Next Article in Journal
Polymorphism and Phase-Transition Thermodynamic Properties of Phenazone (Antipyrine)
Previous Article in Journal
An Overview of α-Pyrones as Phytotoxins Produced by Plant Pathogen Fungi
Previous Article in Special Issue
Application of Intelligent Response Fluorescent Probe in Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Triazine Calixarene as a Dual-Channel Chemosensor for the Reversible Detection of Cu2+ and I Ions via Water Content Modulation

1
Department of Brewing Engineering, Moutai Institute, Renhuai 564500, China
2
Key Laboratory of Macrocycle and Supramolecular Chemistry of Guizhou Province, Guiyang 550025, China
3
Biological Physics Laboratory, Department of Physics and Astronomy, School of Natural Science, The University of Manchester, Oxford Road, Manchester M13 9PL, UK
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(13), 2815; https://doi.org/10.3390/molecules30132815
Submission received: 8 May 2025 / Revised: 26 June 2025 / Accepted: 27 June 2025 / Published: 30 June 2025

Abstract

Rationally designing and synthesizing chemosensors capable of simultaneously detecting both anions and cations via water content modulation is challenging. In this study, we synthesized and characterized a novel triazine calixarene derivative-based iodide and copper ion-selective fluorescent “turn-off” sensor. This dual-channeled fluorescent probe is able to recognize Cu2+ and I ions simultaneously in aqueous systems. The fluorescent sensor s4 was synthesized by displacement reaction of acridine with 1, 3-bis (dichloro-mono-triazinoxy) benzene in acetonitrile. Mass spectrometry (MS), UV-vis, and fluorescence spectra were acquired to characterize the fluorescence response of s4 to different cations and anions, while infrared (IR) spectroscopy and isothermal titration calorimetry (ITC) were employed to study the underlying selectivity mechanism of s4 to Cu2+ and I. In detail, s4 displayed extremely high sensitivity to Cu2+ with over 80% fluorescence decrement caused by the paramagnetic nature of Cu2+ in the aqueous media. The reversible fluorescence response to Cu2+ and the responses to Cu2+ in the solution of other potential interferent cations, such as Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+ were also investigated. Probe s4 also exhibited very good fluorescence selectivity to iodide ions under various anion (F, Cl, Br, NO3, HSO4, ClO4, PF6, AcO, H2PO4) interferences. In addition to the fluorescent response to I, s4 showed a highly selective naked-eye-detectable color change from colorless to yellow with the other tested anions.

1. Introduction

Although some heavy metal ions are vital for maintaining human metabolism and play key roles in living systems, they can be highly toxic, posing significant environmental and health risks [1]. Iron (III), zinc (II), copper (II), cobalt (II), and manganese (II) are essential elements, yet their high concentrations can lead to adverse health effects [1,2,3]. Among these elements, copper (II) stands out for its critical involvement in numerous cellular processes, including gene expression and protein function [4]. Meanwhile, iodide is considered as an important ion among several other anions (PO43+, HCO3 and Cl) for its various biological activities, for example, in thyroid function, normal growth, neurological activity, and brain function [5,6]. On the other hand, consumption of iodide in excess can cause adverse effects on human health [7].
Therefore, the development of increasingly selective and sensitive methods for the detection of iodide and copper is currently receiving considerable attention [4,8,9]. Several methods, including atomic absorption spectroscopy, inductively coupled plasma atomic emission spectrometry, electrochemical sensing, and the use of piezoelectric quartz crystals make it possible to detect low limits [10,11,12]. However, these methods require sophisticated equipment and are time-consuming, limiting their use to trained professionals. As an alternative, colorimetric sensors, which respond visually, offer several advantages, including simplicity, sensitivity, selectivity, and economic viability, without the need for specialized instrumentation [13,14].
Recently, Huang et al. synthesized a rhodamine-B derivative that functions as a “turn-on” probe for monitoring copper (II) ions in living cells [15]. This probe is colorless and exhibits weak fluorescence in acetonitrile. Upon the addition of copper (II) ions, however, the probe induces the appearance of a purple-red color and a pronounced orange-red fluorescence. The observed absorption and fluorescence changes are attributed to a two-step process. Initially, the copper (II) ions promote the ring opening of the probe. This is followed by a redox reaction between the probe and the copper (II) ions, leading to the reduction of copper (II) to copper (I). Lee and colleagues developed a novel iodide chemosensor that operated effectively over a wide concentration range of iodide ions in a CH3CN/H2O (99/1, v/v) solution, with no interference from other anions [16]. To enhance the detection limit of iodide in aqueous solutions, Chen et al. designed a new iodide ion chemosensor based on a hydrazone derivative. This sensor exhibited a distinct colorimetric response to iodide, achieving a detection limit as low as 1.0 × 10−4 mol/L through naked-eye color changes and 1.1 × 10−6 mol/L through changes in absorption spectra [17].
However, the development of a highly sensitive and selective chemosensor capable of simultaneously detecting multiple ions remains a significant challenge, particularly in response to specific conditions. Zhang et al. reported a fluorescent sensor with aggregation-induced emission (AIE) properties, which was highly sensitive to both iodide (I) and mercury (Hg2+) ions. The sensor demonstrated its suitability for detecting low concentrations of I and Hg2+ in real samples [18]. Lee and co-workers developed sequential recognition of Zn2+ and Cu2+ using a new anthracene-containing dipyridylamine-based receptor [19]. This receptor exhibited highly selective and sensitive fluorescent “off–on” recognition of Zn2+, while the resulting receptor–Zn2+ complex displayed high selectivity to Cu2+ through a decrease in fluorescence intensity, demonstrating that the receptor–Zn2+ complex could detect Cu2+ via metal displacement. Qin et al. designed a multifunctional fluorescent chemosensor for detecting Cu2+ and Zn2+. The sensor exhibited significant fluorescence enhancement in the presence of Zn2+ but was quenched by Cu2+, an effect attributed to the paramagnetic nature of the Cu2+ species [20]. The development of a selective chemosensor for both I and Cu2+ ions is challenging due to the distinct properties of these ions, such as the large size and high polarizability of iodide and the paramagnetic nature of copper (II) [21,22]. In most of the reported Cu2+ fluorescent chemosensors, because of the paramagnetic nature of Cu2+, the binding of the metal ion causes a quenching of the fluorescence emission and leads to a “turn-off” signal [1]. However, very few Cu2+ chemosensors have the potential for I detection [23,24,25,26]. Patil and colleagues designed a PPT-1 receptor that showed a naked-eye-detectable color change from colorless to red in the presence of Cu2+ over the other tested cations [27]. In contrast, iodide ions did not change the color of PPT-1, but resulted a new spectroscopic response in the absorption spectrum at 232 nm, and a detection limit as low as 0.22 µM was achieved in aqueous solution. Meanwhile, rationally designing and synthesizing chemosensors capable of simultaneously detecting both anions and cations via water content modulation is still challenging.
1,3,5-Triazines are heterocyclic compounds containing three N atoms in a six-membered ring. These three nitrogen atoms can coordinate with metal ions, and their stability and properties are comparable to benzene due to π–electron conjugation. The compound exhibits high biological activity, including antibacterial, antihypertensive, and analgesic effects. Furthermore, the reactivity of the threew chlorine atoms on the triazine ring varies, allowing hierarchical substitution by controlling reaction conditions to selectively introduce one or more substituents on the triazine ring. This study utilized pyridine as the parent compound and 1,3,5-triazines as the connecting arm, to design and synthesize a triazine-based heteroaromatic compound s4 for the highly sensitive and selective detection of both I ions and Cu2+ via water content modulation.

2. Results and Discussion

2.1. The Solvent Effects of Fluorescent Probe s4

To study the solvent effects on the fluorescent efficiency of s4, the fluorescence spectra of probe s4 were obtained in the presence of various solvents such as CH3CN, DMSO, DMF, 1,4-dioxane, and THF (Figure 1B). It was found that probe s4 showed good solubility (100 µM s4 in CH3CN for the further measurements) and the highest fluorescence emission intensity in CH3CN. Furthermore, the effects of water content and subsequent pH conditions in the solvent systems on the fluorescence intensity of s4 were investigated (Figure 1C). When the water content ranged from 0 to 50%, the probe s4 exhibited a slight increase in fluorescence intensity, and the fluorescence intensity increased further in the water content system > 50%. Based on the stable emission intensity of s4 in the presence of the solvents with different water contents, the mixed solvent system CH3CN/H2O (v/v, 9/1) was selected for the following tests of pH effects (Figure 1D). The results indicated that probe s4 showed relatively stable fluorescence emission in the tested range of pH from 2.5 to 8.8. All the subsequent experiments were carried out in presence of CH3CN/H2O (v/v, 9/1) with a pH at 6.5 to mimic physiological conditions.

2.2. The Fluorescent Emission Quenching of Probe s4 in Response to Cu2+ and I

We evaluated the fluorescent selectivity of s4 in response to various monovalent anions (F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4) in CH3CN solution. As shown in Figure 2A, the fluorescence emission of probe s4 was significantly quenched after the addition of I in comparison to the other anions. The corresponding inset graph showed the phenomenon that under 365 nm UV irradiation, the initial blue solution of probe s4 became darker upon the addition of iodide ions. The results of the fluorescent titration curves showed the gradual fluorescence quenching of s4 with the addition of I due to the heavy atom effect; the quenching efficiency reached 88% with the I/s4 ratio at 40 (Figure 2B). The observed static quenching effect was attributed to the spontaneous bonding of probe s4 and I in the ground state. To verify the experimental results with the static quenching constant (Ksv), which was calculated to be 3.09 × 104 mol/L, the static fluorescence quenching rate constant (Kq) was derived as 3.09 × 1012 mol/L, which was much greater than 2.0 × 1010 mol/L. The quenching trend suggested static quenching aligning with the formation of a s4–I complex. Furthermore, both the molar ratio and the Job method revealed the stoichiometric ratio of I to the probe s4 as 1:1, with a detection limitation of 8.23 × 10−8 mol/L. In the UV–vis spectra of s4 interacting with different anions, only I could lead to the decreased absorption intensity of the probe s4 at 254 nm and cause the obvious red shift shown in Figure 2C. Moreover, the initial solution of the probe s4 was colorless under white light, and it changes to a bright yellow color after the addition of iodide ions (inset graph in Figure 2C). With the increasing concentration of I, the absorption intensity of s4 at 254 nm decreased gradually along with a more obvious red shift. This indicated interactions between the molecules and iodide ions. Spectral titration of the probe s4 with I suggested a 1:1 stoichiometric ratio of I to the probe s4, consistent with the previous fluorescence spectrum titration results. The stability constant of the I complex was calculated to be 4.36 × 104 L·mol−1. Considering the UV spectrum, this red shift may have been due to the spontaneous bonding of probe s4 and I, which induced the formation of a complex between the molecule and iodide, thereby altering the original electronic configuration of the molecules, leading to the red shift in the spectrum. The transition from colorless to yellow under white light illumination suggests the involvement of heavy atom effects. In the excited state, the formation of an iodide anion–probe charge transfer complex, coupled with the reduction in energy due to the heavy atom effect, would have accounted for the observed fluorescence quenching [28,29]. However, the formation of s4–I complex did not change the molecular structure of s4. Meanwhile, the curves of fluorescent emission at high concentrations of iodine ion overlapped, indicating that the coordination of s4 molecules with I aligned with the static quenching mechanism.
In contrast, to improve the multifunctionality of the chemosensor, we simply added acidic aqueous solution into CH3CN to evaluate its cation recognition performance. Probe s4 was further studied for its recognition of different cations. Various cations including Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+, and Cu2+ mixed with s4 in CH3CN/H2O (v/v, 9/1, pH 6.5) were tested via fluorescence emission (Figure 3A). Among all the selected cations, the addition of Cu2+ significantly reduced the fluorescence intensity of the probe s4. The inset graph in Figure 3A shows the disappearance of blue fluorescence in s4 solution, indicating the good detection capability of s4 for Cu2+ in aqueous media. To describe the selectivity of s4 to Cu2+ quantitatively, measurements of s4 responses to Cu2+ with different molar ratios were carried out. The fluorescence quenching rate in s4 reached 85% in the presence of a 100-fold concentration of Cu2+. Both the molar ratio and Job’s method indicated that the coordination ratio of probe s4 to Cu2+ was 1:1. The binding constant of Cu2+ to probe s4 was calculated to be 5.71 × 104 mol/L. Although both I and Cu2+ caused fluorescence quenching, giving rise to the blue fluorescence disappearance in the solution, we also observed distinct colorimetric responses under white light: the solution turned from colorless to yellow upon addition of I, whereas no color changing occurred with Cu2+ ion. This selective chromogenic behavior can be used as a dual-mode colorimetric detection chemosensor.
To evaluate the selectivity of s4, the interference of coexisting metal ions or anions in the recognition of Cu2+ and I, respectively, by the probe s4 was investigated (Figure 4A and Figure 5B). The results of the fluorescent emission showed that s4 exhibited high selectivity for both Cu2+ and I, unperturbed by the addition of other competitive metal ions and anions. This was also quite evident from the iodide-induced shift in UV–vis absorbance maxima at 254 nm remaining unperturbed. These results confirm a great deal of selectivity for the detection of iodide ion compared with other anions (Figure 5A).
Reversibility is important for the continuous recyclability and reusability of chemosensors [10]. The reversible nature of s4 towards Cu2+ in CH3CN/H2O (9/1, v/v, pH 6.5) with EDTA was characterized as shown in Figure 4B,C. Upon the addition of EDTA (20-fold molar concentration of probe s4) to the solution containing s4 and Cu2+, the fluorescence intensity of the s4–Cu2+ complex recovery indicated the regeneration of s4. After that, the addition of 50-fold Cu2+ (relative to probe s4) resulted in a decrease in fluorescence emission at 425 nm as the Cu2+ was initially added. The results indicated that the excess Cu2+ reformed the s4–Cu2+ complex with free probe in solution, indicating that the coordination interaction was a reversible process. The time required for the complete reversible process was 10 s.
To confirm the reversible nature of s4 towards I, Ag+ was used as a competitive reagent (Figure 5C,D). With the addition of Ag+ (with a 20-fold molar ratio of the probe) to the probe s4–I complex solution, the fluorescence spectrum of the system was similar to that of the probe without adding I, indicating that the addition of Ag+ had taken up the I from the complex, forming a AgI compound and releasing free probe. Subsequently, the same operation demonstrated that the system still exhibited good reversibility on fluorescence restoring, because the weak interactions between the probe s4 and iodide allowed iodide ions to be easily displaced by the silver ions.

2.3. The Concentration and Temperature Effects of Fluorescent Probe s4

Under the experimental conditions described above, a calibration curve was established for determining the concentration of Cu2+ and I, using fluorescence spectroscopy to analyze the probe s4 (10 μM) (Figure 6A,C). The fluorescence intensity of the probe s4 showed a linear correlation with the concentration of Cu2+ in the range from 8.0 × 10−6 mol/L to 3.5 × 10−4 mol/L, with a correlation coefficient R = 0.9978 (n = 11). For the probe s4–I calibration curve, there was a linear correlation in range of I concentration from 5.0 × 10−6 mol/L to 2.8 × 10−4 mol/L, with a correlation coefficient R = 0.9971 (n = 13). Ten sets of blank solutions were measured and the relative standard deviation (RSD) was calculated to be 2.31% (Cu2+) and 2.46% (I), with a detection limit as low as 7.08 × 10−8 mol/L for Cu2+ ion and for 8.23 × 10−8 for I ion.
The effects of temperature on the fluorescence intensity of probe s4 in the presence of different concentrations of Cu2+ or I were measured, and the fitting curves were obtained according to the previously published method, as shown in Figure 6B,D [17,21,30]. The slopes of the quenching curves for both probe s4–Cu2+ and s4–I systems decreased with the temperature increase from 293.15 K to 303.15 K, indicating a static quenching led by temperature effect. The following Stern-Volmer equation was applied:
F0/F = 1 + Ksv [Q] = 1 + Kq × τo [Q]
The static quenching constant Ksv was calculated to be 5.97 × 103 mol/L (Cu2+) and 3.09 × 104 mol/L (I), and the static fluorescence quenching rate constant Kq was obtained to be 5.97 × 1011 (Cu2+) and 3.09 × 1012 (I). The Kq values for Cu2+ and I were both greater than 2.0 × 1010 L·mol−1, indicating that the quenching process was caused by the formation of s4–Cu2+ and s4–I complexes, respectively.

2.4. The Mechanism of Cu2+ Fluorescent Probe s4

To explore the underlying mechanisms between s4 and Cu2+ or I for the emission quenching of s4, the corresponding ITC measurements in CH3CN solvent were carried out (Figure 7A,B). The binding of probe s4 to Cu2+ is an endothermic reaction and tends to be equilibrate with the increased Cu2+ concentration. From the nonlinear fitting curve of the molar ratio of probe s4 to Cu2+ based on the reaction heat, the binding constant was calculated to be Ka = (8.55 ± 0.52) × 104 mol/L, number of binding sites n = 0.964 ± 0.030, molar binding enthalpy ΔH° = (−331.2 ± 2.04) kJ/mol, molar binding entropy TΔS° = −302.92 kJ/mol, and molar binding free energy ΔG° = (−28.28 ± 2.04) kJ/mol. The coordination ratio and binding constant were in agreement with the spectral measurement results. Unlike the binding between s4 and Cu2+, the ITC curve of the interaction between probe s4 and I in CH3CN solvent reflected a typically exothermic reaction. Through the non-linear fitting of binding reaction heat between the probe s4 and I, the parameters that were yielded included binding constant Ka = (7.20 ± 0.63) × 104 mol/L, number of binding sites n = 0.921 ± 0.055, molar binding enthalpy ΔH° = (9.35 ± 1.19) kJ/mol, molar binding entropy TΔS° = 37.09 kJ/mol, and molar binding free energy ΔG° = (−27.74 ± 1.19) kJ/mol. These results indicate that the binding of the probe s4 to iodide ion is a spontaneous procedure. The coordination ratio and binding constant were consistent with the results obtained from the fluorescence spectra.
To explore the sensing mechanism of s4 to Cu2+ and I on the molecular level, the IR spectra were further investigated, which revealed characteristic structural changes of s4 occurring upon interaction with different ions (Figure 7C,D). In the IR spectra of s4 before and after addition with Cu2+, the N-H stretching vibrations of the probe s4 at 3259 cm−1 and 3321 cm−1 clearly shifted towards higher wavenumbers at 3416 cm−1 and 3522 cm−1. Additionally, the double peaks of the C-N in s4 (at 1383 cm−1 and 1426 cm−1) merged into a single peak with increased intensity after s4 binding with Cu2+. The peaks of the N-H bending vibrations at 1574 cm−1 and 1610 cm−1 became single peaks with reduced intensity, suggesting that Cu2+ may participate in coordination with the N atom in the secondary amine of the probe s4. Similarly, when s4 coordinated with I, the stretching vibration absorption peaks of N-H shifted to higher wavenumbers along with the peak becoming broader. Moreover, the C-N peaks at 1383 cm−1 and 1426 cm−1 shifted to higher wavenumbers at 1398 cm−1 and 1471 cm−1, and their intensities decreased. The N-H bending vibrations at 1574 cm−1 and 1610 cm−1 became single peaks with weakened intensities. These observations suggest that I may participate in the coordination process with the secondary amine in the probe molecule.

3. Conclusions

In summary, we have developed a novel chemosensor s4, featuring a triazine calixarene structure. This sensor exhibits specific selectivity and high sensitivity in detecting copper (II) ions (Cu2+) and iodide ions (I) via water content modulation. The fluorescent response to Cu2+ is marked by a striking color transition from blue to dark under UV light, which aligns with the paramagnetic properties of Cu2+. The remarkable colorimetric sensing of probe s4 confirmed a 1:1 (probe s4–Cu2+) binding model with the detection limit down to 7.08 × 10−8 mol·L−1. Other than the fluorescent recognition of I, a unique colorimetric response to I based on UV-vis spectra is realized through the coordination with probe s4. In particular, competitive anions such as F, Cl, Br, NO3, HSO4, ClO4, PF6, AcO, H2PO4 did not afford any obvious interference response. The detection limits of I were found to be 8.23 × 10−8 mol·L−1 according to the naked-eye color changes and absorption spectral changes respectively. This recognition behavior makes probe s4 a potential tool to detect Cu2+ and I in environmental and life sciences.

4. Materials and Experimental Methods

All the chemicals were purchased commercially as analytical pure reagents. Triazine calixarene was purchased from Sigma (Sigma Aldrich Technology., Ltd., St. Louis, MO, USA) and used without further purification. The metal ion solution was prepared from its perchlorate. The anion solution was prepared using its tetrabutylammonium salt with acetonitrile. High-purity water (ElgaUltrapure with a resistivity of 18.2 MΩ·cm) was used in all the experiments.
Mass spectra were recorded on an Agilent LC/MSD spectrometer, based on infusion methods. UV–vis absorption studies were carried out on a TU-1901 UV–Vis spectrophotometer (Beijing Puxi General Instrument Co., Ltd., Beijing, China). Fluorescence measurements were performed on a Varian Cary Eclipse spectrofluorimeter equipped with quartz cuvettes of 1 cm path length. The excitation and emission slit widths were 5.0 nm. All absorption and emission spectra were recorded at 24 ± 1 °C. Stock solutions of probe s4 (100 µM in CH3CN/H2O, 9:1, v/v, pH 6.5) were prepared immediately before the experiments. The solutions of metal ions were prepared from perchlorates of Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+. The solutions of anions were prepared from tetrabutylammonium salts of F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4. Stock solutions of Cu2+ and I was prepared with Cu(ClO4)2·6H2O and tetrabutylammonium iodide at a concentration of 2 mM. Tris–HCl solution with a concentration of 2 mM was prepared as buffer solution to use for pH control of the solvent systems.

4.1. Molecular Design and Synthesis

Based on the previous work using calixarene as the parent body, a new macrocycle molecule was synthesized, in which acridine as a main body covalently linked with the triazine derivative group. The above reagents (acridine and triazine derivative) were synthesized according to procedures published previously [26,31]. The synthesis method followed the following procedure. N,N-Diisopropyl-ethylamin (DIPEA, 10.5 mM in acetonitrile) as an acid binding agent was mixed with aminomethyl acridine (0.42 mM in acetonitrile); after that, the triazine derivative (0.42 mM 1, 3-bis (dichloro-mono-triazinoxy) benzene in acetonitrile) was dropped into the solution. The final macrocyclic calixarene was obtained by the nucleophilic substitution reaction in a nitrogen atmosphere under 40 °C for 48 h. The specific synthesis and characterization are described in a previous report [32].

4.2. Fluorescence and UV-Vis Spectra

In a series of 10.0 mL volumetric flasks, 1.0 mL of probe s4 acetonitrile stock solution, 1.0 mL Tris-HCl buffer, and 1.0 mL of Cu2+ or other metal ion (I or other anion solution in case of anion recognization) aqueous solution were added, and diluted to the mark with CH3CN/H2O (v/v, 9/1) for further use in UV–vis and fluorescence spectral analysis. The excitation wavelength of the fluorescence spectra was set at 246 nm.

4.3. Isothermal Titration Calorimetry

The ITC measurements were carried out with Nano ITC (TA Instrument Co., Ltd., USA). Firstly, 1.0 mL probe s4 solution (100 µM in DMF) was injected into the calorimetric titration cell. The titration syringe was filled with 250 µL Cu2+ solution (1 mM in DMF). Each 6 µL Cu2+ solution was injected into the calorimetric cell at an interval of 180 s, while keeping stirring at a rate of 250 rpm until equilibrium. The I titration procedure was the same as above, with the solvent selected as CH3CN. All the tests were carried out at a temperature of 298.15 K.

4.4. IR Spectra

The IR spectra for the chemosensor s4 and the s4–I and s4–Cu2+ complexes were recorded. The IR spectroscopic analysis was conducted by using a Bruker Vertex 70 FTIR Spectrophotometer (Bruker, Billerica, MA, USA). A uniform resolution of 2 cm−1 was maintained in all cases.

Author Contributions

Overall analysis, conceptualization, and manuscript writing: F.W.; funding support: J.G., T.S. and L.C.; financial and analytical resources: H.Z.; analysis: L.Z.; synthesis: L.J.; synthesis route design: R.W.; Data analysis: M.Y.; manuscript revision support: M.L. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge PhD studentship support from the University of Manchester (UoM) and the China Scholarship Council (UoM-CSC joint PhD programme) to M.L. Thanks for the financial support from the Fund of Zunyi Technology and Big data Bureau, Moutai institute Joint Science and Technology Research and Development Project (ZunShiJiaoHe HZ Zi [2024] 371), the Science and Technology Foundation of Guizhou Education Department (No. Qianjiaohe KY zi[2020]035), Modern Baijiu Brewing Technology Engineering Research Center of Guizhou Universities (Qianjiaoji [2023] No. 028) and the Scientific Research Foundation of Moutai Institute (mygccrc[2024]008, mygccrc[2024]014, mygccrc[2022]007, mygccrc(2022)001). Guizhou Engineering Research Center for Specialty Food Resources (KY[2020]022). Guizhou Province Technology Innovation Center for Jiangxiangxing Baijiu, Qiankehe Platform JSZX (2025) 002.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in this study.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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] [PubMed]
  2. Orvig, C.; Abrams, M.J. Medicinal Inorganic Chemistry: Introduction. Chem. Rev. 1999, 99, 2201–2204. [Google Scholar] [CrossRef]
  3. McRae, R.; Bagchi, P.; Sumalekshmy, S.; Fahrni, C.J. In Situ Imaging of Metals in Cells and Tissues. Chem. Rev. 2009, 109, 4780–4827. [Google Scholar] [CrossRef]
  4. Sivaraman, G.; Iniya, M.; Anand, T.; Kotla, N.G.; Sunnapu, O.; Singaravadivel, S.; Gulyani, A.; Chellappa, D. Chemically diverse small molecule fluorescent chemosensors for copper ion. Coord. Chem. Rev. 2018, 357, 50–104. [Google Scholar] [CrossRef]
  5. Dai, G.; Levy, O.; Carrasco, N. Cloning and characterization of the thyroid iodide transporter. Nature 1996, 379, 458–460. [Google Scholar] [CrossRef]
  6. Li, M.; Eastman, C.J. The changing epidemiology of iodine deficiency. Nat. Rev. Endocrinol. 2012, 8, 434–440. [Google Scholar] [CrossRef] [PubMed]
  7. Leung, A.M.; Braverman, L.E. Consequences of excess iodine. Nat. Rev. Endocrinol. 2013, 10, 136–142. [Google Scholar] [CrossRef] [PubMed]
  8. Trusso Sfrazzetto, G.; Satriano, C.; Tomaselli, G.A.; Rizzarelli, E. Synthetic fluorescent probes to map metallostasis and intracellular fate of zinc and copper. Coord. Chem. Rev. 2016, 311, 125–167. [Google Scholar] [CrossRef]
  9. Valdes-García, J.; Rosales-Vázquez, L.D.; Bazany-Rodríguez, I.J.; Dorazco-González, A. Recent Advances in Luminescent Recognition and Chemosensing of Iodide in Water. Chem.-Asian J. 2020, 15, 2925–2938. [Google Scholar] [CrossRef]
  10. Velmurugan, K.; Prabhu, J.; Tang, L.; Chidambaram, T.; Noel, M.; Radhakrishnan, S.; Nandhakumar, R. A simple chalcone-based fluorescent chemosensor for the detection and removal of Fe3+ ions using a membrane separation method. Anal. Methods 2014, 6, 2883–2888. [Google Scholar] [CrossRef]
  11. Vassos, B.H.; Hirsch, R.F.; Letterman, H. X-ray microdetermination of chromium, cobalt, copper, mercury, nickel, and zinc in water using electrochemical preconcentration. Anal. Chem. 1973, 45, 792–794. [Google Scholar] [CrossRef]
  12. Becker, J.S.; Zoriy, M.V.; Pickhardt, C.; Palomero-Gallagher, N.; Zilles, K. Imaging of Copper, Zinc, and Other Elements in Thin Section of Human Brain Samples (Hippocampus) by Laser Ablation Inductively Coupled Plasma Mass Spectrometry. Anal. Chem. 2005, 77, 3208–3216. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Y.; Zhao, Q.; Feng, W.; Li, F. Luminescent Chemodosimeters for Bioimaging. Chem. Rev. 2013, 113, 192–270. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, W.; Guo, S.; Hu, C.; Fan, J.; Peng, X. Recent Development of Chemosensors Based on Cyanine Platforms. Chem. Rev. 2016, 116, 7768–7817. [Google Scholar] [CrossRef]
  15. Yu, M.; Shi, M.; Chen, Z.; Li, F.; Li, X.; Gao, Y.; Xu, J.; Yang, H.; Zhou, Z.; Yi, T.; et al. Highly Sensitive and Fast Responsive Fluorescence Turn-On Chemodosimeter for Cu2+ and Its Application in Live Cell Imaging. Chem.-A Eur. J. 2008, 14, 6892–6900. [Google Scholar] [CrossRef]
  16. Lee, D.Y.; Singh, N.; Kim, M.J.; Jang, D.O. Chromogenic and Fluorescent Recognition of Iodide with a Benzimidazole-Based Tripodal Receptor. Org. Lett. 2011, 13, 3024–3027. [Google Scholar] [CrossRef]
  17. Chen, J.; Lin, Q.; Li, Q.; Li, W.-T.; Zhang, Y.-M.; Wei, T.-B. A highly selective colorimetric chemosensor for detection of iodide ions in aqueous solution. RSC Adv. 2016, 6, 86627–86631. [Google Scholar] [CrossRef]
  18. Zhang, R.X.; Li, P.F.; Zhang, W.J.; Li, N.; Zhao, N. A highly sensitive fluorescent sensor with aggregation-induced emission characteristics for the detection of iodide and mercury ions in aqueous solution. J. Mater. Chem. C 2016, 4, 10479–10485. [Google Scholar] [CrossRef]
  19. Lee, H.G.; Kim, K.B.; Park, G.J.; Na, Y.J.; Jo, H.Y.; Lee, S.A.; Kim, C. An anthracene-based fluorescent sensor for sequential detection of zinc and copper ions. Inorg. Chem. Commun. 2014, 39, 61–65. [Google Scholar] [CrossRef]
  20. Qin, J.-C.; Yang, Z.-Y. Design of a novel coumarin-based multifunctional fluorescent probe for Zn2+/Cu2+/S2− in aqueous solution. Mater. Sci. Eng. C 2015, 57, 265–271. [Google Scholar] [CrossRef]
  21. Barman, S.; Sadhukhan, M. Facile bulk production of highly blue fluorescent graphitic carbon nitride quantum dots and their application as highly selective and sensitive sensors for the detection of mercuric and iodide ions in aqueous media. J. Mater. Chem. 2012, 22, 21832. [Google Scholar] [CrossRef]
  22. Lou, X.; Ou, D.; Li, Q.; Li, Z. An indirect approach for anion detection: The displacement strategy and its application. Chem. Commun. 2012, 48, 8462. [Google Scholar] [CrossRef]
  23. Patil, S.R.; Nandre, J.P.; Jadhav, D.; Bothra, S.; Sahoo, S.K.; Devi, M.; Pradeep, C.P.; Mahulikar, P.P.; Patil, U.D. Imatinib intermediate as a two in one dual channel sensor for the recognition of Cu2+ and I−ions in aqueous media and its practical applications. Dalton Trans 2014, 43, 13299–13306. [Google Scholar] [CrossRef]
  24. Zhou, H.; Wang, S.; Jin, Y.; Pang, X.F.; Zhao, Q.; Zhang, T.; Zhang, J. A near-infrared “turn-on” fluorescent probe for selective detection of copper (II) ions in aqueous media and its application in cell imaging. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2025, 336, 126036. [Google Scholar] [CrossRef] [PubMed]
  25. Zavalishin, M.; Gamov, G.; Kiselev, A.; Nikitin, G. A fluorescein conjugate as colorimetric and red-emissive fluorescence chemosensor for selective recognition Cu2+ ions. Opt. Mater. 2024, 153, 115580. [Google Scholar] [CrossRef]
  26. Yu, M.; Jiang, L.; Mou, L.; Zeng, X.; Wang, R.; Peng, T.; Wu, F.; Shi, T. A “Pincer” Type of Acridine–Triazole Fluorescent Dye for Iodine Detection by Both ‘Naked-Eye’Colorimetric and Fluorometric Modes. Molecules 2024, 29, 1355. [Google Scholar] [CrossRef] [PubMed]
  27. Shi, T.; Xie, Z.; Mo, X.; Feng, Y.; Peng, T.; Wu, F.; Yu, M.; Zhao, J.; Zhang, L.; Guo, J. Synthesis and Application of Salicylhydrazone Probes with High Selectivity for Rapid Detection of Cu2+. Molecules 2024, 29, 2032. [Google Scholar] [CrossRef]
  28. Tang, H.; de Oliveira, C.S.; Sonntag, G.; Gibb, C.L.D.; Gibb, B.C.; Bohne, C. Dynamics of a Supramolecular Capsule Assembly with Pyrene. J. Am. Chem. Soc. 2012, 134, 5544–5547. [Google Scholar] [CrossRef]
  29. Kong, F.; Meng, X.; Chu, R.; Xu, K.; Tang, B. A turn-on fluorescence probe for imaging iodide in living cells based on an elimination reaction. Chem. Commun. 2015, 51, 6925–6927. [Google Scholar] [CrossRef]
  30. Wang, M.-X.; Yang, H.-B. A General and High Yielding Fragment Coupling Synthesis of Heteroatom-Bridged Calixarenes and the Unprecedented Examples of Calixarene Cavity Fine-Tuned by Bridging Heteroatoms. J. Am. Chem. Soc. 2004, 126, 15412–15422. [Google Scholar] [CrossRef]
  31. Laronze-Cochard, M.; Kim, Y.-M.; Brassart, B.; Riou, J.-F.; Laronze, J.-Y.; Sapi, J. Synthesis and biological evaluation of novel 4,5-bis(dialkylaminoalkyl)-substituted acridines as potent telomeric G-quadruplex ligands. Eur. J. Med. Chem. 2009, 44, 3880–3888. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, R.-X.; Mu, L.; Zeng, X.; Sun, Q.; Chen, Y.; Li, J. Synthesis and Crystal Structure of a Novel Homo[2]Heterocalix[2]arene[2]triazines containing Acridine Moiety. Asian J. Chem. 2013, 25, 8566–8568. [Google Scholar] [CrossRef]
Figure 1. (A) Molecular structure of fluorescent probe s4. (B) The solvent effects on the fluorescent efficiency of s4; the test concentration of the compound was 10 µM. (C,D) The effects of H2O and pH on the fluorescent efficiency of fluorescent probe s4 (10 µM), with fluorescent excitation and emission wavelength of λexem = 246/425 nm.
Figure 1. (A) Molecular structure of fluorescent probe s4. (B) The solvent effects on the fluorescent efficiency of s4; the test concentration of the compound was 10 µM. (C,D) The effects of H2O and pH on the fluorescent efficiency of fluorescent probe s4 (10 µM), with fluorescent excitation and emission wavelength of λexem = 246/425 nm.
Molecules 30 02815 g001
Figure 2. (A) Fluorescence and (C) UV–vis absorption spectra of different s4–anion complexes (the anions including F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4), all the measurements were conducted with the concentration of s4 at 10 μM. The inset graphs in (A,C) showed the color change of s4 before and after addition with I under UV light and ambient light, respectively. (B) Fluorescence and (D) UV–vis absorption spectra of s4–I titration curves (I concentration ranging from 0 to 40 or 30 eq, CH3CN). Each fluorescence spectrum was recorded at 246 nm. The two insert graphs in (B,D) are the corresponding mole ratio (the upper part) and Job’s plot (the lower part) curves of s4 reacted with I, according to fluorescence (λexem = 246/425 nm) and UV–vis absorption spectra.
Figure 2. (A) Fluorescence and (C) UV–vis absorption spectra of different s4–anion complexes (the anions including F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4), all the measurements were conducted with the concentration of s4 at 10 μM. The inset graphs in (A,C) showed the color change of s4 before and after addition with I under UV light and ambient light, respectively. (B) Fluorescence and (D) UV–vis absorption spectra of s4–I titration curves (I concentration ranging from 0 to 40 or 30 eq, CH3CN). Each fluorescence spectrum was recorded at 246 nm. The two insert graphs in (B,D) are the corresponding mole ratio (the upper part) and Job’s plot (the lower part) curves of s4 reacted with I, according to fluorescence (λexem = 246/425 nm) and UV–vis absorption spectra.
Molecules 30 02815 g002
Figure 3. (A) The fluorescence spectra of different s4–cation systems (cations including Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+, Cu2+). All the measurements were conducted with the concentration of s4 at 10 μM. The inset graph shows the color change of s4 before and after addition of Cu2+ under UV light. (B) Fluorescence spectra of s4–Cu2+ titration curves (Cu2+ concentration ranging from 0 to 100 eq, CH3CN/H2O, 9/1, v/v, pH 6.5), at λexem = 246/425 nm. Each spectrum was recorded at 246 nm. The insert graphs in (B) show the mole ratio (the upper part) and Job’s plot (the lower part) of s4’s reaction with Cu2+.
Figure 3. (A) The fluorescence spectra of different s4–cation systems (cations including Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+, Cu2+). All the measurements were conducted with the concentration of s4 at 10 μM. The inset graph shows the color change of s4 before and after addition of Cu2+ under UV light. (B) Fluorescence spectra of s4–Cu2+ titration curves (Cu2+ concentration ranging from 0 to 100 eq, CH3CN/H2O, 9/1, v/v, pH 6.5), at λexem = 246/425 nm. Each spectrum was recorded at 246 nm. The insert graphs in (B) show the mole ratio (the upper part) and Job’s plot (the lower part) of s4’s reaction with Cu2+.
Molecules 30 02815 g003
Figure 4. (A) The selective effects of Cu2+ complexing with s4 in the solution environments of various metal ions (200 µM, Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+, Cu2+). The black bars represent the addition of the competing cations to the s4 solution, while the red bars represent the addition of competing cations and Cu2+ to the s4 solution. (B,C) The fluorescence spectra of reversible reaction between s4 and Cu2+ upon addition of EDTA, and the corresponding change in fluorescent intensity between “on” and “off” states. The test concentration of s4 in all the solution systems was 10 µM in CH3CN/H2O (9/1, v/v, pH 6.5), at a wavelength of λexem = 246/425 nm.
Figure 4. (A) The selective effects of Cu2+ complexing with s4 in the solution environments of various metal ions (200 µM, Li+, Na+, K+, Ca2+, Cd2+, Zn2+, Sr2+, Ni2+, Co2+, Cu2+). The black bars represent the addition of the competing cations to the s4 solution, while the red bars represent the addition of competing cations and Cu2+ to the s4 solution. (B,C) The fluorescence spectra of reversible reaction between s4 and Cu2+ upon addition of EDTA, and the corresponding change in fluorescent intensity between “on” and “off” states. The test concentration of s4 in all the solution systems was 10 µM in CH3CN/H2O (9/1, v/v, pH 6.5), at a wavelength of λexem = 246/425 nm.
Molecules 30 02815 g004
Figure 5. The selective effects of I complexing with s4 in the solution environments of various anions (200 μM, F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4), indicated by (A) UV–vis and (B) fluorescence spectra. The black bars represent the addition of the competing anions to the s4 solution, while the red bars represent the addition of competing anions and I to the s4 solution. (C) The fluorescence spectra of the reversible reaction between s4 and I upon addition of Ag+, (D) and the corresponding change of fluorescent intensity between “on” and “off” states. The test concentration of s4 in all the solution systems was 10 µM in CH3CN, at a wavelength of λexem = 246/425 nm.
Figure 5. The selective effects of I complexing with s4 in the solution environments of various anions (200 μM, F, Cl, Br, I, NO3, HSO4, ClO4, PF6, AcO, H2PO4), indicated by (A) UV–vis and (B) fluorescence spectra. The black bars represent the addition of the competing anions to the s4 solution, while the red bars represent the addition of competing anions and I to the s4 solution. (C) The fluorescence spectra of the reversible reaction between s4 and I upon addition of Ag+, (D) and the corresponding change of fluorescent intensity between “on” and “off” states. The test concentration of s4 in all the solution systems was 10 µM in CH3CN, at a wavelength of λexem = 246/425 nm.
Molecules 30 02815 g005
Figure 6. (A,C) Plots of fluorescence intensity at λexem = 246/425 nm for a mixture of s4 (10 μM in CH3CN/H2O, v/v, 9/1, pH 6.5) with Cu2+ and I in CH3CN at different concentrations. (B,D) Stern–Volmer curves of s4 with Cu2+ and I at temperatures of 293.15 and 303.15 K.
Figure 6. (A,C) Plots of fluorescence intensity at λexem = 246/425 nm for a mixture of s4 (10 μM in CH3CN/H2O, v/v, 9/1, pH 6.5) with Cu2+ and I in CH3CN at different concentrations. (B,D) Stern–Volmer curves of s4 with Cu2+ and I at temperatures of 293.15 and 303.15 K.
Molecules 30 02815 g006
Figure 7. (A,B) Binding of s4 to Cu2+ or I. Isothermal calorimetric titration (ICT) of s4 with Cu2+ in DMF and I in CH3CN (upper part). Raw experimental data and calorimetric titration curves for the binding of Cu2+ or I to s4 (lower part). (C,D) IR spectra of compound s4 and s4–Cu2+ or s4–I in KBr disks.
Figure 7. (A,B) Binding of s4 to Cu2+ or I. Isothermal calorimetric titration (ICT) of s4 with Cu2+ in DMF and I in CH3CN (upper part). Raw experimental data and calorimetric titration curves for the binding of Cu2+ or I to s4 (lower part). (C,D) IR spectra of compound s4 and s4–Cu2+ or s4–I in KBr disks.
Molecules 30 02815 g007
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

Wu, F.; Chen, L.; Yu, M.; Zhao, L.; Jiang, L.; Shi, T.; Guo, J.; Zheng, H.; Wang, R.; Liao, M. Triazine Calixarene as a Dual-Channel Chemosensor for the Reversible Detection of Cu2+ and I Ions via Water Content Modulation. Molecules 2025, 30, 2815. https://doi.org/10.3390/molecules30132815

AMA Style

Wu F, Chen L, Yu M, Zhao L, Jiang L, Shi T, Guo J, Zheng H, Wang R, Liao M. Triazine Calixarene as a Dual-Channel Chemosensor for the Reversible Detection of Cu2+ and I Ions via Water Content Modulation. Molecules. 2025; 30(13):2815. https://doi.org/10.3390/molecules30132815

Chicago/Turabian Style

Wu, Fuyong, Long Chen, Mei Yu, Liang Zhao, Lu Jiang, Tianzhu Shi, Ju Guo, Huayan Zheng, Ruixiao Wang, and Mingrui Liao. 2025. "Triazine Calixarene as a Dual-Channel Chemosensor for the Reversible Detection of Cu2+ and I Ions via Water Content Modulation" Molecules 30, no. 13: 2815. https://doi.org/10.3390/molecules30132815

APA Style

Wu, F., Chen, L., Yu, M., Zhao, L., Jiang, L., Shi, T., Guo, J., Zheng, H., Wang, R., & Liao, M. (2025). Triazine Calixarene as a Dual-Channel Chemosensor for the Reversible Detection of Cu2+ and I Ions via Water Content Modulation. Molecules, 30(13), 2815. https://doi.org/10.3390/molecules30132815

Article Metrics

Back to TopTop