Next Article in Journal
The Network Basis for the Structural Thermostability and the Functional Thermoactivity of Aldolase B
Next Article in Special Issue
Efficient Removal of Hexavalent Chromium (Cr(VI)) from Wastewater Using Amide-Modified Biochar
Previous Article in Journal
Fatty Acid and Multi-Isotopic Analysis (C, H, N, O) as a Tool to Differentiate and Valorise the Djebel Lamb from the Mountainous Region of Tunisia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Benzothiadiazole-Based Self-Assembled Cage for Cadmium Detection

Department of Biology and Chemistry, Hunan University of Science and Engineering, Yongzhou 415199, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(4), 1841; https://doi.org/10.3390/molecules28041841
Submission received: 30 January 2023 / Revised: 9 February 2023 / Accepted: 12 February 2023 / Published: 15 February 2023

Abstract

:
A turn-on fluorescent probe, cage 1, was efficiently self-assembled by condensing 4,4′-(benzothiadiazole-4,7-diyl)dibenzaldehyde and TREN in chloroform. The formation of cage 1 was characterized and confirmed by NMR spectroscopy, mass spectrometry, and theoretical calculations. The yield of cage 1 could be controlled by tuning the reaction conditions, such as the precursor concentration. Interestingly, the addition of 10 equiv of Cd2+ relative to cage 1 could increase the fluorescence almost seven-fold. 1H NMR and fluorescence experiments indicating fluorescence enhancement may be caused by the decomposition of cage 1. Such a high selectivity toward Cd2+ implies that the cage could potentially be employed in cadmium detection.

1. Introduction

Molecular cages, as one of the supramolecular architectures, provide a synthetic three-dimensional cavity for binding guests, and offer a window for mimicking biological processes in nature [1,2,3,4,5,6,7,8]. Since the first synthesis of cryptands and cavitands in 1969by Jean-Marie Lehn and co-workers [9], various synthetic molecular cages have been reported and have found a variety of potential applications in gas separation [10,11,12,13,14], catalysis [15,16,17,18,19], molecular sensing, and other fields [20,21,22,23,24,25]. Among them, the synthesis of molecular cages by the self-assembly method has attracted extensive attention in recent years due to its simple synthesis and high yield. In 1988, MacDowell and Nelson [26] reported the condensation of tris(2-aminoethyl)amine (TREN) and aldehydes to a make TREN-based [2 + 3] cage in a yield of about 50%. Subsequently, reversible covalent linkages, such as the condensation of amine and aldehydes, have been proved to be a good choice for the high-yield synthesis of cage compounds. Dynamic bonds are formed in a reversible forming/cleaving manner, allowing the self-assembled products to search for their thermodynamic minimum. Recently, Li and coworkers [27] reported the controllable self-assembly of pills and cages via imine condensation for silver cation detection. Later, their group [28] also reported the in situ detection of silver cations in water by a capsule-shaped cage formed by condensing a trialdehyde precursor and TREN. Khashab’s group [29] realized the separation of a mixture of benzene and cyclohexane by thienothiophene cages which were synthesized by a one-step reaction of thieno[2,3-b]thiophene-2,5-dicarboxaldehyde and TREN in acetonitrile via an imine condensation reaction. Moreover, an azobenzene cage [30] for energy efficient and highly selective p-xylene separation was also prepared by condensing (E)-4,4′-(diazene-1,2-diyl)dibenzaldehyde and TREN. Although great efforts have been made in the preparation of functional cages, the construction of new cages with novel properties in high yield remains challenging.
Cadmium is an essential resource on earth, widely used in fertilizers and batteries [31,32,33,34,35]. With the increasing demand for batteries, the pollution of the environment with cadmium ions is becoming increasingly serious. Cadmium is harmful to human health and can cause lung, prostate, breast, or endometrial cancer [36,37]. Therefore, developing new methods to detect cadmium in the environment is urgent. Among the various detection techniques, fluorescent probe detection has been proven to be the best choice, not only for its simplicity and low detection limit, but also for its characteristics of intracellular detection [38,39,40]. Although considerable efforts have been devoted to Cd2+ fluorescent probes, the development of selective fluorescent probes for Cd2+ remains a great challenge due to it being easily interfered with by other transition metals, especially Zn2+ ions in the same group [41,42,43].
Recently, Li and coworkers reported the synthesis of a benzothiadiazole-based macrocycle and found that the intense fluorescence of the macrocycle in the solid state was higher than the monomer [44]. We questioned whether the intense fluorescence of the benzothiadiazole-based cage would increase compared to the monomers. Herein, as a continuation of our interest for supramolecular chemistry [45,46,47,48,49], we report a turn-on fluorescent probe based on a self-assembled cage for the selective detection of Cd2+. The self-assembled cage 1 was obtained by the condensation of 4,4′-(benzothiadiazole-4,7-diyl)dibenzaldehyde and TREN in chloroform via the [3 + 2] approach (Scheme 1). A low precursor concentration can improve the yield of self-assembled cage 1. Interestingly, cage 1 showed a selective response to cadmium ions. The fluorescence intensity of cage 1 was increased almost seven-fold after the addition of cadmium ions, implying that the self-assembled cage represents a promising fluorescent probe for cadmium ion detection.

2. Results and Discussion

First, 4,4′-(benzothiadiazole-4,7-diyl)dibenzaldehyde 2 was prepared according to the reported literature [50]. TREN was used as received without further purification. As shown in Figure 1II, after mixing 2 (1.55 mg, 7.5 mM) and TREN (0.44 mg, 5 mM) in CDCl3 (0.6 mL) at room temperature for 12 h, a small amount of yellow precipitate was formed and the 1H NMR spectrum of the mixture showed one set of new signals that were different from those of 2, suggesting that cage 1 with a relatively high symmetry was formed. However, according to the 1H NMR spectrum results (Figure 1I),compound 2 was not completely converted into cage 1, and a large amount of yellow precipitate would be generated if the reaction time was further prolonged. These results made us doubt whether 2 and TREN could be completely converted to cage 1. We then studied the self-assembly of 2 and TREN at low concentrations. To our surprise, when 2 (1.24 mg, 6 mM) and TREN (0.36 mg, 4 mM) were combined in CDCl3 (0.6 mL) at room temperature for 12 h, only a set of relatively sharp resonances of cage 1 were observed in the 1H NMR spectrum, indicating that 2 and TREN were completely converted to cage 1, and the yield was quantitative (Figure 1III). Protons a-c corresponding to cage 1 shifted remarkably up field (Δδ ≈ 1.82 ppm, 0.61 ppm, 0.71 ppm, and 1.23 ppm, respectively) compared with protons a-c corresponding to 2, indicating that the benzothiadiazole moiety experienced a shielded magnetic environment in the cavity of cage 1. The structure of cage 1 was further confirmed by two-dimensional NMR spectroscopy and mass spectrometry (see Supplementary Materials Figures S3, S4 and S6). Moreover, at a lower concentration, 2 (0.62 mg, 3 mM) and TREN (0.18 mg, 2 mM) in CDCl3 (0.6 mL) could still completely self-assemble into cage 1 (Figure 1IV).
We attempted to obtain a single crystal of cage 1 that is suitable for X-ray diffraction analysis; however, this was unsuccessful. Density functional theory (DFT) methods were used to gain further insights into the formation of cage 1 by using Gaussian 09 software and choosing6-311G as the basis sets. In the optimized structure (Figure 2), the three benzothiazole units of cage 1 are oriented in the manner of a three-blade propeller which may reduce the repulsive force between the benzothiazole units and facilitate the formation of the cage.
In order to expand the application of cage 1, we first tested the coordination properties of cage 1 withvarious metal cations by fluorescence experiments. A number of metal cations, such as Co2+, Ba2+, Pb2+, Mg2+, Zn2+, Fe2+, Ni2+, Ag+, Cu2+, and Cd2+, were added to a solution of cage 1. As shown in Figure 3, it was found that the fluorescence of cage 1 is slightly quenched after the addition of Ag+ or Cu2+, which is different from other results where Ag+ can completely quench the fluorescence of the TREN-based cage [27]. In contrast, Co2+, Ba2+, Pb2+, Mg2+, Zn2+, Fe2+, Ni2+, or Cd2+ are able to enhance the fluorescence of cage 1. It is noteworthy that the addition of 10 equiv of Cd2+ relative to cage 1 could increase the fluorescence by almost seven times, while other cations could only slightly increase the fluorescence of cage 1. This enhancing behavior might potentially be employed to the selective detect of cadmium ions without interference by zinc and other cations. Moreover, 1H NMR experiments were further carried out to investigate the mechanism of Cd2+ to enhance the fluorescence of the cage. As shown in Supplementary Materials Figure S7, after the addition of 0.5 equiv of Cd2+ cation, the protons corresponding to cage 1 shifted remarkably downfield and a small amount of TREN was formed, suggesting that decomposition of the cage may have occurred. Moreover, the fluorescence of cage 1 that is coordinated with the Cd2+ cation is similar to that of compound 2 (Figure 3d), indicating that the fluorescence enhancement may be caused by the decomposition of cage 1. To the best of our knowledge, this system is the first example of cadmium detection based on self-assembled cage formed by tris(2-aminoethyl)amine (TREN) and aldehyde condensation.

3. Materials and Methods

3.1. General Considerations

Unless otherwise noted, all reagents were obtained from commercial suppliers and used without further purification. 1H NMR, 13C NMR spectra were recorded on a Bruker DMX400 NMR spectrometer. 4,4′-(Benzothiadiazole-4,7-diyl)dibenzaldehyde 2 was prepared according to the reported literature [50]. Electrospray ionization mass spectra (ESI-MS) were recorded on a Thermo Fisher®Exactive LC-MS spectrometer (Thermo Fisher Scientific, Waltham, MA, USA).

3.2. Typical Procedure for the Synthesis of Cage 1

Cage 1 was obtained by condensing 2 (1.24 mg, 6 mM) and tris(2-aminoethyl)amine(TREN) (0.36 mg, 4 mM) in CDCl3 (0.6 mL). The solution was sealed in an NMR tube for 12 h without stirring to allow the system to reach equilibrium. 1H NMR (400 MHz, Chloroform-d) δ 8.34 (s, 6H), 7.59 (d, J = 8.0 Hz, 12H), 7.39 (d, J = 8.0 Hz, 12H), 6.71 (s, 6H), 3.91 (s, 12H), 2.90 (s, 12H). 13C NMR (101 MHz, CDCl3) δ 161.96, 153.58, 137.89, 136.15, 132.17, 129.05, 128.41, 127.11, 57.90, 52.22. ESI-MS cald. for [M + Na]+: 1239.42, found: 1239.15.

4. Conclusions

In summary, a turn-on fluorescent probe, cage 1, was self-assembled by condensing 4,4′-(benzothiadiazole-4,7-diyl)dibenzaldehyde and TREN via imine condensation in chloroform. The yield of cage 1 could be improved by lowering the precursor concentration. Moreover, the addition of cadmium ions could remarkably enhance the fluorescence intensity of the system by decomposing cage 1. An easy preparation of cage 1 and a selective turn-on fluorescent sensor for Cd2+ over other cations were achieved using the present system, which is promising for practical applications in the selective detection of Cd2+ in petrochemical industry wastes. Further work will focus on the development of a water-soluble self-assembled cage for the detection of Cd2+ in water and living cells.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28041841/s1, Synthesis of cage 1, Scheme S1; 1H and 13C spectra of cage 1, Figures S1 and S2; 1H-1H COSY and 1H-1H NOESY spectra of cage 1, Figures S3 and S4; Fluorescence photographs of cage 1 after addition of 10 equiv of Cd2+, Figure S5; ESI-MS spectrum of cage 1, Figure S6; 1H NMR spectroscopic of cage 1 after addition of Cd2+, Figure S7.

Author Contributions

Conceptualization, Z.-C.W. and Y.-Z.T.; methodology, H.Y.; investigation, L.-L.T. and W.-H.B.; writing—original draft preparation, F.Z.; supervision, F.Z.; project administration, F.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 China (21602055); Natural Science Foundation of Hunan Province (2017JJ3094).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Durot, S.; Taesch, J.; Heitz, V. Multiporphyrinic Cages: Architectures and Functions. Chem. Rev. 2014, 114, 8542–8578. [Google Scholar] [CrossRef] [PubMed]
  2. Mastalerz, M. Porous Shape-Persistent Organic Cage Compounds of Different Size, Geometry, and Function. Acc. Chem. Res. 2018, 51, 2411–2422. [Google Scholar] [CrossRef] [PubMed]
  3. Montá-Gonzále, G.; Sancenón, F.; Martínez-Máñez, R.; Martí-Centelles, V. Purely Covalent Molecular Cages and Containers for Guest Encapsulation. Chem. Rev. 2022, 122, 13636–13708. [Google Scholar] [CrossRef] [PubMed]
  4. Lehn, J.-M. Supramolecular Chemistry: Concepts and Perspectives; VCH: Weinheim, Germany, 1995. [Google Scholar]
  5. Steed, J.W.; Atwood, J.L. Supramolecular Chemistry, 2nd ed.; Wiley: Chichester, UK, 2009. [Google Scholar]
  6. Zhang, G.; Mastalerz, M. Organic Cage Compounds—From Shape-Persistency to Function. Chem. Soc. Rev. 2014, 43, 1934–1947. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, W.; Stoddart, J.F. Review Emergent Behavior in Nanoconfined Molecular Containers. Chem 2021, 7, 919–947. [Google Scholar] [CrossRef]
  8. Zhang, C.; Chen, C.F. Synthesis and Structure of a TriptyceneBased Nanosized Molecular Cage. J. Org. Chem. 2007, 72, 9339–9341. [Google Scholar] [CrossRef]
  9. Dietrich, B.; Lehn, J.M.; Sauvage, J.P. Les Cryptates. Tetrahedron Lett. 1969, 10, 2889–2892. [Google Scholar] [CrossRef]
  10. Little, M.A.; Cooper, A.I. The Chemistry of Porous Organic Molecular Materials. Adv. Funct. Mater. 2020, 30, 1909842. [Google Scholar] [CrossRef]
  11. Zhao, X.; Liu, Y.; Zhang, Z.Y.; Wang, Y.; Jia, X.; Li, C. One-pot and Shape-controlled Synthesis of Organic Cages. Angew. Chem. Int. Ed. 2021, 60, 17904–17909. [Google Scholar] [CrossRef]
  12. Schneider, M.W.; Oppel, I.M.; Mastalerz, M. ExoFunctionalized Shape-Persistent 2 + 3 Cage Compounds: Influence of Molecular Rigidity on Formation and Permanent Porosity. Chem. Eur. J. 2012, 18, 4156–4160. [Google Scholar] [CrossRef]
  13. Ma, J.X.; Li, J.; Chen, Y.F.; Ning, R.; Ao, Y.F.; Liu, J.M.; Sun, J.; Wang, D.X.; Wang, Q.Q. Cage Based Crystalline Covalent Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 3843–3848. [Google Scholar] [CrossRef]
  14. Fang, S.; Wang, M.; Wu, Y.; Guo, Q.H.; Li, E.; Li, H.; Huang, F. Cagearenes: Synthesis, characterization, and application for programmed vapour release. Chem. Sci. 2022, 13, 6254–6261. [Google Scholar] [CrossRef]
  15. Yu, Y.; Yang, J.M.; Rebek, J. Molecules in Confined Spaces: Reactivities and Possibilities in Cavitands. Chem 2020, 6, 1265–1274. [Google Scholar] [CrossRef]
  16. Warmuth, R. Reactions Inside Carcerands. In Molecular Encapsulation; John Wiley & Sons, Ltd.: Chichester, UK, 2010; pp. 227–268. [Google Scholar]
  17. Pappalardo, A.; Puglisi, R.; Sfrazzetto, G.T. Catalysis inside Supramolecular Capsules: Recent Developments. Catalysts 2019, 9, 630. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, Q.; Catti, L.; Tiefenbacher, K. Catalysis inside the Hexameric Resorcinarene Capsule. Acc. Chem. Res. 2018, 51, 2107–2114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Luo, N.; Ao, Y.F.; Wang, D.X.; Wang, Q.Q. Exploiting anion–π interactions for efficient and selective catalysis with chiral molecular cages. Angew. Chem. Int. Ed. 2021, 60, 20650–20655. [Google Scholar] [CrossRef]
  20. Galan, A.; Ballester, P. Stabilization of Reactive Species by Supramolecular Encapsulation. Chem. Soc. Rev. 2016, 45, 1720–1737. [Google Scholar] [CrossRef]
  21. Ballester, P. Anion Binding in Covalent and Self-Assembled Molecular Capsules. Chem. Soc. Rev. 2010, 39, 3810–3830. [Google Scholar] [CrossRef]
  22. Chakrabarty, R.; Mukherjee, P.S.; Stang, P.J. Supramolecular Coordination: Self-Assembly of Finite Two- and Three-Dimensional Ensembles. Chem. Rev. 2011, 111, 6810–6918. [Google Scholar] [CrossRef] [Green Version]
  23. Cook, T.R.; Stang, P.J. Recent Developments in the Preparation and Chemistry of Metallacycles and Metallacages via Coordination. Chem. Rev. 2015, 115, 7001–7045. [Google Scholar] [CrossRef]
  24. Percástegui, E.G.; Ronson, T.K.; Nitschke, J.R. Design and Applications of Water-Soluble Coordination Cages. Chem. Rev. 2020, 120, 13480–13544. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, Y.; Zhao, W.; Chen, C.H.; Flood, A.H. Chloride capture using a C-H hydrogen-bonding cage. Science 2019, 365, 159–161. [Google Scholar] [PubMed]
  26. MacDowell, D.; Nelson, J. Facile synthesis of a new family of cage molecules. Tetrahedron Lett. 1988, 29, 385–386. [Google Scholar] [CrossRef]
  27. Cao, N.; Wang, Y.; Zheng, X.; Jiao, T.; Li, H. Controllable Self-Assembly of Pills and Cages via Imine Condensation for Silver Cation Detection. Org. Lett. 2018, 20, 7447–7450. [Google Scholar] [CrossRef]
  28. Fang, S.; Sun, W.; Lin, C.; Huang, F.; Li, H. Self-Assembled Cage for In Situ Detecting Silver Cation in Water. Inorg. Chem. 2022, 62, 1776–1780. [Google Scholar] [CrossRef]
  29. Ding, Y.; Alimi, L.O.; Moosa, B.; Maaliki, C.; Jacquemin, J.; Huang, F.; Khashab, N.M. Selective adsorptive separation of cyclohexane over benzene using thienothiophene cages. Chem. Sci. 2021, 12, 5315–5318. [Google Scholar] [CrossRef]
  30. Moosa, B.A.; Alimi, L.O.; Shkurenko, A.; Fakim, A.; Bhatt, P.M.; Zhang, G.; Eddaoudi, M.; Khashab, N.M. A polymorphic azobenzene cage for energy efficient and highly selective p-Xylene separation. Angew. Chem. Int. Ed. 2020, 59, 21367–21371. [Google Scholar] [CrossRef]
  31. Nordberg, G.F.; Herber, R.F.M.; Alessio, L. Cadmium in the Human Environment; Oxford University Press: Oxford, UK, 1992. [Google Scholar]
  32. Shi, C.T.; Huang, Z.Y.; Wu, A.B.; Hu, Y.X.; Wang, N.C.; Zhang, Y.; Shu, W.M.; Yu, W.C. Recent progress in cadmium fluorescent and colorimetric probes. RSC Adv. 2021, 11, 29632–29660. [Google Scholar] [CrossRef]
  33. Kim, H.N.; Ren, W.X.; Kim, J.S.; Yoon, J. Fluorescent and colorimetric sensors for detection of lead, cadmium, and mercury ions. Chem. Soc. Rev. 2012, 41, 3210–3244. [Google Scholar] [CrossRef]
  34. Goshisht, M.K.; Patra, G.K.; Tripathi, N. Fluorescent Schiff base sensors as a versatile tool for metal ion detection: Strategies, mechanistic insights, and applications. Mater. Adv. 2022, 3, 2612–2669. [Google Scholar] [CrossRef]
  35. Chaney, R.L.; Ryan, J.A.; Li, Y.M.; Brown, S.L. Cadmium in Soils and Plants; McLaughlin, M.J., Singh, B., Eds.; Kluwer: Boston, MA, USA, 1999. [Google Scholar]
  36. Goyer, R.A.; Liu, J.; Waalkes, M.P. Cadmium and cancer of prostate and testis. BioMetals 2004, 17, 555–558. [Google Scholar] [CrossRef] [PubMed]
  37. Satarug, S.; Baker, J.R.; Urbenjapol, S.; Haswell-Elkins, M.; Reilly, P.E.B.; Williams, D.J.; Moore, M.R. A global perspective on cadmium pollution and toxicity in non-occupationally exposed population. Toxicol. Lett. 2003, 137, 65–83. [Google Scholar] [CrossRef] [PubMed]
  38. Zhou, Y.; Xu, Z.; Yoon, J. Fluorescent and colorimetric chemosensors for detection of nucleotides, FAD and NADH: Highlighted research during 2004–2010. Chem. Soc. Rev. 2011, 40, 2222–2235. [Google Scholar] [CrossRef]
  39. Quang, D.T.; Kim, J.S. Fluoro- and Chromogenic Chemodosimeters for Heavy Metal Ion Detection in Solution and Biospecimens. Chem. Rev. 2010, 110, 6280–6301. [Google Scholar] [CrossRef]
  40. Zhang, J.F.; Zhou, Y.; Yoon, J.; Kim, J.S. Recent progress in fluorescent and colorimetric chemosensors for detection of precious metal ions (silver, gold and platinum ions). Chem. Soc. Rev. 2011, 40, 3416–3429. [Google Scholar] [CrossRef] [PubMed]
  41. Krishnaveni, K.; Iniya, M.; Jeyanthi, D.; Siva, A.; Chellappa, D. A new multifunctional benzimidazole tagged coumarin as ratiometric fluorophore for the detection of Cd2+/F ions and imaging in live cells. Spectrochim. Acta Part A 2018, 205, 557–567. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, D.B.; Li, S.Y.; Lu, R.M.; Liu, G.; Pu, S.Z. A multi-functional hydrazinobenzothiazole-based diarylethene derivative: Highly efficient discrimination cadmium ion from zinc ion and near-infrared absorption detection of hydroxide ion. Dye. Pigment. 2017, 146, 305–315. [Google Scholar] [CrossRef]
  43. Zhang, X.X.; Wang, R.J.; Fan, C.B.; Liu, G.; Pu, S.Z. A highly selective fluorescent sensor for Cd2+ based on a new diarylethene with a 1,8-naphthyridine unit. Dyes Pigment. 2017, 139, 208–217. [Google Scholar] [CrossRef]
  44. Li, S.; Liu, K.; Feng, X.-C.; Li, Z.-X.; Zhang, Z.Y.; Wang, B.; Li, M.; Bai, Y.-L.; Cui, L.; Li, C. Synthesis and macrocyclization-induced emission enhancement of benzothiadiazole-based macrocycle. Nat. Commun. 2022, 13, 2850. [Google Scholar] [CrossRef]
  45. Ding, M.H.; Tang, L.L.; Liao, J.; Ou, G.C.; Zeng, F. High-yield synthesis of a novel water-soluble macrocycle for selective recognition of naphthalene. Chin. Chem. Lett. 2021, 5, 1665–1668. [Google Scholar] [CrossRef]
  46. Zeng, F.; Liao, J.; Ding, M.H.; Ou, G.C. Self-assembled macrocycle that binds polycyclic aromatic hydrocarbons. Dye. Pigment. 2021, 192, 109430. [Google Scholar] [CrossRef]
  47. Zeng, F.; Cheng, L.; Ou, G.C.; Tang, L.L.; Ding, M.H. Pyromellitic Diimide-Extended Pillar[6]arene: Synthesis, Structure, and Its Complexation with Polycyclic Aromatic Hydrocarbons. J. Org. Chem. 2022, 87, 3863–3867. [Google Scholar] [CrossRef]
  48. Zeng, F.; Cheng, L.; Zhang, W.J.; Tang, L.L.; Wang, X.F. Phenanthrene[2]arene: Synthesis and application as nonporous adaptive crystals in the separation of benzene from cyclohexane. Org. Chem. Front. 2022, 9, 3307–3311. [Google Scholar] [CrossRef]
  49. Zeng, F.; Xiao, X.S.; Gong, S.F.; Yuan, L.; Tang, L.L. An electron-deficient supramolecular macrocyclic host for the selective separation of aromatics and cyclic aliphatics. Org. Chem. Front. 2022, 9, 4829–4833. [Google Scholar] [CrossRef]
  50. Tang, Y.; Huang, H.; Peng, B.; Chang, Y.; Li, Y.; Zhong, C. A thiadiazole-based covalent triazine framework nanosheet for highly selective and sensitive primary aromatic amine detection among various amines. J. Mater. Chem. A 2020, 8, 16542–16550. [Google Scholar] [CrossRef]
Scheme 1. Structural formula and proton designations of cage 1, by condensing the corresponding bisaldehyde precursors 2 with the trisamino linker TREN in CDCl3. Red a, b, c and d are the associated proton signal of precursor 2. Blue a, b, c and d are the associated proton signal of cage 1.
Scheme 1. Structural formula and proton designations of cage 1, by condensing the corresponding bisaldehyde precursors 2 with the trisamino linker TREN in CDCl3. Red a, b, c and d are the associated proton signal of precursor 2. Blue a, b, c and d are the associated proton signal of cage 1.
Molecules 28 01841 sch001
Figure 1. Partial 1H NMR spectra (400 MHz, CDCl3, 298 K) of 2 (spectrum (I)), the self-assembled products by combining 2 (7.5 mM) and TREN (5.0 mM) (spectrum (II)), by combining 2 (6.0 mM) and TREN (4.0 mM) (spectrum (III)), and by combining 2 (3.0 mM) and TREN (2.0 mM) (spectrum (IV)). All spectra were recorded after mixing the corresponding starting materials for 12 h at room temperature, in order to allow the systems to reach equilibrium. Red a, b, c, and d are the associated proton signal of precursor 2. Blue a, b, c, and d are the associated proton signal of cage 1.
Figure 1. Partial 1H NMR spectra (400 MHz, CDCl3, 298 K) of 2 (spectrum (I)), the self-assembled products by combining 2 (7.5 mM) and TREN (5.0 mM) (spectrum (II)), by combining 2 (6.0 mM) and TREN (4.0 mM) (spectrum (III)), and by combining 2 (3.0 mM) and TREN (2.0 mM) (spectrum (IV)). All spectra were recorded after mixing the corresponding starting materials for 12 h at room temperature, in order to allow the systems to reach equilibrium. Red a, b, c, and d are the associated proton signal of precursor 2. Blue a, b, c, and d are the associated proton signal of cage 1.
Molecules 28 01841 g001
Figure 2. DFT-optimized structures of cage 1: (a) side view, (b) top view. Color code: C, gray; N, green; S, yellow. Protons have been omitted for clarity.
Figure 2. DFT-optimized structures of cage 1: (a) side view, (b) top view. Color code: C, gray; N, green; S, yellow. Protons have been omitted for clarity.
Molecules 28 01841 g002
Figure 3. (a) Fluorescence emission spectra of cage 1 after the addition of 10 equiv of various metal cations in the mixture solvent of CHCl3 and CH3CN (v/v = 10:1) at 295 K. Excitation at 372 nm. [1]0 = 1.0 × 10−5 M. (b) Emission spectra (λex = 372 nm) of 1 (5.0 × 10−6 M) in the presence of Cd2+ in the mixture solvent of CHCl3 and CH3CN (v/v = 10:1) at 295 K. The concentration of Cd2+ for curves from top to bottom are 0, 0.5 (0.1 eq of 1), 1.0 (0.2 eq of 1), 1.5 (0.3 eq of 1), 2.0 (0.4 eq of 1), 2.5 (0.5 eq of 1), 3.0 (0.6 eq of 1), 3.5 (0.7 eq of 1), 4.0 (0.8 eq of 1), 4.5 (0.9 eq of 1), 5.0 (1.0 eq of 1), 6.0 (1.2 eq of 1), 7.0 (1.4 eq of 1), 8.0 (1.6 eq of 1), 10.0 (2.0 eq of 1), 12.0 (2.4 eq of 1), and 15.0 (3.0 eq of 1) (× 10−6 M). (c) UV–visspectrum of cage 1. [1]0 = 1.0 × 10−5 M. (d) Fluorescence emission spectra of 2, cage 1, and 1-Cd2+. [2]0 = 1.0 × 10−5 M, [1]0 = 1.0 × 10−5 M. (e) Photograph of the solutions used for (a) under a 365 nm ultraviolet lamp, demonstrating the relative fluorescence intensity of cage 1 upon coordination with various metal cations.
Figure 3. (a) Fluorescence emission spectra of cage 1 after the addition of 10 equiv of various metal cations in the mixture solvent of CHCl3 and CH3CN (v/v = 10:1) at 295 K. Excitation at 372 nm. [1]0 = 1.0 × 10−5 M. (b) Emission spectra (λex = 372 nm) of 1 (5.0 × 10−6 M) in the presence of Cd2+ in the mixture solvent of CHCl3 and CH3CN (v/v = 10:1) at 295 K. The concentration of Cd2+ for curves from top to bottom are 0, 0.5 (0.1 eq of 1), 1.0 (0.2 eq of 1), 1.5 (0.3 eq of 1), 2.0 (0.4 eq of 1), 2.5 (0.5 eq of 1), 3.0 (0.6 eq of 1), 3.5 (0.7 eq of 1), 4.0 (0.8 eq of 1), 4.5 (0.9 eq of 1), 5.0 (1.0 eq of 1), 6.0 (1.2 eq of 1), 7.0 (1.4 eq of 1), 8.0 (1.6 eq of 1), 10.0 (2.0 eq of 1), 12.0 (2.4 eq of 1), and 15.0 (3.0 eq of 1) (× 10−6 M). (c) UV–visspectrum of cage 1. [1]0 = 1.0 × 10−5 M. (d) Fluorescence emission spectra of 2, cage 1, and 1-Cd2+. [2]0 = 1.0 × 10−5 M, [1]0 = 1.0 × 10−5 M. (e) Photograph of the solutions used for (a) under a 365 nm ultraviolet lamp, demonstrating the relative fluorescence intensity of cage 1 upon coordination with various metal cations.
Molecules 28 01841 g003
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.-C.; Tan, Y.-Z.; Yu, H.; Bao, W.-H.; Tang, L.-L.; Zeng, F. A Benzothiadiazole-Based Self-Assembled Cage for Cadmium Detection. Molecules 2023, 28, 1841. https://doi.org/10.3390/molecules28041841

AMA Style

Wang Z-C, Tan Y-Z, Yu H, Bao W-H, Tang L-L, Zeng F. A Benzothiadiazole-Based Self-Assembled Cage for Cadmium Detection. Molecules. 2023; 28(4):1841. https://doi.org/10.3390/molecules28041841

Chicago/Turabian Style

Wang, Zong-Cheng, Ying-Zi Tan, Hui Yu, Wen-Hu Bao, Lin-Li Tang, and Fei Zeng. 2023. "A Benzothiadiazole-Based Self-Assembled Cage for Cadmium Detection" Molecules 28, no. 4: 1841. https://doi.org/10.3390/molecules28041841

Article Metrics

Back to TopTop