Previous Article in Journal
Aromaticity and Antiaromaticity: How to Define Them
Previous Article in Special Issue
Self-Reporting H2S Donors: Integrating H2S Release with Real-Time Fluorescence Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phenylthiadiazole-Based Schiff Base Fluorescent Chemosensor for the Detection of Al3+ and Zn2+ Ions

by
Jorge Heredia-Moya
1,*,
Ariana Fiallos-Ayala
1 and
Amanda Cevallos-Vallejo
2
1
Centro de Investigación Biomédica (CENBIO), Facultad de Ciencias de la Salud Eugenio Espejo, Universidad UTE, Quito 170527, Ecuador
2
Escuela de Ciencias Químicas, Facultad de Ciencias Exactas y Naturales, Pontificia Universidad Católica del Ecuador, Quito 170525, Ecuador
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(4), 128; https://doi.org/10.3390/chemistry7040128
Submission received: 16 June 2025 / Revised: 31 July 2025 / Accepted: 6 August 2025 / Published: 8 August 2025
(This article belongs to the Special Issue Organic Chalcogen Chemistry: Recent Advances)

Abstract

Aluminum (Al) and zinc (Zn) are two of the most widely used metals in industry, and their excessive accumulation in the body has been linked to serious diseases like Alzheimer’s, Parkinson’s, and cancer. This highlights the need for effective ways to detect and measure them. In this study, we synthesized the fluorescent chemosensor 1, which contains a Schiff base and a 1,3,4-thiadiazole ring in its structure, and evaluated its fluorescent response in the presence of various metal ions. The chemosensor enabled the selective quantification of Al3+ and Zn2+ ions through excitations at different wavelengths, yielding differentiated fluorescent emissions. For Al3+, excitation at 370 nm generated a strong emission at 480 nm, whereas for Zn2+, excitation at 320 nm led to a new small broad emission at 560 nm. We established detection limits of 2.22 × 10−6 M for Al3+ and 1.62 × 10−5 M for Zn2+; their binding stoichiometry was found to be 1:1 for Al3+ and 2:1 for Zn2+, based on Job’s plot analysis. These results show that chemosensor 1 is a promising tool for detecting Al3+ and Zn2+.

1. Introduction

Aluminum (Al) is the third most abundant element after silicon and oxygen, and the most predominant metallic element in the Earth’s crust [1,2]. As a metal, it exhibits several characteristics, such as high electrical conductivity, corrosion resistance, non-adsorptiveness, and low density, which enable its extensive use in everyday applications, including water treatment, kitchenware, vehicle and computer manufacturing, and medical devices [3]. However, it is also considered to be one of the most important environmental toxicants, and since it is a non-essential element for humans, its excessive accumulation in the body has been associated with various medical conditions, such as Alzheimer’s disease, Parkinson’s disease, Wilson syndrome, osteoporosis, osteomalacia, colic, gastrointestinal problems, kidney damage, and certain cancers, among others [4,5,6,7,8]. According to the WHO, the tolerable weekly intake of aluminum is estimated at 2 mg/kg of body weight, based on a non-observed adverse-effect level of 30 mg/kg per day [9]. Therefore, the detection of Al3+ in solution is crucial for both environmental and biomedical monitoring.
In contrast, zinc (Zn) is the fourth most consumed metal worldwide and is important due to its role in biological, environmental, and industrial contexts. Zinc’s properties, which include its high corrosion resistance and ability to form alloys, make it invaluable for galvanizing steel and in battery manufacturing [10]. Biologically, Zn2+ plays an important role in many physiological processes, including enzyme function, the immune response, and cellular metabolism [11]. However, overexposure to zinc has been linked to several diseases, including prostate cancer, neurodegenerative disorders, and oxidative stress conditions [12].
Several analytical methods for detecting Al3+ and Zn2+ ions have been reported, such as atomic absorption spectroscopy (AAS), inductively coupled plasma mass spectrometry (ICP-MS), and colorimetric techniques [13,14,15]. Despite their accuracy, these methods are often costly, time-consuming, and require specialized instrumentation and training. The detection of Al3+ ions is particularly challenging due to its poor coordination ability, strong hydration shell, and lack of distinct spectroscopic characteristics [16,17]. Zinc, although easier to coordinate than aluminum, may also suffer from interference by other bivalent or transition metals in complex matrices [18,19].
Numerous fluorescent chemosensors that are capable of selectively detecting Al3+ and Zn2+ ions have been developed over recent years, employing a variety of molecular scaffolds and sensing mechanisms. Schiff base derivatives, in particular, have been extensively studied due to their facile synthesis and strong coordination ability with metal ions, facilitating fluorescence “turn-on” responses [20,21,22]. Various mechanisms, such as excited-state intramolecular proton transfer (ESIPT), chelation-enhanced fluorescence (CHEF), and photoinduced electron transfer (PET), have been exploited to achieve high selectivity and sensitivity [23]. Despite these advances, challenges remain in designing dual chemosensors that can simultaneously detect Al3+ and Zn2+ ions with distinct fluorescence signals in aqueous media [21,24]. Furthermore, the interference from competing ions and the stability of the complexes are important factors that affect sensor performance [25]. Therefore, developing simple, robust, and selective solution-phase fluorescent probes remains an active and relevant area of research.
Among the various scaffolds explored for chemosensing, nitrogen-containing heterocyclics such as 1,3,4-thiadiazole have demonstrated notable fluorescence activity and coordination properties [26,27]. Fluorescent probes incorporating thiadiazole units have been used for selective Al3+ [28] and Zn2+ detection [29]. However, to the best of our knowledge, no dual-functional thiadiazole-based chemosensor capable of detecting both Al3+ and Zn2+ ions in solution with distinct responses has been reported to date.
Herein, we report a selective fluorescent “turn-on” probe, based on a 1,3,4-thiadiazole Schiff base derivative, which is capable of functioning as a dual-analyte chemosensor for the detection of Al3+ and Zn2+ ions in solution. This system demonstrates high selectivity and distinct emission behavior for each metal ion, and operates effectively under mild conditions, offering a practical alternative to more complex detection methodologies.

2. Materials and Methods

2.1. Synthesis of Ligand 1

The fluorescent probe shown in Scheme 1 was synthesized as previously reported [30,31]. The experimental procedure, along with the spectroscopic data, is provided in the Supplementary Materials (Figures S1–S4).

2.2. UV–Vis and Fluorescent Measurements

The thiadiazole 1 was dissolved in methanol to make a 2 mM stock solution. The stock solutions of the testing cation were prepared from CuCl2·2H2O, CoCl2·6H2O, MnCl2·4H2O, FeCl3·6H2O, NiCl2·6H2O, CaCl2·2H2O, SnCl2·2H2O, Pb(NO3)2, FeSO4·7H2O, HgCl2, AgNO3, ZnCl2, and AlCl3, dissolved in methanol (20 mM). All the solutions were stored at −20 °C and heated to room temperature before use. Prior to taking spectroscopic measurements, fresh solutions of the cations (1 mM) were prepared by diluting the stock solutions in methanol.
The UV-vis absorption and fluorescence emission spectra were determined with a 96-well plate using a microplate reader, the Cytation 5 (Agilent BioTek, Winooski, VT, USA) spectrophotometer. The absorption and emission spectra of sensor 1 in the presence of various metal ions were measured in methanol solvent in concentrations of 100 μM and 25 μM, respectively.

3. Results and Discussion

The UV-vis absorption spectrum of thiadiazole 1 was measured in methanol as solvent, and its behavior in the presence of various metal cations, such as Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+ (25 μM in methanol), was evaluated (Figure 1). The free ligand exhibited two main absorption bands at 320 and 370 nm. Upon the addition of metal ions to the ligand, both absorption bands decreased and a hypsochromic shift of the absorption band at 320 nm was observed (310 nm for Cu2+, Co2+, Mn2+, Pb2+, Ag+, and Zn2+ ions, 300 nm for Fe3+, Sn2+ and Al3+ ions, and 290 nm for Ni2+ and Ca2+ ions). The main decrease in the absorption band at 370 nm was observed for Sn2+; however, this result suggested that 1 could not be used for the selective detection of any ions by absorbance.
The fluorescence spectra of thiadiazole 1 were determined after excitation at both wavelengths, 320 and 370 nm. After excitation at 320 nm, a strong emission band at 380 nm was observed. In contrast, for excitation at 370 nm, a small broad emission band at 560 nm was observed (Figure 2). A 100 μM methanolic solution of 1 was treated with various metal ions (25 μM), and fluorescence was measured upon excitation at 320 nm. A moderate enhancement (30–40%) in the emission at 380 nm was observed in the presence of Fe3+, Sn2+, and Al3+, while other ions showed a negligible response. Notably, Zn2+ induced a distinct broad emission band centered at λem = 560 nm (Figure 3A and Figure S5), corresponding to a 5.7-fold increase in intensity at that wavelength, relative to the baseline of the free ligand. This enhancement, although moderate in absolute intensity, is highly selective and reproducible (Figure S6). Conversely, when it is excited at 370 nm, a strong emission is observed for Al3+, with a maximum emission at 480 nm (Figure 3B), indicating that 1 exhibits an “off–on” mode in response to this ion, exhibiting a high sensitivity toward Al3+ over other metal ions that were used.
The fluorescent response of 1 was evaluated in the presence and absence of Al3+ (Figure S7) and Zn2+ (Figure S8) in various solvents. When excited at 320 nm and in the presence of Al3+, the emission band at 380 nm was observed only in ethanol and methanol; however, the response in methanol was 13.77 times greater than that observed in ethanol. No fluorescence was observed in the other solvents, except in the case of water, where the ligand only showed a broad signal with two maxima at 560 and 590 nm. When 1 was excited at 370 nm, no fluorescence was observed, except when using water. In this solvent, a broad emission band was observed again, with two maxima at 560 and 590 nm. In the presence of Zn2+ in ethanol and acetonitrile, an emission band was observed at 550 nm, accompanied by reductions in the intensity of the emission band by 45% and 36%, respectively, compared to the emission band observed at 560 nm in methanol. For both metal ions, the highest fluorescent response of chemosensor 1 was observed in methanol.
The fluorescent response of chemosensor 1 to Al3+ and Zn2+ was evaluated in a methanol–water system at varying ratios. For the 1-Al3+ complex, it was observed that the intensity of the fluorescence emission signal decreased as the water concentration increased, compared to a solution of pure methanol (Figure S9). At a 10% water concentration, the signal intensity dropped by 30% compared to pure methanol (Figure S10). As the water ratio increased, the fluorescence intensity declined significantly. This decrease in intensity may be due to the aggregation of compounds at higher water concentrations. This leads to an aggregation-caused quenching (ACQ) effect, which results in reduced fluorescence intensity [32].
In the case of Zn2+, a similar decrease in fluorescence emission intensity was observed with increasing water concentration, but the reduction was more pronounced (Figure S11). In the presence of 10% water, the signal intensity decreased by 55% compared to that observed in pure methanol (Figure S12). When the water concentration exceeded 60%, the same broad emission signal observed in pure water began to appear, so complete fluorescence quenching was not observed.
The fluorescence response of 1 to various concentrations of Al3+ (0–200 µM) and Zn2+ (0–1000 µM) was investigated. With increasing concentrations of Al3+, the fluorescence intensity at 480 nm of the solution gradually increased (Figure 4A). The chemosensor 1 showed a notable linear relationship (R2 = 0.9986) between the fluorescence intensity at 480 nm and the concentrations of Al3+ from 0 to 100 µM, suggesting that this thiadiazole was potentially valuable for the quantitative determination of Al3+. The same behavior was seen when detecting Zn2+ at 560 nm, showing a linear relationship of R2 = 0.9992 between the fluorescence intensity at 560 nm and Zn2+ concentrations ranging from 15 to 1000 µM (Figure 4B). The detection limits (S/N = 3) of chemosensor 1 were determined to be 2.22 × 10−6 M for Al3+ and 1.62 × 10−5 M for Zn2+.
The reversibility of the chemosensor in the presence of Al3+ and Zn2+ was evaluated through a reversible experiment. When an EDTA solution (>4.0 eq) was added to the Al3+ complex, the chemosensor was released, resulting in a loss of fluorescence. Adding Al3+ ions along with the EDTA solution repeatedly led to an on–off fluorescence response, and this cycle was successfully repeated up to four times (Figure 5). Conversely, when an EDTA solution was added to the 1-Zn2+ complex, the fluorescence disappeared after the addition of one equivalent of EDTA. However, in contrast to what happened with Al3+, the complex formed with Zn2+ was not reversible, since adding more Zn2+ ions did not restore the original fluorescence (see Figures S13 and S14).
Additionally, we evaluated the interaction time between chemosensor 1 and Al3+ in methanol and a methanol–water mixture (9:1). As shown in Figure 6, the fluorescence response of 1 to the Al3+ ion increases steadily with time after adding the chemosensor, and a gradual rise in the fluorescent signal occurs in both solvents; however, the fluorescence is higher in pure methanol, as expected. In methanol, the fluorescent signal increases progressively during the first 60 min, after which it stabilizes. Conversely, in the case of the methanol–water system (9:1), the fluorescent response stabilizes after the first 30 min and remains constant throughout the remainder of the experiment. When compared to results in the literature, it can be observed that many probes exhibit response times that are often within seconds. However, probes with longer response times may still be suitable for applications where immediate detection is not critical, and signal stability is more important.
The combination mode between 1 and Al3+ and Zn2+ ions was investigated using Job’s plot at 480 nm for Al3+ and 560 nm for Zn2+. The fluorescence intensity reached its maximum value at 480 nm when the mole fraction of the Al3+ ion was 0.5 (Figure 7A). This suggests the formation of a 1:1 complex between the chemosensor and Al3+. This finding contrasts with the report published by Manna et al., which indicated the formation of a 2:1 complex with this metal ion despite the similar ligand structure [28]. The difference in coordination modes between the two studies suggests variations in the binding affinity of the ligands for Al3+, which may be attributed to factors such as electronic or solvation effects. Further studies are required to elucidate the precise mechanism behind this observation. Conversely, the fluorescence intensity of Zn2+ reached its maximum value at 560 nm when the mole fraction of the Zn2+ ion was about 0.30 (Figure 7C). This indicates the formation of a 2:1 complex between 1 and Zn2+.
Linear fitting of the fluorescence titration profiles using a Benesi–Hildebrand plot was also examined using the following equation [33,34]:
F m a x F 0 F X F 0 = 1 + 1 K [ M ] n
where Fmax, F0, and FX are the fluorescence intensities of the probe in the presence of a metal ion at saturation, free probe, and any intermediate metal ion concentration, respectively, and n is 1 for Al3+ and ½ for Zn2+.
Plotting of (FmaxF0)/(FXF0) versus 1/[Al3+] showed a linear relationship (R2 = 0.9889), which also strongly supports the 1:1 binding stoichiometry of 1 and Al3+ (Figure 7B), and the binding constant K was calculated to be 2.93 × 109 M−1. In the case of Zn2+, (FmaxF0)/(FXF0) versus 1/[Zn2+]1/2 was plotted, also exhibiting a good linear relationship (R2 = 0.9819), which is expected in the case of 2:1 binding stoichiometry of 1 and Zn2+ (Figure 7C). The binding constant K for this ion is 3.5 M−2.
To further understand the recognition mechanism of chemosensor 1 for Al3+, studies using 1H NMR and Fourier transform infrared (FT-IR) spectroscopy were conducted. The 1H NMR spectra of chemosensor 1 were investigated in both the absence and presence of Al3+ in DMSO-d6. As shown in Figure 8, when Al3+ was added to 1, the imine signal (Ha) at a chemical shift of 9.27 ppm shifted to 10.25 ppm. Also, the phenolic hydroxyl proton (Hb) was not seen, which suggests that the oxygen atom of the phenolic hydroxyl group and the nitrogen atom of the imine were involved in the coordination between chemosensor 1 and Al3+.
In the IR spectra (Figure 9), the characteristic C=N stretching vibration observed at 1602 cm−1 in the free ligand shifted to 1626 cm−1 in the complex, which was consistent with metal coordination through the imine nitrogen. Simultaneously, the broad phenolic O–H band diminished, indicating deprotonation and coordination through the phenolic oxygen. The appearance of new bands in the 400–500 cm−1 region, assignable to Al–O and Al–N vibrations, confirms the formation of a coordination complex.
Considering the bidentate nature of the ligand and the preference of Al3+ for six-coordinate geometries, the formation of an octahedral complex is proposed. The coordination sphere is plausibly completed by two chloride anions and one methanol molecule, resulting in a distorted octahedral arrangement stabilized by both electronic and steric factors.
The IR spectrum of the 1-Zn2+ complex (Figure S15) shows the appearance of new absorption bands at 454 and 551 cm−1, which are absent in the free ligand and can be assigned to Zn–S vibrational modes. These bands indicate coordination through the sulfur atom of the benzothiazole moiety. Notably, the bands associated with the hydroxyl and imine groups remain largely unchanged, which is consistent with the absence of significant chemical shift variations in the 1H NMR spectrum of the complex (Figure S16), suggesting that these groups are not directly involved in coordination. In addition, the disappearance or shifts of specific ligand peaks (e.g., at 691 and 1418 cm−1) further support complex formation. These results confirm metal–ligand coordination and are consistent with the formation of a 2:1 ligand-to-metal assembly, in which two ligand molecules chelate a single Zn2+ ion, predominantly via sulfur atoms. These features, combined with the d10 electronic configuration of Zn2+, suggest a distorted tetrahedral geometry for the [L2Zn] complex, as commonly observed in similar N,S-donor Schiff base zinc complexes [35,36].
The selectivity of thiadiazole 1 (50 µM) for Al3+ over other metal ions (1.0 eq.) was evaluated by a competition experiment. The addition of Al3+ to the ligand–metal mixture generated a significant increase in fluorescence for almost all ions, reaching values close to those observed for the pure Al3+ ligand complex, except in the case of Fe2+ (Figure 10A). This pattern suggests that Al3+ displaces most other metals from the coordination site, forming a highly fluorescent complex, likely due to the ligand’s higher affinity toward Al3+, possibly due to its high charge, small radius, and tendency to rigid coordination geometries [37]. In the case of the 1–Fe2+ complex, the addition of Al3+ produced a fluorescence that was 17.8 times lower than that observed in the 1–Al3+ complex (Figure 10B). This suggests that the complex formed between the ligand and Fe2+ was much more stable, probably due to a stronger, specific, or irreversible coordination mode, which prevented the efficient displacement of Fe2+ by Al3+.
When evaluating the selectivity of thiadiazole 1 (50 mM) for Zn2+ compared to other metal ions (1.0 eq.), a more limited selectivity profile was observed than that for Al3+ (Figure 11). This reduced selectivity was particularly evident when the metal ions formed more stable complexes with the ligand, which hindered the effective coordination of Zn2+. For Cu2+, Fe3+, Sn2+, Fe2+, and Al3+, there was no significant increase in fluorescence at 560 nm after adding Zn2+. The fluorescence emission was between 1.7 and 15.4 times lower than what was seen in the 1–Zn2+ complex. A similar effect occurred with Co2+ and Ni2+; however, the increase in fluorescence after adding Zn2+ was slightly higher. Nonetheless, the observed emission was still 1.7 times lower than that from the 1–Zn2+ complex. This suggests a preference of 1 for hard metal ions, which prevents Zn2+ coordination by forming stronger or more rigid complexes with the ligand, due to partial competition for the coordination site [18] or possibly due to incomplete or reversible displacement by Zn2+ [38].
In contrast, the addition of Zn2+ to the complex of chemosensor 1 with Mn2+, Ca2+, Pb2+, Hg2+, and Ag+ produced a higher fluorescent response, with fluorescence emissions between 1.2 and 2.9 times higher than that observed in the complex 1–Zn2+. This suggests that these metal ions form complexes of high kinetic lability or low thermodynamic stability with the ligand and can easily be replaced by Zn2+. However, the results also suggest that the mechanism is not limited to the complete replacement of the initial metal ion by Zn2+, but that additional phenomena may be occurring that enhance the fluorescent emission. Possible explanations include the presence of multiple fluorescent species, such as supramolecular aggregates or heterobimetallic complexes, which contribute to overall emission [39,40]. Additionally, these metal ions might cause structural changes in the ligand, allowing it to interact more effectively with Zn2+, increasing the emission intensity.
The free ligand exhibited negligible fluorescence at 480 nm, likely due to the operation of an excited-state intramolecular proton transfer (ESIPT) process, which enables non-radiative decay pathways and quenches emission [41]. Upon complexation with Al3+, a significant fluorescence enhancement was observed. This effect is attributable to the inhibition of ESIPT via the coordination of Al3+ to the phenolic oxygen and imine nitrogen atoms, which rigidifies the ligand framework, suppresses intramolecular proton transfer, and enhances the conjugation of the π-system (Scheme 2). The resulting increase in emission is characteristic of the chelation-enhanced fluorescence (CHEF) mechanism, as reported for related Schiff-base systems [41].
Nevertheless, in the absence of competing ions, the 1–Zn2+ complex exhibits a marked fluorescence enhancement compared to the free ligand, emission from which is otherwise suppressed. This turning-on behavior is attributable to a mechanism involving the inhibition of non-radiative deactivation pathways, particularly photo-induced electron transfer (PET). In the free ligand, fluorescence quenching is likely due to PET from the imine nitrogen lone pair to the excited chromophore, and possibly due to internal conversion, facilitated by conformational flexibility. Upon Zn2+ coordination, as suggested by the appearance of Zn–N vibrational modes in the IR spectrum (bands at 454 and 551 cm−1) and the absence of significant chemical shift changes in the OH and N=CH signals in the 1H NMR spectrum, the metal center binds preferentially through the nitrogen atom of the benzothiazole ring (see Scheme 2). This interaction inhibits PET by reducing the electron-donating capacity of the nitrogen atom and rigidifies the structure, thereby enhancing conjugation and favoring radiative decay. This coordination-induced conformational restriction leads to increased fluorescence emission, which is consistent with the behavior observed in other Zn2+ complexes involving nitrogen donor atoms in similar heteroaromatic systems [36,42].
The use of compound 1 in real sample applications that involve Zn2+ is limited by its low selectivity for this metal ion compared to other metal ions. The literature, however, discusses several methods that could improve this selectivity. These methods include adding nanomaterials, making structural changes, altering functional groups, or using auxiliary receptors [43]. More research is necessary to understand the true capabilities of this chemosensor in relation to Zn2+.
Chemosensor 1 was compared to other fluorescent probes made from salicylaldehyde for detecting Al3+, as shown in Table S1 in the Supplementary Materials. The limits of detection (LOD) show that sensor 1 is in a good range, with values below the World Health Organization (WHO) guideline for drinking water, which is 7.41 µM [44]. In contrast, analysis of the binding constants reveals that 1 exhibits a significantly higher affinity for Al3+ compared to the other probes, suggesting that it may be more effective for detecting this metal ion in solutions.

4. Conclusions

The fluorescent probe, 2-(2-hydroxybenzal)amino-5-phenyl-1,3,4-thiadiazole (1), shows potential for selectively detecting Al3+ and Zn2+ ions in methanol. This chemosensor allows the identification of these metal ions by displaying different fluorescence responses upon excitation at specific wavelengths. The excitation of 1 in the presence of Al3+ at 370 nm increases fluorescence emission at 480 nm, whereas excitation in the presence of Zn2+ at 320 nm results in a new emission band at 560 nm. Detection limits for Al3+ and Zn2+ were established at 2.22 × 10−6 M and 1.62 × 10−5 M, respectively. The binding stoichiometry differs between the two ions, forming a 1:1 complex with Al3+ and a 2:1 complex with Zn2+, and the calculated binding constant for Al3+ is 2.93 × 109 M−1, and for Zn2+ is 3.5 M−2.
Competitive experiments further show the ligand’s preferential affinity for Al3+, whereas interference from other metal ions limits Zn2+ coordination. These findings underscore the utility of thiadiazole 1 in metal detection, emphasizing the need for further research to optimize its selectivity and explore its applications in sensing technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry7040128/s1: Synthesis of chemosensor 1. Figure S1: 1H-NMR spectrum of 2; Figure S2: 13C-NMR spectrum of 2; Figure S3: 1H-NMR spectrum of 1; Figure S4: 13C-NMR spectrum of 1; Figure S5: Zoom of the fluorescence emission spectra (λex = 320 nm) of thiadiazol 1 (100 µM) in the presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 µM) in methanol solvent; Figure S6: Relative fluorescence emission of thiadiazole 1 (100 μM) in presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 μM) in methanol solvent. (λex = 320 nm; λem = 560 nm); Figure S7: Fluorescence emission spectrum (λex = 370 nm) of thiadiazole 1 (100 μM) in the presence and absence of Al3+ (25 μM) in different solvents; Figure S8: Fluorescence emission spectrum (λex = 320 nm) of thiadiazole 1 (100 μM) in the presence and absence of Zn2+ (25 μM) in different solvents; Figure S9: Fluorescence emission spectrum (λex = 370 nm) of thiadiazole 1 (100 μM) in the presence of Al3+ (25 μM) as a function of the water fraction in the binary MeOH/water system; Figure S10: Fluorescence intensity vs. water fraction curve for thiadiazole 1 (100 μM) in the presence of Al3+ (25 μM); Figure S11: Fluorescence emission spectrum (λex = 320 nm) of thiadiazole 1 (100 μM) in the presence of Zn2+ (25 μM) as a function of the water fraction in the binary MeOH/water system; Figure S12: Fluorescence intensity vs. water fraction curve for thiadiazole 1 (100 μM) in the presence of Zn2+ (25 μM); Figure S13: Fluorescence emission spectra of thiadiazole 1 (100 μM) and Al3+ ion (25 μM) in methanol–water (9:1) solvent (λex = 370 nm) after the addition of EDTA solution; Figure S14: Fluorescence intensity of 1 at 560 nm by the alternate addition of Zn2+ and EDTA; Figure S15: FTIR spectra of compound 1 and 1 + Zn2+; Figure S16: 1H-NMR titration of compound 1 with Zn2+ in DMSO-d6. Table S1: Comparison of reported fluorescent probes derived from salicylaldehyde derivatives for Al3+ in this work.

Author Contributions

Conceptualization, J.H.-M.; methodology, J.H.-M.; formal analysis, A.F.-A. and J.H.-M.; investigation, A.F.-A. and J.H.-M.; writing—original draft preparation, A.C.-V. and J.H.-M.; writing—review and editing, A.C.-V. and J.H.-M.; supervision, J.H.-M.; project administration, J.H.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
NMRNuclear magnetic resonance
TMSTetramethylsilane
PETPhoto-induced electron transfer
RIRRestricts intramolecular rotations
CHEFChelation-enhanced fluorescence
ESIPTExcited-state intramolecular proton transfer
IRInfrared spectra

References

  1. Chan, W.C.; Ng, M.P.; Ang, C.W.; Sim, K.S.; Tan, K.W. From Lab to Life: Safe and Efficient Optical Based Dual-Mode Chemosensor for the Detection of Aluminium (III) and Copper (II) Ions. Inorganica Chim. Acta 2023, 557, 121703. [Google Scholar] [CrossRef]
  2. Sultana, T.; Mahato, M.; Tohora, N.; Ahamed, S.; Das, S.K. An Azine-Based Chromogenic, Fluorogenic Probe for Specific Cascade Detection of Al3+ and PO43− Ions. J. Photochem. Photobiol. A Chem. 2023, 444, 114951. [Google Scholar] [CrossRef]
  3. Hwang, G.W.; Jeon, J.; Neupane, L.N.; Lee, K.-H. Sensitive Ratiometric Detection of Al(III) Ions in a 100% Aqueous Buffered Solution Using a Fluorescent Probe Based on a Peptide Receptor. New J. Chem. 2018, 42, 1437–1445. [Google Scholar] [CrossRef]
  4. Flaten, T.P. Aluminium as a Risk Factor in Alzheimer’s Disease, with Emphasis on Drinking Water. Brain Res. Bull. 2001, 55, 187–196. [Google Scholar] [CrossRef] [PubMed]
  5. Inan-Eroglu, E.; Ayaz, A. Is Aluminum Exposure a Risk Factor for Neurological Disorders? J. Res. Med. Sci. 2018, 23, 51. [Google Scholar] [CrossRef] [PubMed]
  6. Bortoli, P.M.; Alves, C.; Costa, E.; Vanin, A.P.; Sofiatti, J.R.; Siqueira, D.P.; Dallago, R.M.; Treichel, H.; Vargas, G.D.L.P.; Kaizer, R.R. Ilex Paraguariensis: Potential Antioxidant on Aluminium Toxicity, in an Experimental Model of Alzheimer’s Disease. J. Inorg. Biochem. 2018, 181, 104–110. [Google Scholar] [CrossRef]
  7. Roszak, J.; Domeradzka-Gajda, K.; Smok-Pieniążek, A.; Kozajda, A.; Spryszyńska, S.; Grobelny, J.; Tomaszewska, E.; Ranoszek-Soliwoda, K.; Cieślak, M.; Puchowicz, D.; et al. Genotoxic Effects in Transformed and Non-Transformed Human Breast Cell Lines after Exposure to Silver Nanoparticles in Combination with Aluminium Chloride, Butylparaben or Di-n-Butylphthalate. Toxicol. Vitr. 2017, 45, 181–193. [Google Scholar] [CrossRef]
  8. Perl, D.P.; Brody, A.R. Alzheimer’s Disease: X-Ray Spectrometric Evidence of Aluminum Accumulation in Neurofibrillary Tangle-Bearing Neurons. Science 1980, 208, 297–299. [Google Scholar] [CrossRef]
  9. World Health Organization. Evaluation of Certain Food Additives and Contaminants. In WHO Technical Report Series; WHO Press: Geneve, Switzerland, 2011; Volume 966, pp. 7–18. [Google Scholar]
  10. Raju, L.; Deviga, G.; Mariappan, M.; Rajkumar, E. Zinc in Industry. In Zinc; Sukumar, E., Vinothkumar, K., Manickavasagan, A., Eds.; CRC Press: Boca Raton, FL, USA, 2024; p. 18. [Google Scholar]
  11. Patil, R.; Sontakke, T.; Biradar, A.; Nalage, D. Zinc: An Essential Trace Element for Human Health and Beyond. Food Health 2023, 5, 13. [Google Scholar] [CrossRef]
  12. Schoofs, H.; Schmit, J.; Rink, L. Zinc Toxicity: Understanding the Limits. Molecules 2024, 29, 3130. [Google Scholar] [CrossRef]
  13. Aydin, D. A Novel Turn on Fluorescent Probe for the Determination of Al3+ and Zn2+ Ions and Its Cells Applications. Talanta 2020, 210, 120615. [Google Scholar] [CrossRef] [PubMed]
  14. Carpenter, M.C.; Lo, M.N.; Palmer, A.E. Techniques for Measuring Cellular Zinc. Arch. Biochem. Biophys. 2016, 611, 20–29. [Google Scholar] [CrossRef]
  15. Sen, S.; Mukherjee, T.; Chattopadhyay, B.; Moirangthem, A.; Basu, A.; Marek, J.; Chattopadhyay, P. A Water Soluble Al3+ Selective Colorimetric and Fluorescent Turn-on Chemosensor and Its Application in Living Cell Imaging. Analyst 2012, 137, 3975. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, L.; Yang, J.; Wang, H.; Ran, C.; Su, Y.; Zhao, L. A Highly Selective Turn-on Fluorescent Probe for the Detection of Aluminum and Its Application to Bio-Imaging. Sensors 2019, 19, 2423. [Google Scholar] [CrossRef]
  17. Balamurugan, G.; Velmathi, S.; Thirumalaivasan, N.; Wu, S.P. New Phenazine Based AIE Probes for Selective Detection of Aluminium(III) Ions in Presence of Other Trivalent Metal Ions in Living Cells. Analyst 2017, 142, 4721–4726. [Google Scholar] [CrossRef] [PubMed]
  18. Leuci, R.; Brunetti, L.; Laghezza, A.; Loiodice, F.; Tortorella, P.; Piemontese, L. Importance of Biometals as Targets in Medicinal Chemistry: An Overview about the Role of Zinc (II) Chelating Agents. Appl. Sci. 2020, 10, 4118. [Google Scholar] [CrossRef]
  19. Wu, Y.; Peng, X.; Guo, B.; Fan, J.; Zhang, Z.; Wang, J.; Cui, A.; Gao, Y. Boron Dipyrromethene Fluorophore Based Fluorescence Sensor for the Selective Imaging of Zn(II) in Living Cells. Org. Biomol. Chem. 2005, 3, 1387. [Google Scholar] [CrossRef]
  20. Liu, M.; Zhu, H.; Fang, Y.; Liu, C.; Li, X.; Zhang, X.; Ma, L.; Wang, K.; Yu, M.; Sheng, W.; et al. An Ultra-Sensitive Fluorescent Probe for Recognition of Aluminum Ions and Its Application in Environment, Food, and Living Organisms. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2024, 307, 123578. [Google Scholar] [CrossRef]
  21. Shi, Y.; Zhang, W.; Xue, Y.; Zhang, J. Fluorescent Sensors for Detecting and Imaging Metal Ions in Biological Systems: Recent Advances and Future Perspectives. Chemosensors 2023, 11, 226. [Google Scholar] [CrossRef]
  22. Roy, N.; Pramanik, H.A.R.; Paul, P.C.; Singh, T.S. A Highly Sensitive and Selective Fluorescent Chemosensor for Detection of Zn2+ Based on a Schiff Base. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 140, 150–155. [Google Scholar] [CrossRef]
  23. Liu, Q.; Liu, Y.; Xing, Z.; Huang, Y.; Ling, L.; Mo, X. A Novel Dual-Function Probe for Fluorescent Turn-on Recognition and Differentiation of Al3+ and Ga3+ and Its Application. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 287, 122076. [Google Scholar] [CrossRef] [PubMed]
  24. Tang, Y.; Sun, J.; Yin, B. A Dual-Response Fluorescent Probe for Zn2+ and Al3+ Detection in Aqueous Media: PH-Dependent Selectivity and Practical Application. Anal. Chim. Acta 2016, 942, 104–111. [Google Scholar] [CrossRef]
  25. Wen, J.; Hua, Q.; Ding, S.; Sun, A.; Xia, Y. Recent Advances in Fluorescent Probes for Zinc Ions Based on Various Response Mechanisms. Crit. Rev. Anal. Chem. 2024, 54, 3313–3344. [Google Scholar] [CrossRef]
  26. Chhikara, A.; Tomar, D.; Bartwal, G.; Chaurasia, M.; Sharma, A.; Gopal, S.; Chandra, S. Thiadiazole Functionalized Salicylaldehyde-Schiff Base as a PH-Responsive and Chemo-Reversible “Turn-Off” Fluorescent Probe for Selective Cu (II) Detection: Logic Gate Behaviour and Molecular Docking Studies. J. Fluoresc. 2023, 33, 25–41. [Google Scholar] [CrossRef] [PubMed]
  27. Aktara, M.N.; Das, S.; Nayim, S.; Sahoo, N.K.; Beg, M.; Jana, G.C.; Maji, A.; Jha, P.K.; Hossain, M. A Sensorial Colorimetric Detection Method for Hg2+ and Cu2+ Ions Using Single Probe Sensor Based on 5-Methyl-1,3,4-Thiadiazole-2-Thiol Stabilized Gold Nanoparticles and Its Application in Real Water Sample Analysis. Microchem. J. 2019, 147, 1163–1172. [Google Scholar] [CrossRef]
  28. Manna, A.K.; Chowdhury, S.; Patra, G.K. Combined Experimental and Theoretical Studies on a Phenyl Thiadiazole-Based Novel Turn-on Fluorescent Colorimetric Schiff Base Chemosensor for the Selective and Sensitive Detection of Al3+. New J. Chem. 2020, 44, 10819–10832. [Google Scholar] [CrossRef]
  29. Kaur, P.; Kaur, S.; Mahajan, A.; Singh, K. Highly Selective Colorimetric Sensor for Zn2+ Based on Hetarylazo Derivative. Inorg. Chem. Commun. 2008, 11, 626–629. [Google Scholar] [CrossRef]
  30. Mullick, P.; Khan, S.A.; Verma, S.; Alam, O. Thiadiazole Derivatives as Potential Anticonvulsant Agents. Bull. Korean Chem. Soc. 2011, 32, 1011–1016. [Google Scholar] [CrossRef]
  31. Muglu, H.; Vurdu, C.D.; Sayiner, G.; Cavus, M.S.; Kandemirli, F.; Ahmedzade, M. Synthesis and Theoretical Study of 5-Phenyl-1,3,4-Thiadiazole Derivatives. J. Mater. Environ. Sci. 2015, 6, 184–190. [Google Scholar]
  32. Zhang, Z.; Liu, C.; Lu, Y.; Zhao, W.; Zhu, Q.; He, H.; Chen, Z.; Wu, W. In Vivo Fluorescence Imaging of Nanocarriers in Near-Infrared Window II Based on Aggregation-Caused Quenching. J. Nanobiotechnol. 2024, 22, 488. [Google Scholar] [CrossRef]
  33. Leng, X.; Xu, W.; Qiao, C.; Jia, X.; Long, Y.; Yang, B. New Rhodamine B-Based Chromo-Fluorogenic Probes for Highly Selective Detection of Aluminium(III) Ions and Their Application in Living Cell Imaging. RSC Adv. 2019, 9, 6027–6034. [Google Scholar] [CrossRef]
  34. Li, Y.; Wu, J.; Jin, X.; Wang, J.; Han, S.; Wu, W.; Xu, J.; Liu, W.; Yao, X.; Tang, Y. A Bimodal Multianalyte Simple Molecule Chemosensor for Mg2+, Zn2+, and Co2+. Dalt. Trans. 2014, 43, 1881–1887. [Google Scholar] [CrossRef]
  35. Sing Lai, C.; Tiekink, E.R.T. Crystallographic Report: (2,9-Dimethyl-1,10-phenanthroline)Bis-(N,N-pyrrolidinedithiocarbamato)Zinc(II) Chloroform Hemihydrate. Appl. Organomet. Chem. 2003, 17, 255–256. [Google Scholar] [CrossRef]
  36. Caruso, U.; Panunzi, B.; Roviello, A.; Tingoli, M.; Tuzi, A. Two Aminobenzothiazole Derivatives for Pd(II) and Zn(II) Coordination. Inorg. Chem. Commun. 2011, 14, 46–48. [Google Scholar] [CrossRef]
  37. Ayers, P.W.; Parr, R.G.; Pearson, R.G. Elucidating the Hard/Soft Acid/Base Principle: A Perspective Based on Half-Reactions. J. Chem. Phys. 2006, 124, 194107. [Google Scholar] [CrossRef]
  38. Damu, K.V.; Shaikjee, M.S.; Michael, J.P.; Howard, A.S.; Hancock, R.D. Control of Metal Ion Selectivity in Ligands Containing Neutral Oxygen and Pyridyl Groups. Inorg. Chem. 1986, 25, 3879–3883. [Google Scholar] [CrossRef]
  39. Liu, C.; An, X.; Cui, Y.; Xie, K.; Dong, W. Novel Structurally Characterized Hetero-bimetallic [Zn(II) 2 M(II)] (M = Ca and Sr) Bis (Salamo)-type Complexes: DFT Calculation, Hirshfeld Analyses, Antimicrobial and Fluorescent Properties. Appl. Organomet. Chem. 2020, 34, e5272. [Google Scholar] [CrossRef]
  40. Majumder, I.; Chakraborty, P.; Álvarez, R.; Gonzalez-Diaz, M.; Peláez, R.; Ellahioui, Y.; Bauza, A.; Frontera, A.; Zangrando, E.; Gómez-Ruiz, S.; et al. Bioactive Heterometallic Cu II –Zn II Complexes with Potential Biomedical Applications. ACS Omega 2018, 3, 13343–13353. [Google Scholar] [CrossRef] [PubMed]
  41. Li, H.-Q.; Yang, S.-H.; Li, Y.; Ye, W.-X.; Liao, Z.-Y.; Lu, J.-Q.; Wang, Z.-Y. Schiff Base Compounds Derived from 5-Methyl Salicylaldehyde as Turn-On Fluorescent Probes for Al3+ Detection: Experimental and DFT Calculations. Molecules 2025, 30, 1128. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, Y.; Ma, Y.; Li, L.; Sang, W.; Feng, S.; Wang, Y.; Zhang, C.; Yang, S.; Xu, L.; Lu, W. A Benzothiazole-Modified Quinoline Schiff Base Fluorescent Probe for Selective Detection of Zn2+ Ions, DFT Studies and Its Application in Live Cell Imaging. New J. Chem. 2025, 49, 2192–2200. [Google Scholar] [CrossRef]
  43. Musikavanhu, B.; Liang, Y.; Xue, Z.; Feng, L.; Zhao, L. Strategies for Improving Selectivity and Sensitivity of Schiff Base Fluorescent Chemosensors for Toxic and Heavy Metals. Molecules 2023, 28, 6960. [Google Scholar] [CrossRef] [PubMed]
  44. World Health Organization. Guidelines for Drinking-Water Quality: Fourth Edition Incorporating the First and Second Addenda, 4th ed.; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
Scheme 1. Synthesis of ligand 1 from 2-amino-5-phenyl-1,3,4-thiadiazole (2).
Scheme 1. Synthesis of ligand 1 from 2-amino-5-phenyl-1,3,4-thiadiazole (2).
Chemistry 07 00128 sch001
Figure 1. UV–vis absorbance spectra of thiadiazole 1 (100 μM) in the presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 μM) in methanol solvent.
Figure 1. UV–vis absorbance spectra of thiadiazole 1 (100 μM) in the presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 μM) in methanol solvent.
Chemistry 07 00128 g001
Figure 2. Fluorescence emission spectra of thiadiazole 1 (A) λex = 320 nm and (B) λex = 370 nm.
Figure 2. Fluorescence emission spectra of thiadiazole 1 (A) λex = 320 nm and (B) λex = 370 nm.
Chemistry 07 00128 g002
Figure 3. Fluorescence emission spectra of thiadiazole 1 (100 μM) in the presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 μM) in methanol solvent: (A) λex = 320 nm; (B) λex = 370 nm.
Figure 3. Fluorescence emission spectra of thiadiazole 1 (100 μM) in the presence of different metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, Zn2+, and Al3+) (25 μM) in methanol solvent: (A) λex = 320 nm; (B) λex = 370 nm.
Chemistry 07 00128 g003
Figure 4. Fluorescence emission spectra of thiadiazole 1 (100 μM) upon the addition of Al3+ ion (0–200 μM) in methanol solvent (λex = 370 nm, λem = 480 nm) (A) and Zn2+ ion (0–1000 μM) in methanol solvent (λex = 320 nm, λem = 560 nm) (B). Plots of fluorescence of 1 at 480 nm with Al3+ ions (0–100 μM) (C) and at 560 nm with Zn2+ ions (0–1000 μM) (D).
Figure 4. Fluorescence emission spectra of thiadiazole 1 (100 μM) upon the addition of Al3+ ion (0–200 μM) in methanol solvent (λex = 370 nm, λem = 480 nm) (A) and Zn2+ ion (0–1000 μM) in methanol solvent (λex = 320 nm, λem = 560 nm) (B). Plots of fluorescence of 1 at 480 nm with Al3+ ions (0–100 μM) (C) and at 560 nm with Zn2+ ions (0–1000 μM) (D).
Chemistry 07 00128 g004aChemistry 07 00128 g004b
Figure 5. Fluorescence emission spectra of thiadiazole 1 (100 μM) and the Al3+ ion (50 μM) in methanol–water (9:1) solvent (λex = 370 nm) after the addition of EDTA solution (A). Fluorescence intensity of 1 at 480 nm by the alternate addition of Al3+ and EDTA (B).
Figure 5. Fluorescence emission spectra of thiadiazole 1 (100 μM) and the Al3+ ion (50 μM) in methanol–water (9:1) solvent (λex = 370 nm) after the addition of EDTA solution (A). Fluorescence intensity of 1 at 480 nm by the alternate addition of Al3+ and EDTA (B).
Chemistry 07 00128 g005
Figure 6. Response time spectra of fluorescent chemosensor 1 to Al3+ in methanol and methanol–water mixtures (9:1).
Figure 6. Response time spectra of fluorescent chemosensor 1 to Al3+ in methanol and methanol–water mixtures (9:1).
Chemistry 07 00128 g006
Figure 7. (A) Job’s plot of thiadiazole 1 and Al3+ (the total concentration is 100 μM) (λex = 370 nm, λem = 480 nm). (B) Benesi–Hildebrand plot (fluorescence at 480 nm) of thiadiazole 1, assuming 1:1 binding stoichiometry with Al3+. (C) Job’s plot of thiadiazole 1 and Zn2+ (the total concentration is 100 μM) (λex = 320 nm, λem = 560 nm). (D) Benesi–Hildebrand plot (fluorescence at 560 nm) of Zn2+, assuming 2:1 binding stoichiometry with thiadiazole 1.
Figure 7. (A) Job’s plot of thiadiazole 1 and Al3+ (the total concentration is 100 μM) (λex = 370 nm, λem = 480 nm). (B) Benesi–Hildebrand plot (fluorescence at 480 nm) of thiadiazole 1, assuming 1:1 binding stoichiometry with Al3+. (C) Job’s plot of thiadiazole 1 and Zn2+ (the total concentration is 100 μM) (λex = 320 nm, λem = 560 nm). (D) Benesi–Hildebrand plot (fluorescence at 560 nm) of Zn2+, assuming 2:1 binding stoichiometry with thiadiazole 1.
Chemistry 07 00128 g007
Figure 8. Partial hydrogen spectra of compound 1 and 1 + Al3+.
Figure 8. Partial hydrogen spectra of compound 1 and 1 + Al3+.
Chemistry 07 00128 g008
Figure 9. FTIR spectra of compound 1 and 1 + Al3+.
Figure 9. FTIR spectra of compound 1 and 1 + Al3+.
Chemistry 07 00128 g009
Figure 10. Selectivity of the thiadiazole 1 (50 µM) toward 1 eq. of Al3+ and other metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, and Zn2+). In these experiments, the fluorescence measurement was taken at λex = 370 nm in methanol at room temperature. (A) Emission spectra of thiadiazole 1 upon the addition of 1.0 eq. of various metal ions in the absence and presence of 1.0 eq. of the Al3+ ion. (B) Bars represent the final (I480 nm) emission intensity. The blue bars represent the emission after the addition of various metal ions (1 eq.) to a solution of 1. The green bars represent the emission after the addition of 1 eq. of Al3+ to the above solution.
Figure 10. Selectivity of the thiadiazole 1 (50 µM) toward 1 eq. of Al3+ and other metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, and Zn2+). In these experiments, the fluorescence measurement was taken at λex = 370 nm in methanol at room temperature. (A) Emission spectra of thiadiazole 1 upon the addition of 1.0 eq. of various metal ions in the absence and presence of 1.0 eq. of the Al3+ ion. (B) Bars represent the final (I480 nm) emission intensity. The blue bars represent the emission after the addition of various metal ions (1 eq.) to a solution of 1. The green bars represent the emission after the addition of 1 eq. of Al3+ to the above solution.
Chemistry 07 00128 g010
Figure 11. Selectivity of the thiadiazole 1 (50 µM) toward 1 eq. of Zn2+ and other metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, and Al3+). In these experiments, the fluorescence measurement was taken at λex = 320 nm in methanol at room temperature. (A) Emission spectra of thiadiazole 1 upon the addition of 1.0 eq. of various metal ions in the absence and presence of 1.0 eq. of Zn2+ ion. (B) Bars represent the final (I560 nm) emission intensity. The blue bars represent the emission after the addition of various metal ions (1 eq.) to a solution of 1. The red bars represent the emission after the addition of 1 eq. of Zn2+ to the above solution.
Figure 11. Selectivity of the thiadiazole 1 (50 µM) toward 1 eq. of Zn2+ and other metal ions (Cu2+, Co2+, Mn2+, Fe3+, Ni2+, Ca2+, Sn2+, Pb2+, Fe2+, Hg2+, Ag+, and Al3+). In these experiments, the fluorescence measurement was taken at λex = 320 nm in methanol at room temperature. (A) Emission spectra of thiadiazole 1 upon the addition of 1.0 eq. of various metal ions in the absence and presence of 1.0 eq. of Zn2+ ion. (B) Bars represent the final (I560 nm) emission intensity. The blue bars represent the emission after the addition of various metal ions (1 eq.) to a solution of 1. The red bars represent the emission after the addition of 1 eq. of Zn2+ to the above solution.
Chemistry 07 00128 g011
Scheme 2. The proposed mechanism of the sensing of Zn2+ and Al3+.
Scheme 2. The proposed mechanism of the sensing of Zn2+ and Al3+.
Chemistry 07 00128 sch002
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

Heredia-Moya, J.; Fiallos-Ayala, A.; Cevallos-Vallejo, A. Phenylthiadiazole-Based Schiff Base Fluorescent Chemosensor for the Detection of Al3+ and Zn2+ Ions. Chemistry 2025, 7, 128. https://doi.org/10.3390/chemistry7040128

AMA Style

Heredia-Moya J, Fiallos-Ayala A, Cevallos-Vallejo A. Phenylthiadiazole-Based Schiff Base Fluorescent Chemosensor for the Detection of Al3+ and Zn2+ Ions. Chemistry. 2025; 7(4):128. https://doi.org/10.3390/chemistry7040128

Chicago/Turabian Style

Heredia-Moya, Jorge, Ariana Fiallos-Ayala, and Amanda Cevallos-Vallejo. 2025. "Phenylthiadiazole-Based Schiff Base Fluorescent Chemosensor for the Detection of Al3+ and Zn2+ Ions" Chemistry 7, no. 4: 128. https://doi.org/10.3390/chemistry7040128

APA Style

Heredia-Moya, J., Fiallos-Ayala, A., & Cevallos-Vallejo, A. (2025). Phenylthiadiazole-Based Schiff Base Fluorescent Chemosensor for the Detection of Al3+ and Zn2+ Ions. Chemistry, 7(4), 128. https://doi.org/10.3390/chemistry7040128

Article Metrics

Back to TopTop