossRef]
  23. Gao, Z.-Z.; Lin, R.L.; Bai, D.; Tao, Z.; Liu, J.X.; Xiao, X. Host-guest complexation of cucurbit[8]uril with two enantiomers. Sci. Rep. 2017, 7, 44717. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, J.; Han, X.; Wen, X.; Yu, H.; Li, B.; Wang, M.; Liu, M.; Wu, G. Chiral ring-in-ring complexes with torsion-induced circularly polarized luminescence. Chem. Sci. 2025, 16, 7858–7863. [Google Scholar] [CrossRef]
  25. Tang, B.; Xu, W.; Xu, J.-F.; Zhang, X. Transforming a Fluorochrome to an Efficient Photocatalyst for Oxidative Hydroxylation: A Supramolecular Dimerization Strategy Based on Host-Enhanced Charge Transfer. Angew. Chem. Int. Edit. 2021, 60, 9384–9388. [Google Scholar] [CrossRef]
  26. Xu, W.; Du, Y.; Ma, H.; Tang, X.; Xu, J.-F.; Zhang, X. A cucurbit[8]uril-mediated host-guest complex for red-light photocatalysis. Org. Chem. Front. 2024, 11, 6327–6332. [Google Scholar] [CrossRef]
  27. Zhou, X.; Bai, X.; Shang, F.; Zhang, H.-Y.; Wang, L.-H.; Xu, X.; Liu, Y. Supramolecular assembly activated single-molecule phosphorescence resonance energy transfer for near-infrared targeted cell imaging. Nat. Commun. 2024, 15, 4787. [Google Scholar] [CrossRef] [PubMed]
  28. Li, F.-F.; Huo, M.; Kong, J.; Liu, Y. Cucurbituril-Confined Tetracation Supramolecular 2D Organic Framework for Dual-Emission TS-FRET. Adv. Opt. Mater. 2024, 12, 2400453. [Google Scholar] [CrossRef]
  29. Niu, Q.; Ye, Y.; Su, L.; He, X.; Li, Z.; Liu, Y. Ultrastrong room-temperature phosphorescence in cucurbit[8]uril-mediated crystalline supramolecules for ratiometric detection of phenethylamine. Sci. China Chem. 2025, 68, 2671–2679. [Google Scholar] [CrossRef]
  30. Lin, F.; Yu, S.-B.; Liu, Y.-Y.; Liu, C.-Z.; Lu, S.; Cao, J.; Qi, Q.-Y.; Zhou, W.; Li, X.; Liu, Y.; et al. Porous Polymers as Universal Reversal Agents for Heparin Anticoagulants through an Inclusion-Sequestration Mechanism. Adv. Mater. 2022, 34, 2200549. [Google Scholar] [CrossRef]
  31. Yu, S.-B.; Lin, F.; Tian, J.; Yu, J.; Zhang, D.-W.; Li, Z.-T. Water-soluble and dispersible porous organic polymers: Preparation, functions and applications. Chem. Soc. Rev. 2022, 51, 434–449. [Google Scholar] [CrossRef] [PubMed]
  32. He, J.; Pu, Z.-H.; Dai, X.; Bai, Q.-H.; Tao, Z.; Xiao, X. Cucurbit[8]uril-modulated discrete pyrene dimers fluorescent sensor for nitroaromatic compounds and latent fingerprints visualization. Sens. Actuator B-Chem. 2024, 418, 136312. [Google Scholar] [CrossRef]
  33. Sun, Y.; Chen, Q.; Pan, D.; Xu, X.; Bai, Q.-H.; Wang, C.-H.; Zeng, X.; Xiao, X. Supramolecular phosphorescent assemblies based on cucurbit[8]uril and bromophenylpyridine derivatives for dazomet recognition. Spectrochim. Acta A. 2025, 330, 125695. [Google Scholar] [CrossRef]
  34. Zhang, H.; Yang, D.; Guan, M.; Li, Q.; Xu, J.; Cai, M.; Xu, J.; Liu, Q. Study of supramolecular organic frameworks for purification of hydrogen through molecular dynamics simulations. Sep. Purif. Technol. 2024, 335, 126106. [Google Scholar] [CrossRef]
  35. Zhang, D.-S.; Zhang, W.-Z.; Ding, X.-Y.; Xiao, J.; Wang, W.; Zhang, Y.-K.; Zhang, R.-H.; Hu, H.; Zhang, Y.-Z.; Geng, L.; et al. A novel flake-shaped supramolecular building block based metal-organic framework: Structure analysis and selective dye adsorption properties. Polyhedron 2023, 244, 116595. [Google Scholar] [CrossRef]
  36. Gao, Z.-Z.; Xu, Y.-Y.; Wang, Z.-K.; Wang, H.; Zhang, D.-W.; Li, Z.-T. Porous [Ru(bpy)3]2+-Cored Metallosupramolecular Polymers: Preparation and Recyclable Photocatalysis for the Formation of Amides and 2 Diazo-2-phenylacetates. ACS Appl. Polym. Mater. 2020, 2, 4885–4892. [Google Scholar] [CrossRef]
  37. Gao, Z.-Z.; Shen, L.; Hu, Y.-L.; Sun, J.-F.; Wei, G.; Zhao, H. Supramolecular Crystal Networks Constructed from Cucurbit[8]uril with Two Naphthyl Groups. Molecules 2023, 28, 63. [Google Scholar] [CrossRef]
  38. Zhao, H.; Shen, F.-F.; Sun, J.-F.; Gao, Z.-Z. Cucurbit[8]uril-controlled[2+2]photodimerization of styrylpyridinium molecule. Inorg. Chem. Commun. 2022, 141, 109536. [Google Scholar] [CrossRef]
  39. Liu, Y.; Wang, Z.-K.; Gao, Z.-Z.; Zong, Y.; Sun, J.-D.; Zhou, W.; Wang, H.; Ma, D.; Li, Z.-T.; Zhang, D.W. Porous organic polymer overcomes the post-treatment phototoxicity of photodynamic agents and maintains their antitumor efficiency. Acta Biomater. 2022, 150, 254–264. [Google Scholar] [CrossRef]
  40. Gao, Z.-Z.; Wang, Z.-K.; Wei, L.; Yin, G.; Tian, J.; Liu, C.-Z.; Wang, H.; Zhang, D.-W.; Zhang, Y.-B.; Li, X.; et al. Water-Soluble 3D Covalent Organic Framework that Displays an Enhanced Enrichment Effect of Photosensitizers and Catalysts for the Reduction of Protons to H2. ACS Appl. Mater. Interfaces 2020, 12, 1404–1411. [Google Scholar] [CrossRef]
  41. Day, A.I.; Arnold, A.P.; Blanch, R.J.; Snushall, B.J. Controlling Factors in the Synthesis of Cucurbituril and Its Homologues. J. Org. Chem. 2001, 66, 8094–8100. [Google Scholar] [CrossRef] [PubMed]
  42. Kim, J.; Jung, I.-S.; Kim, S.-Y.; Lee, E.; Kang, J.-K.; Sakamoto, S.; Yamaguchi, K.; Kim, K. New Cucurbituril Homologues: Syntheses, Isolation, Characterization, and X-ray Crystal Structures of Cucurbit[n]uril (n = 5, 7, and 8). J. Am. Chem. Soc. 2000, 122, 540–541. [Google Scholar] [CrossRef]
  43. Tian, J.; Xu, Z.-Y.; Zhang, D.-W.; Wang, H.; Xie, S.-H.; Xu, D.-W.; Ren, Y.-H.; Wang, H.; Liu, Y.; Li, Z.-T. Supramolecular metalorganic frameworks that display high homogeneous and heterogeneous photocatalytic activity for H2 production. Nat. Commun. 2016, 7, 11580. [Google Scholar] [CrossRef]
  44. Zhang, Y.-C.; Wu, Z.-Y.; Wang, Z.-K.; Wang, H.; Zhang, D.-W.; Liu, Y.; Li, Z.-T. A woven supramolecular metal-organic framework comprising a ruthenium bis(terpyridine) complex and cucurbit[8]uril: Enhanced catalytic activity toward alcohol oxidation. ChemPlusChem 2020, 85, 1498–1503. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures for cucurbit[8]uril (Q[8]) and 1-methyl-4-(4-(1-methylpyridin-1-ium-4-yl)benzamido)pyridin-1-ium (PB2+).
Figure 1. Structures for cucurbit[8]uril (Q[8]) and 1-methyl-4-(4-(1-methylpyridin-1-ium-4-yl)benzamido)pyridin-1-ium (PB2+).
Organics 06 00026 g001
Figure 2. 1H NMR spectrum (400 MHz) of (a) PB2+ (0.5 mM), (b–g) in the presence of 0.2, 0.3, 0.4, 0.5, and 0.6 equiv. of Q[8] in D2O at 25 °C.
Figure 2. 1H NMR spectrum (400 MHz) of (a) PB2+ (0.5 mM), (b–g) in the presence of 0.2, 0.3, 0.4, 0.5, and 0.6 equiv. of Q[8] in D2O at 25 °C.
Organics 06 00026 g002
Figure 3. (a) Changes in the UV–vis absorption spectra of PB2+ (2.0 × 10–5 M) in aqueous solution upon gradual addition of Q[8] (0–1.5 equiv); (b) changes in absorbance at 298 nm with various molar ratios of [Q[8]]/[PB2+].
Figure 3. (a) Changes in the UV–vis absorption spectra of PB2+ (2.0 × 10–5 M) in aqueous solution upon gradual addition of Q[8] (0–1.5 equiv); (b) changes in absorbance at 298 nm with various molar ratios of [Q[8]]/[PB2+].
Organics 06 00026 g003
Figure 4. PB2+ (1.0 mM) titration of the Q[8] (0.1 mM) isothermal titration heat curve and nonlinear fitting result of molar ratios at 298 K. (a) Corrected Heat Rate; (b) Enthalpy and Fit.
Figure 4. PB2+ (1.0 mM) titration of the Q[8] (0.1 mM) isothermal titration heat curve and nonlinear fitting result of molar ratios at 298 K. (a) Corrected Heat Rate; (b) Enthalpy and Fit.
Organics 06 00026 g004
Figure 5. ESI–mass spectrum of the complex PB2+@Q[8] (2:1) in aqueous solution. Chemical formula: C86H86N38O184+. Exact mass: 1938.6982 (484.6746).
Figure 5. ESI–mass spectrum of the complex PB2+@Q[8] (2:1) in aqueous solution. Chemical formula: C86H86N38O184+. Exact mass: 1938.6982 (484.6746).
Organics 06 00026 g005
Figure 6. Theoretical calculation structure of self-assembly PB2+@Q[8] viewed from the (left) front and (right) side (To facilitate observation, the atoms of C, N, O, and H in Q8 are uniformly represented in red).
Figure 6. Theoretical calculation structure of self-assembly PB2+@Q[8] viewed from the (left) front and (right) side (To facilitate observation, the atoms of C, N, O, and H in Q8 are uniformly represented in red).
Organics 06 00026 g006
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

Wang, Z.; Yang, M.; Yang, W.; Gao, Z.; Zhao, H.; Wei, G.; Sun, J. A Study of the Inclusion Complex Formed Between Cucurbit[8]uril and N,4-Di(pyridinyl)benzamide Derivative. Organics 2025, 6, 26. https://doi.org/10.3390/org6020026

AMA Style

Wang Z, Yang M, Yang W, Gao Z, Zhao H, Wei G, Sun J. A Study of the Inclusion Complex Formed Between Cucurbit[8]uril and N,4-Di(pyridinyl)benzamide Derivative. Organics. 2025; 6(2):26. https://doi.org/10.3390/org6020026

Chicago/Turabian Style

Wang, Zhikang, Mingjie Yang, Weibo Yang, Zhongzheng Gao, Hui Zhao, Gang Wei, and Jifu Sun. 2025. "A Study of the Inclusion Complex Formed Between Cucurbit[8]uril and N,4-Di(pyridinyl)benzamide Derivative" Organics 6, no. 2: 26. https://doi.org/10.3390/org6020026

APA Style

Wang, Z., Yang, M., Yang, W., Gao, Z., Zhao, H., Wei, G., & Sun, J. (2025). A Study of the Inclusion Complex Formed Between Cucurbit[8]uril and N,4-Di(pyridinyl)benzamide Derivative. Organics, 6(2), 26. https://doi.org/10.3390/org6020026

Article Metrics

Back to TopTop