Next Article in Journal
The Power of Field-Flow Fractionation in Characterization of Nanoparticles in Drug Delivery
Previous Article in Journal
Screening of Tyrosinase, Xanthine Oxidase, and α-Glucosidase Inhibitors from Polygoni Cuspidati Rhizoma et Radix by Ultrafiltration and HPLC Analysis
Previous Article in Special Issue
Synergism of Photo-Induced Electron Transfer and Aggregation-Induced Quenching Mechanisms for Highly Sensitive Detection of Silver Ion and Captopril
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Highly Selective and Sensitive Sequential Recognition Probe Zn2+ and H2PO4 Based on Chiral Thiourea Schiff Base

1
School of Chemical Engineering and Technology, Hebei University of Technology, Tianjin 300401, China
2
Tianjin Key Laboratory of Structure and Performance for Functional Molecules, College of Chemistry, Tianjin Normal University, Tianjin 300387, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(10), 4166; https://doi.org/10.3390/molecules28104166
Submission received: 5 May 2023 / Revised: 15 May 2023 / Accepted: 16 May 2023 / Published: 18 May 2023
(This article belongs to the Special Issue Design and Synthesis of Novel Fluorescent Molecules)

Abstract

:
A series of novel chiral thiourea fluorescent probes HL1HL6 were designed and synthesized from (1R,2R)-1,2-diphenylethylenediamine, phenyl isothiocyanate, and different substituted salicylic aldehydes. All of the compounds were confirmed by 1H NMR, 13C NMR, and HRMS. They exhibit high selectivity and sensitivity to Zn2+ in the presence of nitrate ions with the detection limit of 2.3 × 10−8 M (HL5). Meanwhile, their zinc (II) complexes (L-ZnNO3) showed continuous response to H2PO4 in acetonitrile solution. The identification processes could further be verified by supramolecular chemistry data analysis, X-ray single-crystal diffraction analysis, and theoretical study. The research provides reliable evidence for an explanation of the mechanism of action of thiourea involved in coordination, which is important for the application of thiourea fluorescent probes. In short, the sensors HL1HL6 based on chiral thiourea Schiff base will be promising detection devices for Zn2+ and H2PO4.

Graphical Abstract

1. Introduction

In recent years, widespread research has been undertaken on the detection of metal ions [1,2,3,4,5] and inorganic anions [6,7,8,9], particularly within the development of selective ion probes, which has more generally become a popular topic in scientific research. Zn2+ is the second most common element in the human body, after iron, and not only plays an important role in protein structure, catalysis, transcription, and regulation, but is also closely associated with nerve signal transmission, enzyme regulation, and gene expression [10]. Excessive concentrations of Zn2+ can cause neurodegenerative diseases such as Alzheimer’s disease and Parkinson’s disease [11,12,13,14]. Therefore, the development of methods to accurately determine trace amounts of Zn2+ and for its removal prior to entering the body are important research goals.
Among the various anions, the detection of phosphates most commonly the subject of research. Phosphates are the key substrates of many biochemical reactions and the main components of biomolecules [15,16,17,18]. Phosphate and its derivatives are important nutrients, and the basic life activities of organisms are closely related to phosphate. Phosphate is not only a component of some important structures of plants but is also a catalyst for many key biochemical reactions in plants, which can promote the development of plant roots and the early growth of seedlings. The life activities of organisms are also closely related to phosphates [18]. Phosphates can regulate the level of 1,25-(OH) 2-vitamin D in human plasma [19] and react with hydroxyl groups to form phosphate esters, thus allowing monomers to polymerize into relatively stable long chain skeletons (DNA and RNA skeletons) [20]. In addition, phosphate is also involved in ATP energy supply, ion channel regulation, enzymatic reactions, and intercellular signal transduction [8]. Unfortunately, a large amount of phosphate deposition can lead to a lack of dissolved oxygen in water and algae eutrophication [18].
At present, there are many conventional methods to detect Zn2+ ions including ion chromatography, atomic absorption spectrometry, and ion exchange [21]. Among these methods, chemical sensors are widely used to detect many biological and environmental heavy metal pollutants because of their advantages such as higher sensitivity, lower cost, and faster detection [22,23]. In particular, the multi-functional sensors used for the sequential recognition of cations and anions have remarkable advantages in special applications for their low cost, quick response, and convenient operation. Most of the currently available fluorescent probes can only analyze certain types of anions or cations, and there are few fluorescent sensors with the sequential recognition of Zn2+ and H2PO4 [24,25].
In this work, a series of chiral thiourea Schiff base fluorescent probes (HL1HL6) were designed and synthesized starting from (1R,2R)-1,2-diphenylethylenediamine, phenyl isothiocyanate, and different substituted salicylic aldehydes (Figure 1). They were explored for the first time for fluorescent-responsive Zn2+ detection in the presence of nitrate ions in acetonitrile. Meanwhile, their zinc (II) complexes (HL-Zn2+-NO3) showed continuous response to H2PO4 in acetonitrile solution. Up to now, the detection mechanisms of thiourea probes have been reported mostly by 1H NMR titration or Job’s plot titration experiments to speculate the possible coordination mode, and some of the assays do not provide practical and reliable data [26,27]. Here, we successfully elucidated the mechanism of metal ion recognition by this kind of thiourea Schiff base fluorescent probe using supramolecular chemistry data analysis, X-ray single-crystal diffraction analysis, and theoretical study. The research provides reliable evidence for an explanation of the mechanism of action of thiourea involved in coordination, which is important for the application of thiourea fluorescent probes. The probes in our work have the advantages of high sensitivity and low detection limits. In short, the sensors HL1HL6 based on chiral thiourea Schiff base will be promising detection devices for Zn2+ and H2PO4.

2. Results and Discussion

2.1. Fluorescence Spectroscopic Studies of HL1HL6 and 3 in the Presence of Metal Ions

2.1.1. Cation Selectivity Experiments

Selectivity experiments were carried out for HL1–HL6 and 3 by examining the emission spectra in the presence of different metal ions (Ag+, Na+, K+, Co2+, Fe3+, Cu2+, Cd2+, Cr3+, Mg2+, Pb2+, Ni2+, Zn2+, Fe2+, Hg2+, Ca2+) in filtered Milli-Q water. As shown in Figure 2a, HL1 displayed a weak emission peak at 447 nm in acetonitrile solution when excited at 380 nm. Upon the addition of various metal ions in HL1, the emission spectrum was only significantly enhanced with Zn2+, whereas Cd2+ and Pb2+ caused a very small enhancement in the emission intensity at wavelengths corresponding to 400 and 600 nm, respectively. This observation demonstrates the high selectivity of HL1 towards Zn2+ compared to other metal ions, including Cd2+ and Pb2+. Molecular probes that show turn-on fluorescence signaling upon interaction with Zn2+ also often interact with Cd2+/Pb2+, resulting in turn-on signaling [28]. In contrast, blue fluorescence was observed in the probe solution upon interaction with Zn2+ under a 365 nm UV lamp (Figure 2a, inset). In the case of other metal cations, the fluorescence spectra almost remained unchanged (Figure 2a). Subsequently, the selectivity of the HL1–HL6 and 3 probes for various metal ions was investigated (Figures S1 and S2). The addition of Zn2+ resulted in a prominent luminescence enhancement (Figures S1 and S2), whereas the addition of a large excess of other competitive cations caused only slight luminescent changes (Figure S2), with the exception of 3 (Figures S1f and S2g). The fluorescence emission produced by the interaction between HL5 and Zn2+ was the strongest, whereas that with HL2 was the weakest. The strength of the other observed emissions was at a medium level, except for compound 3. The difference in the fluorescence intensity of other compounds mainly depends on the difference in the side substituents of salicylaldehyde. The fluorescence intensity of HL5 is better than that of HL3, and obviously stronger than that of HL4, HL6, and HL2. When salicylaldehyde contains electron-donating groups, such as –OH, –CH3, and –OCH3, it can be increased the density of the electron cloud on the benzene ring, resulting in the weakening of the coordination ability of HL with zinc ions. Under the combined influence of the inductive effect and mesomeric effect, HL5 containing –Cl has the strongest coordination ability with Zn2+. Compound 3 without the –OH in the benzene ring structure of the benzylidene group was not able to recognize Zn2+, which demonstrates that the phenolic hydroxyl group was involved in the formation of complexes. HL1HL6 probes showed high selectivity in sensing Zn2+ and can be used as “turn-on” luminescent molecular probes toward Zn2+.

2.1.2. Cation-Competitive Experiments

The competitive experiments between the HL1HL6 probes and coexisting metal ions were investigated on the basis of emission spectra. As shown in Figure 3a, whether in the absence or presence of competitive metal ions, the strong increase in the emission intensity of the HL1 probe was observed upon the addition of Zn2+. Although Cd2+ and Pb2+, among the verified cations, triggered enhancements in emission, these were negligible compared to that of Zn2+. When HL2HL6 were treated with 10 equiv. of other metal ions and 2 equiv. of Zn2+, similar luminescence changes were also observed. The effect was almost the same as that resulting from the addition of Zn2+ (Figure S3). In all, the coexistence of these cations had a negligible effect on the detection of Zn2+ in the tested media.

2.1.3. Study of the Reversibility and Dependence of Fluorescence on Solvent, pH, Different Zinc Salts

To examine whether the complexation of HL and Zn2+ is reversible, we added an aqueous solution of Na2EDTA during detection. Upon the addition of Na2EDTA (1.0 equiv.) to the solution of HL1 containing Zn2+, the emission intensity of the mixture solution decreased significantly with the luminescence quenching.t It seems logical that the fluorescent intensity observed in Figure 4a is inversely proportional to the stability of the complex formed by L-Zn with the anions as the fluorescence of the complex can be quenched by Na2EDTA [29]. Changes in the solvent and pH had non-negligible effects on the detection effect of the fluorescent probe. Therefore, we investigated the dependence of the fluorescence emission intensity on solvent and pH during the recognition of Zn2+ by HL1. Interestingly, the intensity of the fluorescence emission spectrum in the acetonitrile solvent was significantly enhanced, and CH3CH2OH was also a good choice under some conditions (Figure 4b). Further investigation of the effect of pH on the detection of Zn2+ was carried out by adjusting the required pH using 0.01 M HCl or 0.01 M NaOH. The fluorescence emission intensity was stable within the pH range of 5.5–9.0, which includes physiological range basically (Figure 4c). Under strong acid or base conditions, the stability of the C=N bond and the presence of -OH can have a significant impact, resulting in fluorescence quenching. To judge whether the probe can selectively recognize Zn2+ in different zinc salts, an examination of the sensitivity of the L1 probe to several different zinc salts (zinc chloride, zinc gluconate, and zinc nitrate) was undertaken using fluorescence spectroscopy. However, the fluorescence spectrum enhancement of the L1 probe was not obvious when it was complexed with zinc chloride and zinc gluconate, and the fluorescence enhancement phenomenon only appeared when it interacted with zinc nitrate (Figure 4d). Therefore, we preliminarily speculate that the HL probe recognizes Zn2+ in the presence of NO3, which is involved in the coordination process.

2.1.4. Zn2+ Titration Analysis

Fluorescence titration spectra were examined to obtain additional information regarding the binding form of the HL1 probe with Zn2+ (Figure 3b). After adding Zn2+ to the solution, HL1 exhibited a broad emission profile peaking at 447 nm in acetonitrile solution and displayed a linear enhancement in the emission intensity. The increase in the intensity of the emission spectra stopped after the addition of 1 equivalent of metal ions, which suggested 1:1 binding stoichiometry between HL1 and Zn2+ (Figure 3b, inset). In addition, the photoluminescence titration experiments of individual complexes of HL2HL6 with Zn2+ in varying concentrations were carried out by monitoring the emission intensity changes. Similarly, the emission intensity of HL2HL6 increased continuously until the addition of 1 equiv. of Zn2+; the further addition of Zn2+ induced only minor changes in the luminescence spectra (Figure S4).

2.2. Binding Mechanism

According to literature reports [30,31,32], the supramolecular chemical data analysis method was adopted to analyze the combination of host and guest, and the binding constants of host and guest 1:1, 1:2, and 2:1 were simulated, respectively. The results are shown in Table 1. The detection limit of HL1 for Zn2+ was calculated as 3σ/k, where σ is the standard deviation of the blank measurement and k is the slope of the plot of the emission intensity ratio versus Zn2+ concentration [33]. The limit of detection (LOD) of HL2HL6 for Zn2+ are 2.8 × 10−8 M, 2.2 × 10−6 M, 3.3 × 10−8 M, 1.4 × 10−7 M, 2.3 × 10−8 M, and 1.74 × 10−7 M, respectively. The fluorescence intensity of the HL probe had a good linear relationship with the concentration of Zn(NO3)2 and had a lower detection limit, which could be used for the quantitative detection of Zn(NO3)2.
To further confirm the coordination mechanism of ligand HL with zinc nitrate, the crystals were obtained and analyzed. Single crystals of HL4 (CCDC: 2077389) and L4-ZnNO3 (CCDC: 2077390) suitable for X-ray diffraction study were obtained by the slow evaporation of mixed solutions of acetonitrile/methanol at room temperature. L4-ZnNO3 was cultured in proportions of equal amounts and an excessive amount of Zn(NO3)2, and the complex single crystal with a coordination mode of 1:1 was obtained. The ellipsoid diagrams of HL4 and L4-ZnNO3 shown in Figure 5 were obtained by X-ray single crystal diffraction analysis. The corresponding crystallographic data are summarized in Table S1, and selected bond lengths and angles are listed in Table S2. The single crystal of the complex L4-ZnNO3 clearly showed the formation of a six-membered ring and a seven-membered ring. The central Zn atom adopted a four-coordination method, which was, respectively, connected to the S atom (Zn-S = 2.314(16) Å), the N atom of C=N (Zn-N3 = 2.020(5) Å), the O atom of the salicylaldehyde phenolic hydroxyl group (Zn-O1 = 1.946(5) Å), and the O atom of NO3 (Zn-O3 = 2.043(5) Å). This coordination mode is consistent with the coordination mode of monodentate ligands reported in the literature [34,35,36]. The crystal structure indicates that the N atom on the thiourea group was not coordinated, and that only the S atom is involved.
Coordination mechanism studies confirm that HL and Zn(NO3)2 form strongly fluorescent complexes, which suggests 1:1 binding stoichiometry between HL and Zn(NO3)2. Through the single-crystal structure of the complex, it can be visually seen that the atoms coordinated with Zn2+ include O atoms provided by phenolic hydroxyl groups, N atoms provided by Schiff bases, and S atoms provided by thiourea structural units. Zn2+ with HL forms a six-membered ring and a seven-membered ring with coordination atoms.

2.3. Theoretical Study

Density Functional Theory (DFT) calculations and time-dependent DFT (TD-DFT) using B3LYP were carried out with Gaussian 09 software package ADDIN EN.CITE [37]. The 6-31G(d,p) basis set was employed for C, H, N, O, and S atoms, and the LANL2DZ basis set was employed for the Zn atom [38,39]. Full geometry optimizations of the HL4 and L4-ZnNO3 in the singlet ground-state were carried out using the DFT method (Figure 6). The assignment of the type of each MO is made on the basis of its composition. The frontier molecular orbital results from the extension geometries show that the spatial distributions of the HOMO (HOMO = highest occupied molecular orbital) and LUMO (LUMO = lowest unoccupied molecular orbital) are both localized with the phenol moiety with a bandgap of 4.243 eV. However, after the complexation of HL4 with Zn2+, the HOMOs of L4-ZnNO3 were spread over the phenol moiety, and the LUMOs extended to the Zn2+ and thiourea moiety with the bandgap of 3.732 eV, which indicates the possible electron transfer from ligand to metal (LMCT). The theoretical bond length between S and O was 5.12262 Å before adding Zn2+, and was shortened to 4.05971 Å after forming a complex with HL4, which was consistent with the change in the crystal data.

2.4. Fluorescence Studies of L-ZnNO3 in the Presence of Anions

2.4.1. Anion Selectivity Experiments

Different zinc salt selectivity experiments revealed that different kinds of anions could affect the Zn2+ recognition performance. To investigate the effect of different kinds of anions on zinc complexes, anion selectivity experiments were performed. When excited at 377 nm, L1-ZnNO3 displayed a strong emission peak at 449 nm. Among a set of different anions (Cl, Br, F, CH3COO, NO3, C6H5COO, C6H5SO2, PO42−, H2PO4, HPO42−, SO32−, SO42−, H2PO2 , NO2, NO3, and SCN), a prominent change in the emission spectra of L1-ZnNO3 manifested upon the addition of H2PO4 alone (Figure 7). When L1-ZnNO3 was combined with H2PO4, the bright blue color of the solution disappeared, resulting in a colorless transparent solution (Figure 7, insert and Figure S5). A similar phenomenon also occurred when adding H2PO4 to L2-ZnNO3, L3-ZnNO3, L4-ZnNO3, L5-ZnNO3, and L6-ZnNO3 (Figure S6). The above experimental results proved that this type of fluorescent probe could realize the “turn-on–turn-off” behavior, thus realizing the recognition function of cation and anion.

2.4.2. Anion-Competitive Experiments

To verify the influence of other anions on the probe recognition of H2PO4, the emission intensity of the probe and H2PO4 at 450 nm (λex = 370 nm) was monitored (Figure 8a). The result showed that the L1-ZnNO3 could specifically detect H2PO4 in the presence of other related anions. The anion-competitive experiments of L2-ZnNO3, L3-ZnNO3, L4-ZnNO3, L5-ZnNO3, and L6-ZnNO3 are shown in the Supplementary Materials (Figure S7), which were generally consistent with L1-ZnNO3.
The titration of L1-ZnNO3 with H2PO4 triggered a distinct decrease in the emission intensity at 450 nm (Figure 8b). L2-ZnNO3~L6-ZnNO3 and H2PO4 are exhibited in the Supplementary Materials (Figure S8).
In this work, a series of novel chiral thiourea fluorescent probes HL1HL6 were prepared for the first time for fluorescence-responsive Zn2+, while their complex L-Zn-NO3 was continuously identified and detected the phosphate anion (H2PO4). Compared with other working probes for the detection of analytes (Table S4) [7,40,41,42,43,44,45,46], the probes synthesized in this work have the advantages of high sensitivity and low detection limits.

3. Materials and Methods

3.1. General Procedures

3.1.1. Reagents and Instruments

All of the materials used for synthesis were purchased from commercial suppliers and used as received. Nitrate salts of Ag+, Na+, K+, Co2+, Fe3+, Cu2+, Cd2+, Cr3+, Mg2+, Pb2+, Ni2+, and Zn2+, sulfate salts of Fe2+ and Hg2+, and chlorine salt of Ca2+ were used for the spectroscopic studies. The different sodium salts of F, Cl, Br, CH3COO, C6H5COO, C6H5SO2, PO43−, H2PO4, HPO42−, SO32−, SO42−, H2PO2, NO2, NO3, and SCN were used for investigating anion effects. All salts were dissolved in filtered Milli-Q water to prepare aqueous ion solutions.
The melting points of the products were determined on an X-4 binocular microscope. 1H and 13C NMR spectra were recorded on a Bruker 400 MHz instrument at room temperature. Chemical shifts were measured relative to residual solvent peaks of CDCl3 (1H: δ = 7.26 ppm; 13C: δ = 77.0 ppm) or DMSO-d6 (1H: δ = 2.50 ppm; 13C: δ = 39.5 ppm) with tetramethylsilane (TMS) as internal standard. The following abbreviations are used to describe spin multiplicities in 1H NMR spectra: s = singlet; d = doublet; t = triplet; m = multiplet. Specific rotations were measured on a PerkinElmer 341 MC polarimeter. HRMS data were obtained using a 7.0T FT-ICR MS spectrometer. X-ray crystal diffraction measurements of L were acquired using a Bruker SMART APEX II CCD diffractometer. Fluorescence spectra were examined on a Hitachi FL-2700 spectrophotometer using 10 mm path length quartz cuvettes at 298 K. The pH was calculated by STARTER 3100/F.

3.1.2. Synthesis of HL1HL6 and Compound 3

The synthetic route for HL1HL6 and compound 3 is shown in Figure 1. Under ice-cold conditions, (1R,2R)-1,2-diphenylethylenediamine (1, 4.71 mmol) was dissolved in dichloromethane under ice-cold conditions with constant stirring. Isothiocyanate (4.71 mmol) was dissolved in dichloromethane and then added dropwise to the above amine system via a constant pressure dropping funnel. The reaction mixture was stirred for 12 h and then evaporated under low pressure. The crude product was purified by silica gel column chromatography using dichloromethane as an eluent to give the chiral primary amine thiourea 2 in the form of a white powder. Compound 2 and the substituted aldehydes were dissolved in an ethanol solvent and heated to reflux under nitrogen protection for 24 h. The mixture was cooled and washed with ether (3 × 50 mL) to give the solid products HL1HL6 and compound 3.
1-((1R,2R)-2-(((E)-2-Hydroxybenzylidene)amino)-1,2-diphenylethyl)-3-phenylthiourea (HL1). White solid, 70% yield, m.p. 104–108 °C; [ α ] D 23 = +33.44 (c = 10, CH2Cl2); 1H NMR (CDCl3, 400 MHz) δ 12.54–12.23 (m, 1H), 8.06 (s, 1H), 7.69 (s, 1H), 7.48–7.28 (m, 7H), 7.24–6.74 (m, 13H), 5.99 (s, 1H), 4.81 (d, J = 4.1 Hz, 1H); 13C NMR (CDCl3, 101 MHz) δ 180.5, 167.2, 160.6, 138.6, 138.2, 137.4, 135.7, 133.0, 132.0, 130.3, 129.3, 128.5, 128.3, 127.9, 127.6, 127.3, 127.2, 126.3, 125.4, 124.0, 118.9, 118.4, 117.0, 64.2, 37.1, 34.4, 32.7, 31.9, 30.1, 29.7, 29.4, 26.9, 22.7, 14.1, 11.4; HRMS (ESI) m/z calc’d for C28H25N3OS [M + H]+: 452.1797, found 452.1790.
1-((1R,2R)-2-(((E)-2,3-Dihydroxybenzylidene)amino)-1,2-diphenylethyl-3-phenylthiourea (HL2). Yellow solid, 48% yield, m.p. 146–149 °C; [ α ] D 23 = +36.61 (c = 10, CH2Cl2); 1H NMR (DMSO-d6, 400 MHz) δ 13.20 (d, J = 155.7 Hz, 1H), 9.88 (s, 1H), 9.12 (s, 1H), 8.75–8.36 (m, 2H), 7.59–6.54 (m, 18H), 6.11 (t, J = 8.3 Hz, 1H), 5.15–4.87 (m, 1H); 13C NMR (DMSO-d6, 101 MHz) δ 180.4, 166.9, 149.1, 145.5, 140.2, 139.8, 139.4, 128.7, 128.4 (d, J = 10.9 Hz), 128.3, 127.8, 127.6, 127.3, 126.9, 126.7, 123.9, 122.6, 121.9, 118.8, 118.4, 118.3, 76.5, 62.8; HRMS (ESI) m/z calc’d for C28H25N3O2S [M + H]+: 468.1746, found 468.1749.
1-((1R,2R)-2-(((E)-2-Hydroxy-5-methylbenzylidene)amino)-1,2-diphenylethyl)-3-phe-nylthiourea (HL3). White solid, 81% yield, m.p. 139–142 °C; [ α ] D 23 = +42.78 (c = 10, CH2Cl2); 1H NMR (DMSO-d6, 400 MHz) δ 12.45 (s, 1H), 9.75 (s, 1H), 8.48–8.16 (m, 2H), 7.39–7.03 (m, 17H), 6.82 (d, J = 8.3 Hz, 1H), 6.14 (t, J = 7.6 Hz, 1H), 4.93 (d, J = 6.5 Hz, 1H), 2.22 (s, 3H); 13C NMR (DMSO-d6, 101 MHz) δ 178.1, 164.0, 155.3, 138.1, 137.7, 136.8, 131.1, 129.1, 126.2, 126.0, 125.6, 125.3, 125.2, 125.2, 125.0, 124.6, 121.9, 120.6, 116.2, 114.0, 74.3, 60.5, 17.6; HRMS (ESI) m/z calc’d for C29H27N3OS [M + H]+: 466.1953, found 466.1958.
1-((1R,2R)-2-(((E)-2-Hydroxy-3-methoxybenzylidene)amino)-1,2-diphenylethyl)-3-p-henylthiourea (HL4). Brown solid, 84% yield, m.p. 150–152 °C; [ α ] D 23 = +48.81 (c = 10, CH2Cl2); 1H NMR (DMSO-d6, 400 MHz) δ 12.55 (s, 1H), 9.35 (s, 1H), 8.27–7.34 (m, 2H), 7.23–6.53 (m, 17H), 6.44 (d, J = 7.9 Hz, 1H), 5.74 (s, 1H), 4.55 (d, J = 6.6 Hz, 1H), 3.39 (s, 3H); 13C NMR (DMSO-d6, 101 MHz) δ 180.4, 166.6, 150.1, 147.8, 140.2, 139.9, 139.1, 128.5, 128.3, 127.9, 127.6, 127.5, 127.4, 126.9, 124.1, 123.0, 122.8, 118.6, 118.4, 115.0, 76.3, 62.7, 55.8; HRMS (ESI) m/z calc’d for C29H27N3O2S [M + H]+: 482.1902, found 482.1907.
1-((1R,2R)-2-(((E)-5-Chloro-2-hydroxybenzylidene)amino)-1,2-diphenylethyl)-3-phe-nylthiourea (HL5). Yellow solid, 80% yield, m.p. 158–160 °C; [ α ] D 23 = +37.11 (c = 10, CH2Cl2); 1H NMR (DMSO-d6, 400 MHz) δ 12.58 (s, 1H), 9.73 (s, 1H), 8.30 (d, J = 8.0 Hz, 2H), 7.54 (s, 1H), 7.40–7.14 (m, 16H), 7.09 (d, J = 7.2 Hz, 1H), 6.94 (d, J = 8.8 Hz, 1H), 6.14 (s, 1H), 4.93 (d, J = 6.3 Hz, 1H); 13C NMR (DMSO-d6, 101 MHz) δ 180.5, 164.3, 158.4, 140.2, 132.3, 129.9, 128.5, 128.3, 127.9, 127.6, 127.5, 127.4, 127.0, 123.0, 122.4, 120.3, 118.5, 76.5, 62.7; HRMS (ESI) m/z calc’d for C28H24ClN3OS [M + H]+: 482.1407, found 486.1412.
1-((1R,2R)-2-(((E)-(2-Hydroxynaphthalen-1-yl)methylene)amino)-1,2-diphenylethyl)-3-phenylthiourea (HL6). Yellow solid, 74% yield, m.p. 154–156 °C; [ α ] D 23 = +46.22 (c = 10, CH2Cl2); 1H NMR (DMSO-d6, 400 MHz) δ 14.92 (s, 1H), 9.65 (s, 1H), 9.25 (s, 1H), 8.42 (s, 1H), 8.04 (s, 1H), 7.93–7.71 (m, 2H), 7.49 (s, 1H), 7.36–7.06 (m, 15H), 6.96 (d, J = 9.2 Hz, 2H), 6.22 (s, 1H), 5.24 (s, 1H); 13C NMR (DMSO-d6, 101 MHz) δ 180. 4, 161.3, 139.3, 138.9, 135.9, 133.1, 129.0, 128. 5, 128.0, 127. 9, 127.8, 127.6, 127.5, 127.2, 126.3, 122.9, 123.0, 122.2, 119.2, 107.2, 72.7; HRMS (ESI) m/z calc’d for C32H27N3OS [M + H]+: 502.1953, found 502.1959.
1-((1R,2R)-2-(((E)-Benzylidene)amino)-1,2-diphenylethyl)-3-phenylthiourea (3). White solid, 76% yield, m.p. 175–179 °C; [ α ] D 23 = +27.33 (c = 10, CH2Cl2); 1H NMR (CDCl3, 400 MHz) δ 7.99–7.76 (m, 2H), 7.73–7.27 (m, 16H), 7.26–7.02 (m, 5H), 5.92 (s, 1H), 4.60 (s, 1H); 13C NMR (CDCl3, 101 MHz) δ 180.4, 162.9, 141.2, 140.5 (d, J = 65.6 Hz), 135.3, 131.2, 130.2, 128.5, 128.4, 128.3, 127.6, 127.4, 127.2, 126.7, 126.1, 77.3, 64.9; HRMS (ESI) m/z calc’d for C28H25N3S [M + H]+: 436.1847, found 436.1841.

3.2. Spectroscopic Measurements

Stock solutions of various ions (0.50 mol/L) were prepared in filtered Milli-Q water. A stock solution of HL probe (1.0 × 10−2 mol/L) and compound 3 in acetonitrile was freshly prepared for fluorescence measurement. Fluorescence spectroscopic measurements were recorded for samples at 1 min after the addition of various analytes. All spectroscopic measurements were performed at 25 °C. For fluorescence measurements, the excitation wavelength was fixed at 380 nm (λex = 380 nm) and 357 nm for HL1 and HL2, and at 390 nm (λex = 390 nm) for HL3HL6 and compound 3, respectively. The emission wavelength was recorded from 220 to 600 nm for HL1HL6 and compound 3. In the cation selectivity experiments, the test samples were prepared by interacting 20 μL of the cation stock (0.50 mol/L) with 1 mL of HL solution (1.0 × 10−4 mol/L) and compound 3. In the presence of other metal ions, the competition between Zn2+ and other metal ions was systematically studied by fluorescence emission spectroscopy. In the titration experiments, the solution of HL (1.0 × 10−5 mol/L) was placed in a quartz cuvette, and a certain volume of the metal ion stock solution (5.0 × 10−4 mol/L) was added gradually to achieve a concentration of 1.0 × 10−5 mol/L. Similarly, anion experiments were conducted based on L-ZnNO3 to study the fluorescence spectra after combination with different anions.

3.3. Synthesis of the Sensor Zn(II) Complex and X-ray Crystallography

A methanol solution (2 mL) of Zn(NO3)2·6H2O (0.0297 g, 1 equiv.) was added into the round bottom flask containing acetonitrile (3 mL) and compound HL4 (0.0482 g, 0.1 mmol). Then, it was sealed with unsintered polytetrafluoroethylene (PTFE) tapes and holes were pricked in the tape with a needle. The crystals of L4-ZnNO3 were obtained after the solvent slowly evaporated at room temperature.
The crystal structures were determined using a Bruker SMART APEX II CCD diffractometer with a monochromator and Cu Kα radiation (λ = 0.71073 Å) at 114(2) or 146.4(3) K with an increasing ω (width of 0.3° per frame) at a scan speed of 5 s per frame. Using Olex2 [47], the structure was solved with the ShelXT [48] structure solution program using Direct Methods and refined with the ShelXL [49] refinement package using least squares minimization. All non-hydrogen atoms were refined anisotropically and the hydrogen atoms attached to all carbon atoms were geometrically fixed. The positional and temperature factors were refined isotropically.

4. Conclusions

Chiral thiourea Schiff base compounds were synthesized and their binding properties with cations were studied. The spectral data showed that the HL1HL6 probes have high selectivity and sensitivity to Zn2+ and are not affected by the presence of other coexisting metal ions. Combined with the crystal data, it was determined that the HL and Zn(NO3)2 form a L-ZnNO3 complex in a 1:1 ratio. The formed L-ZnNO3 complex can selectively recognize H2PO4 without interference from other anions. The probe has good prospects for application and provides a new approach for the detection of Zn(NO3)2 and H2PO4 in actual samples.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28104166/s1: Section S1: Fluorescence spectra and photos of L with metal ion (Figures S1–S4); Section S2: Fluorescence spectra and photos of L-ZnNO3 with anion (Figures S5–S8); Section S3: Crystallographic data (Tables S1–S3); Section S4: Examples for detection of Zn2+ and phosphate anions by sensors (Table S4); Section S5: 1H NMR and 13C NMR spectra of HL1HL6 and 3 (Figures S9–S22).

Author Contributions

Project administration, supervision, Z.W.; chemical methodology, S.Y.; formal analysis, S.Y. and Y.H.; writing—original draft, S.Y.; writing—review and editing, H.L. and A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the S&T Program of Hebei (21326504D) and the Natural Science Foundation of Hebei Province (B2020202028).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data used to support the findings of this study are included within the article and Supplementary Materials.

Acknowledgments

The authors also acknowledge the State Key Laboratory of Elemento-Organic Chemistry (Nankai University) for the characterization of the target compounds.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds HL1HL6 and 3 are available from the authors.

References

  1. Chowdhury, S.; Rooj, B.; Dutta, A.; Mandal, U. Review on recent advances in metal ions sensing using different fluorescent probes. J. Fluoresc. 2018, 28, 999–1021. [Google Scholar] [CrossRef] [PubMed]
  2. Hou, L.J.; Feng, J.; Wang, Y.B.; Dong, C.; Shuang, S.M.; Wang, Y. Single fluorescein-based probe for selective colorimetric and fluorometric dual sensing of Al3+ and Cu2+. Sens. Actuators B Chem. 2017, 247, 451–460. [Google Scholar] [CrossRef]
  3. Zhan, Y.C.; Tsai, J.J.; Chen, Y.C. Zinc ion-based switch-on fluorescence-sensing probes for the detection of tetracycline. Molecules 2022, 27, 8403. [Google Scholar] [CrossRef] [PubMed]
  4. Asiri, A.M.; Al-Ghamdi, N.S.M.; Dzudzevic-Cancar, H.; Kumar, P.; Khan, S.A. Physicochemical and Photophysical investigation of newly synthesized carbazole containing pyrazoline-benzothiazole as fluorescent chemosensor for the detection of Cu2+, Fe3+ & Fe2+ metal ion. J. Mol. Struct. 2019, 1195, 670–680. [Google Scholar] [CrossRef]
  5. Shi, W.; Chen, Y.B.; Chen, X.; Xie, Z.F.; Hui, Y.H. Simple-structured, hydrazinecarbothioamide derivatived dual-channel optical probe for Hg2+ and Ag+. J. Lumin. 2016, 174, 56–62. [Google Scholar] [CrossRef]
  6. Gale, P.A.; Caltagirone, C. Fluorescent and colorimetric sensors for anionic species. Coord. Chem. Rev. 2018, 354, 2–27. [Google Scholar] [CrossRef]
  7. Banerjee, M.; Ta, S.; Ghosh, M.; Ghosh, A.; Das, D. Sequential fluorescence recognition of molybdenum(VI), arsenite, and phosphate ions in a ratiometric manner: A facile approach for discrimination of AsO2− and H2PO4−. ACS Omega 2019, 4, 10877–10890. [Google Scholar] [CrossRef]
  8. Wagh, Y.B.; Tayade, K.C.; Kuwar, A.; Sahoo, S.K.; Mayank; Singh, N.; Dalal, D.S. Exploration of highly selective fluorogenic “on–off” chemosensor for H2PO4- ions: ICT-based sensing and ATPase activity profiling. Luminescence 2020, 35, 379–384. [Google Scholar] [CrossRef]
  9. Lu, W.; Zhang, M.Y.; Liu, K.Y.; Fan, B.; Xia, Z.; Jiang, L.M. A fluoride-selective colorimetric and fluorescent chemosensor and its use for the design of molecular-scale logic devices. Sens. Actuators B Chem. 2011, 160, 1005–1010. [Google Scholar] [CrossRef]
  10. Hagimori, M.; Hara, F.; Mizuyama, N.; Fujino, T.; Saji, H.; Mukai, T. High-affinity ratiometric fluorescence probe based on 6-amino-2,2′-bipyridine scaffold for endogenous Zn2+ and its application to living cells. Molecules 2022, 27, 1287. [Google Scholar] [CrossRef]
  11. Roy, A.; Shee, U.; Mukherjee, A.; Mandal, S.K.; Roy, P. Rhodamine-based dual chemosensor for Al3+ and Zn2+ ions with distinctly separated excitation and emission wavelengths. ACS Omega 2019, 4, 6864–6875. [Google Scholar] [CrossRef]
  12. Liu, D.D.; Zhang, M.Z.; Du, W.; Hu, L.; Li, F.; Tian, X.H.; Wang, A.D.; Zhang, Q.; Zhang, Z.P.; Wu, J.Y. A series of Zn(II) terpyridine-based nitrate complexes as two-photon fluorescent probe for identifying apoptotic and living cells via subcellular immigration. Inorg. Chem. 2018, 57, 7676–7683. [Google Scholar] [CrossRef] [PubMed]
  13. Winnett, M.R.; Mini, P.; Grace, M.R.; Tuck, K.L. Time-resolved terbium-based probe for the detection of zinc(II) ions: Investigation of the formation of a luminescent ternary complex. Inorg. Chem. 2020, 59, 118–127. [Google Scholar] [CrossRef]
  14. Bourassa, D.; Elitt, C.M.; McCallum, A.M.; Sumalekshmy, S.; McRae, R.L.; Morgan, M.T.; Sigel, N.; Perry, J.W.; Rosenberg, P.A.; Fahrni, C.J. Chromis-1, a ratiometric fluorescent probe optimized for two-photon microscopy reveals dynamic changes in labile Zn(II) in differentiating oligodendrocytes. ACS Sens. 2018, 3, 458–467. [Google Scholar] [CrossRef]
  15. Cho, J.; Verwilst, P.; Kang, M.J.; Pan, J.L.; Sharma, A.; Hong, C.S.; Kim, J.S.; Kim, S. Crown ether-appended calix[2]triazolium[2]arene as a macrocyclic receptor for the recognition of the H2PO4− anion. Chem. Commun. 2020, 56, 1038–1041. [Google Scholar] [CrossRef] [PubMed]
  16. Asthana, S.K.; Pandey, A.; Kumar, A.; Upadhyay, K.K. An incisive optical recognition of monohydrogen phosphate by a fluorescein-based chemodosimeter. New J. Chem. 2020, 44, 2201–2205. [Google Scholar] [CrossRef]
  17. Yao, H.; Zhou, Q.; Wang, J.; Chen, Y.Y.; Kan, X.T.; Wei, T.B.; Zhang, Y.M.; Lin, Q. Highly selective Fe3+ and F-/H2PO4- sensor based on a water-soluble cationic pillar[5]arene with aggregation-induced emission characteristic. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 221, 117215–117220. [Google Scholar] [CrossRef]
  18. Zhang, D.W.; Cochrane, J.R.; Martinez, A.; Gao, G.H. Recent advances in H2PO4− fluorescent sensors. RSC Adv. 2014, 4, 29735–29749. [Google Scholar] [CrossRef]
  19. Gray, R.W.; Wilz, D.R.; Caldas, A.E.; Jacob Lemann, J.R. The importance of phosphate in regulating plasma 1,25-(OH)2-vitamin D levels in humans: Studies in healthy subjects, in calcium-stone formers and in patients with primary hyperparathyroidism. J. Clin. Endocrinol. Metab. 1976, 45, 299–306. [Google Scholar] [CrossRef]
  20. Shi, B.B.; Zhang, Y.M.; Wei, T.B.; Lin, Q.; Yao, H.; Zhang, P.; You, X.M. A fluorescent and colorimetric chemosensor for dihydrogen phosphate ions based on 2-pyridine-1H-imidazo[4,5-b]phenazine–zinc ensemble. Sens. Actuators B Chem. 2014, 190, 555–561. [Google Scholar] [CrossRef]
  21. Liu, F.F.; Fan, C.B.; Pu, S.Z. A new “turn-on” fluorescent chemosensor for Zn2+ based on a diarylethene derivative and its practical applications. J. Photochem. Photobiol. A Chem. 2019, 371, 248–254. [Google Scholar] [CrossRef]
  22. Kumar, S.S.; Kumar, R.S.; Kumar, S.A. An “Off-On-Off” type fluorescent chemosensor for the relay detection of Zn2+ and H2PO4− in aqueous environment. Inorg. Chim. Acta 2020, 502, 119348. [Google Scholar] [CrossRef]
  23. Wang, S.; Ma, L.L.; Liu, G.; Pu, S.Z. Diarylethene-based fluorescent and colorimetric chemosensor for the selective detection of Al3+ and CN−. Dyes Pigments 2019, 164, 257–266. [Google Scholar] [CrossRef]
  24. Liu, Y.; Wang, X.; Feng, E.T.; Fan, C.B.; Pu, S.Z. A highly selective sequential recognition probe for Zn2+ and HSO4−/H2PO4− based on a diarylethene chemosensor. Spectrochim. Acta A 2021, 246, 119052. [Google Scholar] [CrossRef] [PubMed]
  25. Arabahmadi, R. Antipyrine-based Schiff base as fluorogenic chemosensor for recognition of Zn2+, Cu2+ and H2PO4− in aqueous media by comparator, half subtractor and integrated logic circuits. J. Photochem. Photobiol. A Chem. 2022, 426, 113762. [Google Scholar] [CrossRef]
  26. Liu, G.; Zhao, L. A simple colorimetric and on–off fluorescent chemosensor for biologically important anions based on thiourea groups. Spectrosc. Lett. 2012, 45, 424–429. [Google Scholar] [CrossRef]
  27. Chowdhury, B.; Sinha, S.; Ghosh, P. Substitution effect on near infrared absorbance based selective fluoride sensing of indole functionalized thiourea molecules. Eur. J. Org. Chem. 2019, 5, 1008–1015. [Google Scholar] [CrossRef]
  28. Sinha, S.; Dey, G.; Kumar, S.; Mathew, J.; Mukherjee, T.; Mukherjee, S.; Ghosh, S. Cysteamine-based cell-permeable Zn2+-specific molecular bioimaging materials: From animal to plant cells. ACS Appl. Mater. Interfaces 2013, 5, 11730–11740. [Google Scholar] [CrossRef]
  29. Powell, K.J.; Brown, P.L.; Byrne, R.H.; Gajda, T.; Hefter, G.; Leuz, A.-K.; Sjoberg, S.; Wanner, H. Chemical speciation of environmentally significant metals with inorganic ligands. Part 5: The Zn2+ + OH−, Cl−, CO32−, SO42−, and PO43− systems (IUPAC Technical Report). Pure Appl. Chem. 2013, 85, 2249–2311. [Google Scholar] [CrossRef]
  30. Hibberta, D.B.; Thordarson, P. The death of the Job plot, transparency, open science and online tools, uncertainty estimation methods and other developments in supramolecular chemistry data analysis. Chem. Commun. 2016, 52, 12792–12805. [Google Scholar] [CrossRef]
  31. Available online: http://supramolecular.org (accessed on 9 May 2016).
  32. Thordarson, P. Determining association constants from titration experiments in supramolecular chemistry. Chem. Soc. Rev. 2011, 40, 1305–1323. [Google Scholar] [CrossRef] [PubMed]
  33. Datta, B.K.; Thiyagarajan, D.; Ramesh, A.; Das, G. A sole multi-analyte receptor responds with three distinct fluorescence signals: Traffic signal like sensing of Al3+, Zn2+ and F−. Dalton Trans. 2015, 44, 13093–13099. [Google Scholar] [CrossRef]
  34. Valencia, L.; Pérez-Lourido, P.; Bastida, R.; Macias, A. Dinuclear Zn(II) polymer consisting of channels formed by π,π-stacking interactions with a flow of nitrate anions through the channels. Cryst. Growth Des. 2008, 8, 2080–2082. [Google Scholar] [CrossRef]
  35. Dolai, M.; Mistri, T.; Panja, A.; Ali, M. Diversity in supramolecular self-assembly through hydrogen-bonding interactions of non-coordinated aliphatic –OH group in a series of heterodinuclear CuIIM (M=NaI, ZnII, HgII, SmIII, BiIII, PbII and CdII). Inorg. Chim. Acta 2013, 399, 95–104. [Google Scholar] [CrossRef]
  36. Zhu, Y.; Xia, C.K.; Meng, S.C.; Chen, J.; Chen, J.; Xie, J.M. Syntheses, structures, properties and DFT calculations of coordination polymers constructed by 2,6-bis(benzimidazolyl)pyridine. Polyhedron 2013, 61, 181–187. [Google Scholar] [CrossRef]
  37. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. Revision A.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  38. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  39. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results obtained with the correlation energy density functionals of becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200–206. [Google Scholar] [CrossRef]
  40. Paulpandi, R.Q.; Ramasamy, S.; Paulraj, M.S.; Baños, F.G.D.; Villora, G.; Cerón-Carrasco, J.P.; Pérez-Sanchez, H.; Enoch, I.V.M.V. Enhanced Zn2+ ion-sensing behavior of a benzothiazole derivative on encapsulation by β- cyclodextrin. RSC Adv. 2016, 6, 15670–15677. [Google Scholar] [CrossRef]
  41. Dong, Z.; Le, X.; Zhou, P.; Dong, C.; Ma, J. An “off–on–off” fluorescent probe for the sequential detection of Zn2+ and hydrogen sulfide in aqueous solution. New J. Chem. 2014, 38, 1802–1808. [Google Scholar] [CrossRef]
  42. Dong, Z.; Le, X.; Zhou, P.; Dong, C.; Ma, J. Sequential recognition of zinc ion and hydrogen sulfide by a new quinoline derivative with logic gate behavior. RSC Adv. 2014, 4, 18270–18277. [Google Scholar] [CrossRef]
  43. Shyamal, M.; Mazumdar, P.; Maity, S.; Samanta, S.; Sahoo, G.P.; Misra, A. Highly selective turn-on fluorogenic chemosensor for robust quantification of Zn(II) based on aggregation induced emission enhancement feature. ACS Sens. 2016, 1, 739–747. [Google Scholar] [CrossRef]
  44. He, X.; Xie, Q.; Fan, J.; Xu, C.; Xu, W.; Li, Y.; Ding, F.; Deng, H.; Chen, H.; Shen, J. Dual-functional chemosensor with colorimetric/ratiometric response to Cu(II)/Zn(II) ions and its applications in bioimaging and molecular logic gates. Dyes Pigments 2020, 177, 108255. [Google Scholar] [CrossRef]
  45. Moradi, S.; Molavipordanjani, S.; Hosseinimehr, S.J.; Emami, S. Benzo[d]imidazo[2,1-b] thiazole-based fluorescent sensor for Zn2+ ion detection. J. Photochem. Photobiol. A 2020, 389, 112184. [Google Scholar] [CrossRef]
  46. Sethupathi, M.; Jayamani, A.; Muthusankar, G.; Sakthivel, P.; Sekar, K.; Gandhid, S.; Sengottuvelan, N.; Gopu, G.; Selvaraju, C. Colorimetric and fluorescence sensing of Zn2+ ion and its bio-imaging applications based on macrocyclic “tet a” derivative. J. Photochem. Photobiol. B 2020, 207, 11854. [Google Scholar] [CrossRef] [PubMed]
  47. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  48. Sheldrick, G.M. SHELXT-Integrated space-group and crystal-structure determination. Acta Crystallogr. A-Found. Adv. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, C71, 3–8. [Google Scholar] [CrossRef]
Figure 1. Synthesis of HL1HL6 and compound 3.
Figure 1. Synthesis of HL1HL6 and compound 3.
Molecules 28 04166 g001
Figure 2. (a) Change in the emission spectra of HL1 upon the addition of different metal ions in acetonitrile. Inset: Visual color change of HL1 before and after the addition of Zn2+ under a 365 nm UV lamp; (b) fluorescence spectra of HL1HL6 and compound 3 after the addition of Zn2+; (c) color of HL1HL6 and compound 3 solution with the addition of Zn2+ under a UV lamp.
Figure 2. (a) Change in the emission spectra of HL1 upon the addition of different metal ions in acetonitrile. Inset: Visual color change of HL1 before and after the addition of Zn2+ under a 365 nm UV lamp; (b) fluorescence spectra of HL1HL6 and compound 3 after the addition of Zn2+; (c) color of HL1HL6 and compound 3 solution with the addition of Zn2+ under a UV lamp.
Molecules 28 04166 g002
Figure 3. (a) Fluorescence intensities of HL1 excited at 377 nm upon the addition of Zn(NO3)2 in the presence of nitrates subject to interference. The black bars represent the emission intensities of HL1 in the presence of nitrates of interest (10.0 equiv.). The red bars represent the change in emission upon the subsequent addition of Zn(NO3)2 (2.0 equiv.) to the above solution; (b) emission spectra of HL1 in acetonitrile solution in the presence of an increasing amount of Zn2+ in filtered Milli-Q water. Inset: Fluorescence intensity of HL1 depending on the Zn2+ in the range from 0 to 2.0 equiv.
Figure 3. (a) Fluorescence intensities of HL1 excited at 377 nm upon the addition of Zn(NO3)2 in the presence of nitrates subject to interference. The black bars represent the emission intensities of HL1 in the presence of nitrates of interest (10.0 equiv.). The red bars represent the change in emission upon the subsequent addition of Zn(NO3)2 (2.0 equiv.) to the above solution; (b) emission spectra of HL1 in acetonitrile solution in the presence of an increasing amount of Zn2+ in filtered Milli-Q water. Inset: Fluorescence intensity of HL1 depending on the Zn2+ in the range from 0 to 2.0 equiv.
Molecules 28 04166 g003
Figure 4. (a) Zn2+ and EDTA were added to a solution of HL1 in acetonitrile; (b) emission spectra of the HL1 probe in different solvents, excitation at 380 nm; (c) fluorescence spectra of HL1 interacting with Zn2+ at different pH in acetonitrile; (d) effects on fluorescence spectra of HL1 using different zinc salts in acetonitrile.
Figure 4. (a) Zn2+ and EDTA were added to a solution of HL1 in acetonitrile; (b) emission spectra of the HL1 probe in different solvents, excitation at 380 nm; (c) fluorescence spectra of HL1 interacting with Zn2+ at different pH in acetonitrile; (d) effects on fluorescence spectra of HL1 using different zinc salts in acetonitrile.
Molecules 28 04166 g004
Figure 5. Oka Ridge Thermal Ellipsoidal plot (ORTEP) diagrams of HL4 (a) and L4-ZnNO3 (b) with the atom numbering schemes at the 40% probability level. Hydrogen atoms and solvent molecules are omitted for clarity.
Figure 5. Oka Ridge Thermal Ellipsoidal plot (ORTEP) diagrams of HL4 (a) and L4-ZnNO3 (b) with the atom numbering schemes at the 40% probability level. Hydrogen atoms and solvent molecules are omitted for clarity.
Molecules 28 04166 g005
Figure 6. Frontier molecular orbitals of probe HL4 and L4-ZnNO3 as obtained from DFT calculations.
Figure 6. Frontier molecular orbitals of probe HL4 and L4-ZnNO3 as obtained from DFT calculations.
Molecules 28 04166 g006
Figure 7. Fluorescence spectra of L1-Zn2+-NO3 in acetonitrile with different anions dissolved in filtered Milli-Q water.
Figure 7. Fluorescence spectra of L1-Zn2+-NO3 in acetonitrile with different anions dissolved in filtered Milli-Q water.
Molecules 28 04166 g007
Figure 8. (a) Comparison of fluorescence intensity before and after the addition of L1-ZnNO3 in acetonitrile; (b) fluorescence emission spectra of L1-ZnNO3 in acetonitrile solution with the addition of H2PO4. Inset: Fluorescence intensity of L1-ZnNO3 depending on the H2PO4 in the range from 0 to 2.0 equiv.
Figure 8. (a) Comparison of fluorescence intensity before and after the addition of L1-ZnNO3 in acetonitrile; (b) fluorescence emission spectra of L1-ZnNO3 in acetonitrile solution with the addition of H2PO4. Inset: Fluorescence intensity of L1-ZnNO3 depending on the H2PO4 in the range from 0 to 2.0 equiv.
Molecules 28 04166 g008
Table 1. Job plot shapes of HL1 at various concentrations and K1:1, K1:2 to K2:1 ratios. K1:1 = 1000, K1:2 = 100, K2:1 = 100 were assumed.
Table 1. Job plot shapes of HL1 at various concentrations and K1:1, K1:2 to K2:1 ratios. K1:1 = 1000, K1:2 = 100, K2:1 = 100 were assumed.
Parameter (Bounds)OptimizedErrorInitial
K (0 → ∞)8607.30 M−1±9.1006%100.00 M−1
K11 (0 → ∞)3376.76 M−1±5.2092%1000.00 M−1
K12 (0 → ∞)4790.08 M−1±26.9494%100.00 M−1
K11 (0 → ∞)3576.53 M−1±9.1090%1000.00 M−1
K21 (0 → ∞)−277,004.73 M−1±−4.6365%100.00 M−1
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

Yang, S.; Huang, Y.; Lu, A.; Wang, Z.; Li, H. A Highly Selective and Sensitive Sequential Recognition Probe Zn2+ and H2PO4 Based on Chiral Thiourea Schiff Base. Molecules 2023, 28, 4166. https://doi.org/10.3390/molecules28104166

AMA Style

Yang S, Huang Y, Lu A, Wang Z, Li H. A Highly Selective and Sensitive Sequential Recognition Probe Zn2+ and H2PO4 Based on Chiral Thiourea Schiff Base. Molecules. 2023; 28(10):4166. https://doi.org/10.3390/molecules28104166

Chicago/Turabian Style

Yang, Shan, Yichuan Huang, Aidang Lu, Ziwen Wang, and Hongyan Li. 2023. "A Highly Selective and Sensitive Sequential Recognition Probe Zn2+ and H2PO4 Based on Chiral Thiourea Schiff Base" Molecules 28, no. 10: 4166. https://doi.org/10.3390/molecules28104166

Article Metrics

Back to TopTop