Next Article in Journal
One-Pot Solvent-Involved Synthesis of 5-O-Substituted 5H-Chromeno[2,3-b]pyridines
Next Article in Special Issue
Recent Advances in Supramolecular-Macrocycle-Based Nanomaterials in Cancer Treatment
Previous Article in Journal
Microwave-Assisted Cu-Catalyzed Diaryletherification for Facile Synthesis of Bioactive Prenylated Diresorcinols
Previous Article in Special Issue
Nano-Theranostics Constructed from Terpyridine-Modified Pillar [5]arene-Based Supramolecular Amphiphile and Its Application in Both Cell Imaging and Cancer Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Supramolecular Crystal Networks Constructed from Cucurbit[8]uril with Two Naphthyl Groups

1
College of Chemical and Biological Engineering, Shandong University of Science and Technology, 579 Qianwangang Road, Qingdao 266590, China
2
Commonwealth Scientific and Industrial Research Organisation (CSIRO), Mineral Resources, P.O. Box 218, Lindfield, NSW 2070, Australia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(1), 63; https://doi.org/10.3390/molecules28010063
Submission received: 1 December 2022 / Revised: 16 December 2022 / Accepted: 17 December 2022 / Published: 21 December 2022
(This article belongs to the Special Issue Supramolecular Cancer Nanotheranostics)

Abstract

:
Naphthyl groups are widely used as building blocks for the self-assembly of supramolecular crystal networks. Host–guest complexation of cucurbit[8]uril (Q[8]) with two guests NapA and Nap1 in both aqueous solution and solid state has been fully investigated. Experimental data indicated that double guests resided within the cavity of Q[8], generating highly stable homoternary complexes NapA2@Q[8] and Nap12@Q[8]. Meanwhile, the strong hydrogen-bonding and π···π interaction play critical roles in the formation of 1D supramolecular chain, as well as 2D and 3D networks in solid state.

Graphical Abstract

1. Introduction

Supramolecular self-assembly is a spontaneous process in which basic structural units form ordered structures driven by non-covalent interactions (including electrostatic interactions, van der Waals forces, hydrogen bond, π–π stacking, hydrophobic effects, etc.) [1,2,3,4,5,6]. Based on self-assembly, various supramolecular network structures have been widely developed. Reversible non-covalent bonds endow supramolecular networks with traditional polymer properties, as well as dynamic properties, which can achieve high response to stimuli and self-healing functions [7,8,9,10]. These supramolecular network systems have been constructed to synthesize different functional materials and widely used in drug carriers, nano containers, molecular devices, sensors, adsorption and separation materials, catalysts, and environmental pollutant treatment [11,12,13,14,15]. Although the supramolecular polymerization process has been deeply understood, it is still a huge challenge to design controllable supramolecular polymerization systems and provide supramolecular polymers with controllable structures [16].
The macrocyclic host cucurbit[n]urils (Q[n]s, CB[n]s, n = number of glycoluril units) are spherical macrocyclic compounds adopting a stable rigid structure and composed of n glycoluril units bridged by 2n methylene groups with a hydrophobic cavity [17,18,19,20]. As a kind of rigid macrocyclic host, cucurbit[n]urils have three main structural characteristics, namely, the neutral cavity, the negative electrostatic potential carbonyl portals and the positive electrostatic potential of the outer surface [21,22]. Based on the above three characteristics, the Q[n]s related host-guest chemistry, coordination chemistry and outer surface interaction chemistry have been widely used [23,24,25]. The cavity of the Q[n]s can selectively bind a part of or whole guest molecules to form host–guest complexes [26,27,28,29]. According to the Q[n]’s cavity size, Q[8] exhibits unique molecular recognition characteristics enabling dimerization of specific guest molecules inside the cavity of Q[8] in a controlled manner. The larger cavity of Q[8] can bind two aromatic group, such as phenyl, naphthyl, indole, quinoline, and larger aromatic rings, thus allowing for the formation of homoternary inclusion complexes [30,31,32]. Therefore, Q[8] has been vigorously studied as basic units for the construction of supramolecular networks [33,34,35]. For example, Zhang’s team constructed a variety of supramolecular hyperbranched networks by self-assembly of dendrimers and Q[8] [36,37]. Liu’s team constructed a variety of two-dimensional supramolecular network structures based on triphenylamine and Q[8] [38,39], which have good effects in many fields such as cell imaging, near-infrared lysosome targeted imaging. Li’s team prepared a tetrahedral molecule which was used to co-assemble with Q[8] to afford a new water soluble 3D diamondoid system [40]. Moreover, the hexa-armed [Ru(bpy)3]2+-based derivatives and Q[8] were used to afford another 3D cubelike system, which could be used as heterogeneous catalyst for hydrogen production and organic reaction [41]. In previous work, we used the antiparallel encapsulation of styrene pyridine dimer in Q[8] to construct a variety of supramolecular polymer systems [42,43]. For example, we generated an irreversible covalent component for the construction of the first highly watersoluble 3D supramolecular-covalent organic framework, which was used to highly promote the electron transfer of protons to H2 when loaded on POM catalysts and Ru2+-complex photosensitizers [42]. Having said that, we believe that Q[n]s are ideal as basic building blocks for the construction of Q[n]-based networks.
Herein, we introduced two guests containing naphthyl groups, 4-(4-carboxyphenyl)-1-(naphthalen-2-ylmethyl)pyridin-1-ium bromide (NapA) and 1-(naphthalen-2-ylmethyl)-[4,4′-bipyridin]-1-ium bromide (Nap1) for the formation of supramolecular networks. The Q[8]-induced supramolecular networks were investigated by 1H nuclear magnetic resonance (1H NMR) spectrum, UV-vis absorption spectrum, isothermal titration calorimetry (ITC), dynamic light scattering (DLS), and single-crystal X-ray crystallography. Experimental data indicated that double guests resided within the cavity of Q[8] in both aqueous solution and solid state, generating highly stable homoternary complexes NapA2@Q[8] and Nap12@Q[8]. The strong hydrogen-bonding and π···π interaction played critical roles in the formation of supramolecular networks.

2. Results

2.1. Molecular Binding Behavior and Thermodynamic between Cucurbit[8]uril and NapA or Nap1

It has been established that Q[8] encapsulation for the dimers of naphthyl unit in water can remarkably promote the π–π interaction. Thus, two guests containing naphthyl unit (NapA and Nap1) were designed and synthesized (Scheme 1). Two guests, NapA and Nap1, were prepared in a one-step sequence. Compound NapA was prepared from the reaction of 4-(4-pyridyl)benzoic acid and 2-bromomethyl naphthalene in MeCN in 90% yield and characterized using 1H NMR, 13C NMR, 1H-1H COSY spectra and high-resolution mass spectra (HR-MS). 4,4′-bipyridine and 2-bromomethyl naphthalene were used to synthesize Nap1 in acetone in 85% yield. We then studied their coassembly with Q[8] in water for the formation of two new homoternary supramolecular complexes.
The naphthyl units of two guests can combine to form supramolecular dimers under the action of Q[8]-driven self-assembly. Two examples of homoternary supramolecular complexes were constructed under the coordination of supramolecular interactions such as Q[n]s’ outer surface interaction, hydrogen bond interaction, strong π–π interaction, host-guest interaction, etc. Simply mixing Q[8] and Nap1 or NapA in aqueous solution resulted in the formation of the supramolecular networks depending on the corresponding host–guest interaction between the two species of molecules, which was further characterized by 1H NMR spectroscopy, X-ray single crystal diffraction, and dynamic light scattering analysis (DLS).
The 1H NMR spectroscopy measurements indicated that both Nap1 and NapA form host-guest inclusion complexes with Q[8] host. In the presence of small amount of the Q[8] hosts (Figure 1), the signals of both free and complexed guests were simultaneously observed and were very broad, indicating slow exchange of free and complexed guests on the NMR time scale. On the one hand, the protons of bipyridine (H1-H4) and one of the CH2 protons (H5) of Nap1 moved downfield slightly, which indicated that they were located outside the cavity. On the other hand, at a 2:1 ratio of Nap1 to Q[8], the naphthyl peaks (H6-H12) were completely shifted upfield. These observations suggested that the naphthyl moiety of the Nap1 guest was encapsulated into the cavity of the Q[8] host.
The corresponding 1H NMR titration spectra of the Q[8] with NapA were showed in Figure S1. Due to the low solubility of NapA, the NMR signals became wide peak after adding Q[8]. In the presence of 0.5 equiv of Q[8], the signals corresponding to the naphthyl (H6-H12) protons of the NapA shifted upfield, while those corresponding to the pyridinyl (H3 and H4), phenyl (H1 and H2) protons did not move significantly. This clearly indicated that the naphthyl protons of the guest NapA were buried inside the hydrophobic cavity of Q[8], while the 4-(4-pyridyl)benzoic acid moiety resided outside of the Q[8] portals, forming homoternary supramolecular complexes with 1:2 host-guest binding ratio NapA2@Q[8]. By comparing the NMR titration data of the two self-assemblies, we found that the naphthyl group of the guest molecule was encapsulated in the hydrophobic cavity of Q[8].
Isothermal titration calorimetry (ITC) was also employed to afford the thermodynamic parameters of Q[8] with both Nap1 and NapA. ITC experimental data further confirmed that the binding stoichiometry of Q[8] to both Nap1 and NapA is 1:2 (Figure 2). From the ΔH and TΔS values in Table S1, it was clear that the formation of both homoternary complexes were enthalpically driven. The observed negative enthalpy change (ΔH = −36.17 ± 2.59 kJ·mol−1 for NapA2@Q[8]; ΔH = −22.71 ± 1.37 kJ·mol−1 for Nap12@Q[8]) were probably due to the cooperativity of above mentioned four kinds of weak interactions. On the basis of the corresponding experimental results, we also obtained the association constants of Ka = (2.14 ± 0.62) × 1010 M−2 and (1.48 ± 0.45) × 1010 M−2 for Q[8] with NapA and Nap1, respectively, which was much larger than that of Q[8] with tripeptides reported by Urbach [44] and closer to Q[8] with tripeptides reported by us [30]. Such a high binding constant suggested the relatively strong host–guest interaction between Q[8] and NapA or Nap1, indicating the construction of stable homoternary complexes NapA2@Q[8] and Nap12@Q[8] in aqueous solution.
To better understand the host-guest interaction between Q[8] and both Nap1 and NapA in aqueous solution, we also carried out UV-vis titration experiments. According to the UV-vis absorption spectroscopic results, as shown in Figure 3, the compound Nap1 and NapA displayed an absorption band at 224 nm belonging to naphthyl unit, which decreased markedly in its intensity upon addition of Q[8], due to the strong interaction between Q[8] and naphthyl moiety of guest molecules Nap1 and NapA. When 0.5 equivalent Q[8] was added, the UV−vis absorption spectra intensity did not change significantly. Their Job plots (based on the continuous variation method) clearly showed that UV−vis spectra data of both Nap1 and NapA fitted well to 1:2 stoichiometry of the host-guest inclusion complexes (Figure 3, inset). The binding stoichiometry was also determined by the Job plot analysis at a fixed total concentration of host and guest molecules. The absorption intensity changes (ΔA) were plotted against the molar fraction of Nap1 and NapA to give a peak at a molar fraction of 0.66, indicating a 1:2 stoichiometry for the Q[8]:Nap1 or Q[8]:NapA inclusion complex (Figures S2 and S3), which was consistent with the NMR titration results.
DLS experiments in dilute solutions can be used to monitor the formation of supramolecular networks. DLS results revealed that NapA, Nap1 and Q[8] formed nanoscaled assemblies in water. As can be found, the hydrodynamic diameter (DH) of NapA monomer (1.0 mM) was determined to be 0.6 nm. In Figure 4a, mixing Q[8] and NapA (1:2, [NapA] = 0.1 mM) observed one hydrodynamic diameter distribution centered at 140 nm. It showed that aggregates of this size were formed in the aqueous solution. The DH value decreased with the dilution of the solution. However, even at [NapA] = 25 μM, DH was still as high as 43 nm. These observations supported that NapA and Q[8] coassembled into the nanoscaled supramolecular networks in water. At the same concentration, the hydrated particle size of assembly NapA2@Q[8] is obviously higher than that of assembly Nap12@Q[8], which indicates that the polymerization ability of assembly NapA2@Q[8] is higher than that of assembly Nap12@Q[8] in the aqueous phase. DLS experiment showed that the self-assembly had a certain stability in the solution even at low concentration. The result was consistent with the results of crystal structure description. The DLS experiment of Nap12@Q[8] (Figure 4b) was similar to that of NapA2@Q[8].

2.2. Single-Crystal X-ray Crystallography

X-ray structure analysis provided unequivocal proof of the formation of homoternary complexes between Q[8] and both Nap1 and NapA. Crystal of NapA2@Q[8] was grown by slow evaporation of a solution containing the host Q[8] and the guest NapA under 3.0 M aqueous hydrochloric acid solution. X-ray structural analysis previously established that the NapA2@Q[8] crystallized in the orthorhombic crystal system, space group Pbca. As can be seen in Figure 5a, the naphthyl moiety of the NapA guest located inside the cavity of the Q[8] host, which was in agreement with what we had observed in the aqueous solution by 1H NMR spectroscopy. Furthermore, the π···π interactions between two encapsulated NapA molecules played a critical role in the formation of this host-guest inclusion complex. Obviously, the van der Waals contacted between the naphthyl groups and the inner wall of the Q[8] cavity, and strong hydrogen-bonding, such as C(32)-H···O(6) 2.701 Å (between carbonyl oxygen of host and H on benzene ring of guest), contributed to the formation of the inclusion complex NapA2@Q[8]. As shown in Figure 5b, two adjacent complexes formed a two dimensional assembled host–guest supramolecular network via hydrogen-bonding interactions between the portal carbonyl oxygen atom O6 of Q[8] and between the carboxyl oxygen atom O9, O10 of NapA. It is should be noted that the hydrogen bonds between the carboxyl oxygen atom of the guest NapA and the hydrogen atom on methylene of Q[8] (Figure 5b): C(14)-H···O(9) 2.571 Å, C(16)-H···O(10) 2.430 Å, and between the carbonyl oxygens at the portals of the Q[8] host and the hydrogen atom on the methylene of the adjacent Q[8] (Figure S4): C(11)-H···O(8) 2.696 Å may be largely responsible for the construction of a one-dimensional supramolecular chain (Figure S5) and supramolecular networks (Figure 5c).
Single crystals of complex Nap12@Q[8] were fortunately obtained from hydrochloride acid solution by slow evaporation in the presence of CdCl2. The complex crystallized in the triclinic crystal system, space group P-1, the compounds consists of one Q[8] host and two Nap1 guests. As shown in Figure 6a, X-ray structural analysis revealed that the naphthyl moiety of the guest Nap1 adopted reverse parallel in the Q[8] cavity, the bipyridyl moiety remained outside of its portal, which could be attributed to π···π interactions in the cavity of Q[8]. Outside of the inclusion complexes, neighboring Nap1 molecules contacted with each other through not only π···π interaction, but also C-H···π interactions (Figure 6b), which may serve to stabilize this structure for building a one-dimensional supramolecular chain of the complexes. In the latter, the encapsulated guests were electron donor and acceptor pair, and the major driving force for the ternary complex formation appears to be strong charge-transfer interaction between the guests [45]. Furthermore, the strong hydrogen-bonding played a critical role in the formation of this host-guest inclusion complex, e.g., between H on methylene of Nap1 and the carbonyl oxygens at the portals of the Q[8] host C(59)–H···O(6) 2.691 Å, between H on pyridine ring and the carbonyl oxygens at the portals of the adjacent Q[8] host C(70)–H···O(10) 2.660 Å. Meanwhile, the hydrogen-bonding between two adjacent Q[8]: C(58)–H···O(16) 2.311 Å (Figure S6) and between [CdCl4]2− and Nap1: C(23)–H···Cl(9) 2.692 Å, between [CdCl4]2− and Q[8] C(53)–H···Cl(11) 2.836 Å (Figure S7) cannot be ignored in the construction of one-dimensional supramolecular chains (Figure 6c).

3. Materials and Methods

All reagents were obtained from commercial suppliers and used without further purification unless otherwise noted. All reactions were carried out under a dry nitrogen atmosphere. All solvents were dried before use following standard procedures. 1H and 13C NMR spectra were recorded on 400 MHz spectrometers in the indicated solvents at room temperature (298 K). Dynamic light scattering (DLS) measurement was conducted on Malvern Zetasizer Nano ZS90 using a monochromatic coherent He–Ne laser (633 nm) as the light source and a detector that detected the scattered light at an angle of 90°. An isothermal titration calorimetry (ITC) experiment was carried out using a MicroCal PEAQ-ITC (Malven Panalytical, Worcestershire, UK) instrument. Association constants and associated thermodynamic parameters were obtained through computer simulations (curve fitting) using Micro-Cal ITC analyze software (MicroCal Origin 4.1). UV–vis spectra were detected on a PerkinElmer 750s instrument from 200 to 800 nm at the scan rate of 3 nm/internal.
Single crystals of NapA2@Q[8] and Nap12@Q[8] were grown from hydrochloride acid solution by slow evaporation. The crystal culture conditions showed that the self-assembly NapA2@Q[8] and Nap12@Q[8] had good stability in strong acid solution. Diffraction data of both complexes were collected at 273(2) K with a Bruker SMART Apex-II CCD diffractometer using graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å). Empirical absorption corrections were performed by using the multi-scan program SADABS. Structural solution and full-matrix least-squares refinement based on F2 were performed with the SHELXS-97 and SHELXL-97 program packages, respectively. Non-hydrogen atoms were treated anisotropically in all cases. All hydrogen atoms were introduced as riding atoms with an isotropic displacement parameter equal to 1.2 times that of the parent atom. Hydrogen atoms were given for all isolated water molecules.
CCDC 2223335 (Nap12@Q[8]) and 2223329 (NapA2@Q[8]) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif, which accessed on 3 December 2022.

4. Conclusions

In summary, we investigated the host–guest complexation of Q[8] with two enantiomers, NapA and Nap1, in both aqueous solution and solid state by using NMR, UV-vis spectrum, ITC, DLS, and single-crystal X-ray crystallography. Driven by the cooperativity of electrostatic interactions, multiple C–H···π interactions, and hydrogen-bonding, both NapA and Nap1 can be encapsulated into the cavity of Q[8] to form stable homoternary complexes NapA2@Q[8] and Nap12@Q [8]. Structure analysis shows that hydrogen-bonding interactions and π···π interactions play a critical role not only in the formation of 1D extended chains, but also in the construction of 2D and 3D networks. This study shows that Q[8] host and molecules containing naphthalene units can be used as building units to build a diverse supramolecular network structure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28010063/s1, Figure S1: 1H NMR spectrum (400 MHz) of the mixtures of NapA (1.0 mM) and Q[8] (1:0.5) in D2O at 25 °C; Figures S2 and S3: Job’s plot obtained from the absorption spectra of the mixtures of NapA or Nap1 and Q[8] ([NapA or Nap1] + Q[8] = 50 μM) in water at 25 °C; Figures S4 and S5: Details of NapA2@Q[8] crystal structure; Figures S6 and S7: Details of Nap12@Q[8] crystal structure; Figures S8–S11: Characterization of NapA; Figures S12–S15: Characterization of Nap1; Table S1. ITC measurements of the thermodynamics of NapA2@Q[8] and Nap12@Q[8] interactions in aqueous solution at 298.15 K.

Author Contributions

Conceptualization, Z.-Z.G., Y.-L.H. and L.S.; methodology, Z.-Z.G.; validation, J.-F.S.; formal analysis, Z.-Z.G.; writing—original draft preparation, H.Z.; writing—review and editing, G.W.; supervision, J.-F.S.; funding acquisition, Z.-Z.G. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of Shandong Province (ZR2021QB178, ZR2021QB197) and Shandong University of Science and Technology Research Fund (Grant numbers: skr20-3-023, skr20-3-040).

Institutional Review Board Statement

No applicable.

Informed Consent Statement

No applicable.

Data Availability Statement

No applicable.

Acknowledgments

We thank Pei-Hui Shan of Guizhou University for his help in NMR testing. We also thank Key Laboratory of Macrocyclic and Supramolecular Chemistry of Guizhou Province, Guizhou University for characterization.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Busseron, E.; Ruff, Y.; Moulin, E.; Giuseppone, N. Supramolecular self-assemblies as functional nanomaterials. Nanoscale 2013, 5, 7098–7140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wei, P.; Yan, X.; Huang, F. Supramolecular polymers constructed by orthogonal self-assembly based on host–guest and metal–ligand interactions. Chem. Soc. Rev. 2015, 44, 815–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ariga, K.; Nishikawa, M.; Mori, T.; Takeya, J.; Shrestha, L.K.; Hill, J.P. Self-assembly as a key player for materials nanoarchitectonics. Sci. Technol. Adv. Mater. 2019, 20, 51–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Cao, X.; Gao, A.; Hou, J.-T.; Yi, T. Fluorescent supramolecular self-assembly gels and their application as sensors: A review. Coord. Chem. Rev. 2021, 434, 213792. [Google Scholar] [CrossRef]
  5. Escobar, L.; Ballester, P. Molecular recognition in water using macrocyclic synthetic receptors. Chem. Rev. 2021, 121, 2445–2514. [Google Scholar] [CrossRef]
  6. Li, X.-Z.; Tian, C.-B.; Sun, Q.-F. Coordination-Directed Self-Assembly of Functional Polynuclear Lanthanide Supramolecular Architectures. Chem. Rev. 2022, 122, 6374–6458. [Google Scholar] [CrossRef]
  7. Chen, X.-X.; Zhong, Q.-Y.; Wang, S.-J.; Wu, Y.-S.; Tan, J.-D.; Lei, H.-X.; Huang, S.-Y.; Zhang, Y.-F. Progress in dynamic covalent polymers. Acta Polym. Sin. 2019, 50, 469–484. [Google Scholar]
  8. Cui, L.; Zhao, M.-H.; Zhang, C. Recent advance in applications of host-guest interaction in biochemical analysis. Chin. J. Anal. Chem. 2020, 48, 817–826. [Google Scholar]
  9. Lee, W.-J.; Oh, H.-G.; Cha, S.-H. A brief review of self-healing polyurethane based on dynamic chemistry. Macromol. Res. 2021, 29, 649–664. [Google Scholar] [CrossRef]
  10. Yang, J.; Chen, Y.; Zhao, L.; Zhang, J.; Luo, H. Constructions and Properties of Physically Cross-Linked Hydrogels Based on Natural Polymers. Polym. Rev. 2022, 1–39. [Google Scholar] [CrossRef]
  11. Xia, D.; Wang, P.; Ji, X.; Khashab, N.M.; Sessler, J.L.; Huang, F. Functional supramolecular polymeric networks: The marriage of covalent polymers and macrocycle-based host–guest interactions. Chem. Rev. 2020, 120, 6070–6123. [Google Scholar] [CrossRef] [PubMed]
  12. 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, 2022, 2200549. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Y.-C.; Zeng, P.-Y.; Ma, Z.-Q.; Xu, Z.-Y.; Wang, Z.-K.; Guo, B.; Yang, F.; Li, Z.-T. A pH-responsive complex based on supramolecular organic framework for drug-resistant breast cancer therapy. Drug Delivery 2022, 29, 1–9. [Google Scholar] [CrossRef]
  14. Guo, Y.; Liu, Y.; Zhao, X.; Zhao, J.; Wang, Y.; Zhang, X.; Guo, Z.; Yan, X. Synergistic Covalent-and-Supramolecular Polymers with an Interwoven Topology. ACS Appl. Mater. Interfaces 2022. [Google Scholar] [CrossRef] [PubMed]
  15. Shan, P.; Hu, J.; 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]
  16. Zhang, X. Supramolecular Polymer Chemistry: Past, Present, and Future. Chinese J. Polym. Sci. 2022, 40, 541–542. [Google Scholar] [CrossRef]
  17. Lee, J.W.; Samal, S.; Selvapalam, N.; Kim, H.-J.; Kim, K. Cucurbituril homologues and derivatives: New opportunities in supramolecular chemistry. Acc. Chem. Res. 2003, 36, 621–630. [Google Scholar] [CrossRef]
  18. 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]
  19. 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]
  20. Chen, K.; Hua, Z.-Y.; Zhao, J.-L.; Redshaw, C.; Tao, Z. Construction of cucurbit[n]uril-based supramolecular frameworks via host–guest inclusion and functional properties thereof. Inorg. Chem. Front. 2022, 9, 2753–2809. [Google Scholar] [CrossRef]
  21. Gao, R.H.; Chen, L.X.; Chen, K.; Tao, Z.; Xiao, X. Development of hydroxylated cucurbit[n]urils, their derivatives and potential applications. Coord. Chem. Rev. 2017, 348, 1–24. [Google Scholar] [CrossRef]
  22. Yang, D.; Liu, M.; Xiao, X.; Tao, Z.; Redshaw, C. Polymeric self-assembled cucurbit[n]urils: Synthesis, structures and applications. Coord. Chem. Rev. 2021, 434, 213733. [Google Scholar] [CrossRef]
  23. Liu, M.; Chen, L.; Shan, P.; Lian, C.; Zhang, Z.; Zhang, Y.; Tao, Z.; Xiao, X. Pyridine detection using supramolecular organic frameworks incorporating cucurbit[10]uril. ACS Appl. Mater. Interfaces 2021, 13, 7434–7442. [Google Scholar] [CrossRef]
  24. Huang, Y.; Gao, R.H.; Liu, M.; Chen, L.X.; Ni, X.L.; Xiao, X.; Cong, H.; Zhu, Q.J.; Chen, K.; Tao, Z. Cucurbit[n]uril-Based Supramolecular Frameworks Assembled through Outer-Surface Interactions. Angew. Chem. 2021, 133, 15294–15319. [Google Scholar] [CrossRef]
  25. Zhang, W.; Luo, Y.; Ni, X.-L.; Tao, Z.; Xiao, X. Two-step, Sequential, Efficient, Artificial Light-harvesting Systems Based on Twisted Cucurbit[13]uril for Manufacturing White Light Emission Materials. Chem. Eng. J. 2022, 2022, 136954. [Google Scholar] [CrossRef]
  26. Luo, Y.; Zhang, W.; Yang, M.-X.; Feng, X.-H.; Redshaw, C.; Li, Q.; Tao, Z.; Xiao, X. A Twisted Cucurbit[14]uril-Based Fluorescent Supramolecular Polymer Mediated by Metal Ion. Macromolecules 2022, 55, 1642–1646. [Google Scholar] [CrossRef]
  27. Xu, D.-A.; Zhou, Q.-Y.; Dai, X.; Ma, X.-K.; Zhang, Y.-M.; Xu, X.; Liu, Y. Cucurbit[8]uril-mediated phosphorescent supramolecular foldamer for antibiotics sensing in water and cells. Chin. Chem. Lett. 2022, 33, 851–854. [Google Scholar] [CrossRef]
  28. Peng, M.; Luo, Y.; Rao, Y.; Song, J.; Ni, X.-L. Cucurbit[7]uril-Encapsulation-Controlled Supramolecular Photoproduct and Radical Fluorescence Emission. Chem.-Eur. J. 2022, 28, e202202056. [Google Scholar] [CrossRef]
  29. Liu, C.; Xia, Y.; Tao, Z.; Ni, X.-L. Host-guest interaction tailored cucurbit[6]uril-based supramolecular organic frameworks (SOFs) for drug delivery. Chin. Chem. Lett. 2022, 33, 1529–1532. [Google Scholar] [CrossRef]
  30. 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] [Green Version]
  31. Yang, B.; Yu, S.-B.; Wang, H.; Zhang, D.-W.; Li, Z.-T. 2: 2 Complexes from Diphenylpyridiniums and Cucurbit[8]uril: Encapsulation-Promoted Dimerization of Electrostatically Repulsing Pyridiniums. Chem. Asian J. 2018, 13, 1312–1317. [Google Scholar] [CrossRef] [PubMed]
  32. 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]
  33. 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. Ed. 2020, 59, 9928–9933. [Google Scholar] [CrossRef] [PubMed]
  34. Li, Y.; Li, Q.; Miao, X.; Qin, C.; Chu, D.; Cao, L. Adaptive Chirality of an Achiral Cucurbit[8]uril-Based Supramolecular Organic Framework for Chirality Induction in Water. Angew. Chem. 2021, 133, 6818–6825. [Google Scholar] [CrossRef]
  35. 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]
  36. Yang, H.; Bai, Y.; Yu, B.; Wang, Z.; Zhang, X. Supramolecular polymers bearing disulfide bonds. Polym. Chem. 2014, 5, 6439–6443. [Google Scholar] [CrossRef]
  37. Liu, Y.; Huang, Z.; Liu, K.; Kelgtermans, H.; Dehaen, W.; Wang, Z.; Zhang, X. Porphyrin-containing hyperbranched supramolecular polymers: Enhancing 1O2-generation efficiency by supramolecular polymerization. Polym. Chem. 2014, 5, 53–56. [Google Scholar] [CrossRef]
  38. Dai, X.Y.; Huo, M.; Dong, X.; Hu, Y.Y.; Liu, Y. Noncovalent Polymerization-Activated Ultrastrong Near-Infrared Room-Temperature Phosphorescence Energy Transfer Assembly in Aqueous Solution. Adv. Mater. 2022, 34, 2203534. [Google Scholar] [CrossRef]
  39. 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]
  40. Tian, J.; Zhou, T.-Y.; Zhang, S.-C.; Aloni, S.; Altoe, M.V.; Xie, S.-H.; Wang, H.; Zhang, D.-W.; Zhao, X.; Liu, Y. Three-dimensional periodic supramolecular organic framework ion sponge in water and microcrystals. Nat. Commun. 2014, 5, 5574. [Google Scholar] [CrossRef] [Green Version]
  41. 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 metal-organic frameworks that display high homogeneous and heterogeneous photocatalytic activity for H2 production. Nat. Commun. 2016, 7, 11580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. 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 2019, 12, 1404–1411. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Smith, L.C.; Leach, D.G.; Blaylock, B.E.; Ali, O.A.; Urbach, A.R. Sequence-specific, nanomolar peptide binding via cucurbit[8]uril-induced folding and inclusion of neighboring side chains. J. Am. Chem. Soc. 2015, 137, 3663–3669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ko, Y.H.; Kim, E.; Hwang, I.; Kim, K. Supramolecular assemblies built with host-stabilized charge-transfer interactions. Chem. Commun. 2007, 1305–1315. [Google Scholar] [CrossRef]
Scheme 1. (a) The synthesis of NapA and Nap1 and (b) the structure of host Q[8].
Scheme 1. (a) The synthesis of NapA and Nap1 and (b) the structure of host Q[8].
Molecules 28 00063 sch001
Figure 1. 1H NMR spectrum (400 MHz) of the mixtures of Nap1 (1.0 mM) and Q[8] (0 to 0.6 equiv) in D2O at 25 °C.
Figure 1. 1H NMR spectrum (400 MHz) of the mixtures of Nap1 (1.0 mM) and Q[8] (0 to 0.6 equiv) in D2O at 25 °C.
Molecules 28 00063 g001
Figure 2. (a,b) Nap1 (1.0 mM) and (c,d) NapA (1.0 mM) titration of the Q[8] (0.1 mM) isothermal titration heat curve and nonlinear fitting result of molar ratios at 25 °C.
Figure 2. (a,b) Nap1 (1.0 mM) and (c,d) NapA (1.0 mM) titration of the Q[8] (0.1 mM) isothermal titration heat curve and nonlinear fitting result of molar ratios at 25 °C.
Molecules 28 00063 g002aMolecules 28 00063 g002b
Figure 3. UV-vis spectra of NapA and Nap1 (10 μM) with the addition of Q[8] (0–2.0 equiv.) in water at 25 °C (Inset: Absorbance at 224 nm vs. [Q[8]]/[NapA], and [Q[8]]/[Nap1]).
Figure 3. UV-vis spectra of NapA and Nap1 (10 μM) with the addition of Q[8] (0–2.0 equiv.) in water at 25 °C (Inset: Absorbance at 224 nm vs. [Q[8]]/[NapA], and [Q[8]]/[Nap1]).
Molecules 28 00063 g003
Figure 4. DLS profile of (a) NapA2@Q[8] and (b) Nap12@Q[8] in water at 25 °C. The concentration represents that of NapA or Nap1 of homoternary complexes.
Figure 4. DLS profile of (a) NapA2@Q[8] and (b) Nap12@Q[8] in water at 25 °C. The concentration represents that of NapA or Nap1 of homoternary complexes.
Molecules 28 00063 g004
Figure 5. (a) X-ray crystal structure of the homoternary complex NapA2@Q[8]. Solvate water molecules were omitted for clarity. (b) The strong hydrogen-bonding between NapA and Q[8] molecules. (c) Two-dimensional supramolecular network of complexes.
Figure 5. (a) X-ray crystal structure of the homoternary complex NapA2@Q[8]. Solvate water molecules were omitted for clarity. (b) The strong hydrogen-bonding between NapA and Q[8] molecules. (c) Two-dimensional supramolecular network of complexes.
Molecules 28 00063 g005
Figure 6. (a) X-ray crystal structure of the homoternary complex Nap12@Q[8]. Solvate water molecules and [CdCl4]2− anions were omitted for clarity. (b) The C–H···π interaction and π···π interaction between two neighboring Nap1 molecules. (c) Crystal structure of a one-dimensional supramolecular chain constructed of the Nap1 and Q[8].
Figure 6. (a) X-ray crystal structure of the homoternary complex Nap12@Q[8]. Solvate water molecules and [CdCl4]2− anions were omitted for clarity. (b) The C–H···π interaction and π···π interaction between two neighboring Nap1 molecules. (c) Crystal structure of a one-dimensional supramolecular chain constructed of the Nap1 and Q[8].
Molecules 28 00063 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

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. https://doi.org/10.3390/molecules28010063

AMA Style

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(1):63. https://doi.org/10.3390/molecules28010063

Chicago/Turabian Style

Gao, Zhong-Zheng, Lei Shen, Yu-Lu Hu, Ji-Fu Sun, Gang Wei, and Hui Zhao. 2023. "Supramolecular Crystal Networks Constructed from Cucurbit[8]uril with Two Naphthyl Groups" Molecules 28, no. 1: 63. https://doi.org/10.3390/molecules28010063

Article Metrics

Back to TopTop