Next Article in Journal
Curvature-Insensitive Transparent Surface-Enhanced Raman Scattering Substrate Based on Large-Area Ag Nanoparticle-Coated Wrinkled Polystyrene/Polydimethylsiloxane Film for Reliable In Situ Detection
Next Article in Special Issue
Cs-Doped WO3 with Enhanced Conduction Band for Efficient Photocatalytic Oxygen Evolution Reaction Driven by Long-Wavelength Visible Light
Previous Article in Journal
N-Hydroxypiridinedione: A Privileged Heterocycle for Targeting the HBV RNase H
Previous Article in Special Issue
AgIn5S8/g-C3N4 Composite Photocatalyst Coupled with Low-Temperature Plasma-Enhanced Degradation of Hydroxypropyl-Guar-Simulated Oilfield Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Zn(II) Coordination Polymer for Fluorescent Turn-Off Selective Sensing of Heavy Metal Cation and Toxic Inorganic Anions

School of Petrochemical Engineering, Liaoning Petrochemical University, Fushun 113001, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(12), 2943; https://doi.org/10.3390/molecules29122943
Submission received: 28 May 2024 / Revised: 15 June 2024 / Accepted: 18 June 2024 / Published: 20 June 2024
(This article belongs to the Special Issue Chemical Research on Photosensitive Materials)

Abstract

:
A novel coordination polymer [Zn(atyha)2]n (1) (Hatyha = 2-(2-aminothiazole-4-yl)-2- hydroxyiminoacetic acid) was constructed by hydrothermal reaction of Zn2+ with Hatyha ligand. CP 1 exhibits a 2D (4,4)-connected topological framework with Schläfli symbol of {44·62}, where atyha anions serve as tridentate ligands, bridging with Zn2+ through carboxylate, thiazole and oxime groups. CP 1 displays a strong ligand-based photoluminescence at 390 nm in the solid state, and remains significantly structurally stable in water. Interestingly, it can be utilized as a fluorescent probe for selective and sensitive sensing of Fe3+, Cr2O72− and MnO4 through the fluorescent turn-off effect with limit of detection (LOD) of 3.66 × 10−6, 2.38 × 10−5 and 2.94 × 10−6 M, respectively. Moreover, the efficient recyclability for detection of Fe3+ and Cr2O72− is better than that for MnO4. The mechanisms of fluorescent quenching involve reversible overlap of UV-Vis absorption bands of the analytes (Fe3+, Cr2O72− and MnO4) with fluorescence excitation and emission bands for CP 1, respectively.

Graphical Abstract

1. Introduction

Nowadays, coordination polymers (CPs), as one class of promising functional materials, have received widespread attention not only due to unique topologies, such as multiple active sites, modifiable channels and fascinating architectures, but also due to excellent performance and potential applications in various fields, such as catalyst carrier, proton conducting, sensing, magnetic and fluorescent materials, gas separation and storage [1,2,3,4,5,6,7,8]. In particular, the structural diversity and multi-functional characteristics of CPs aroused intense interest in the design and construction of chemical sensing materials [9,10,11,12]. It has been well documented that due to their fast response, high sensitivity and naked eye monitoring, fluorescent CPs can be one type of good chemical sensor for detecting analytes through fluorescence quenching, enhancement or shift [13,14,15,16]. Reasonable selection of metal nodes and organic linkers is of great significance for the construction of fluorescent CPs [17,18,19]. It was found that fluorescent CPs constructed by heterocyclic carboxylate ligands and d10 metal centers show excellent photoluminescence performance [20,21].
Recently, the rapid development of industry has led to inevitable increases in the discharge of hazardous species into the water environment, including heavy metal ions, inorganic oxo-anions and other industrial pollutants. Improper treating and failure to detect these pollutants in time may pose a threat to the ecosystem and cause various human diseases [22]. Fe3+ plays an important role in metabolic processes of the human body and is one of the indispensable substances in the uptake of oxygen through blood in the body and in the formation of DNA and RNA. Excess or deficiency of Fe3+ in the body can result in serious diseases, such as hereditary hemochromatosis and endotoxemia [23]. Inorganic oxo-anions, especially Cr2O72− and MnO4, have been considered as highly toxic pollutants originating from industrial wastewater, and may cause skin allergies or even induce cancers [24]. Therefore, it is imperative to exploit fluorescent CPs to efficiently detect these hazardous species in aqueous media.
In this work, Hatyha ligand was selected to construct a CP based on three structural characteristics: (a) it contains Lewis basic N sites, which can coordinate with Zn(II); (b) it features carboxylate and oxime groups that may adopt various coordination modes to construct complicated topology; (c) it has N, O donors and acceptors, which can form hydrogen bonds and generate high-dimensional supramolecular structures. We have carried out the coordination reaction of atyha ligand with Zn(II) under hydrothermal condition to successfully synthesize a fluorescent CP [Zn(atyha)2]n (1). CP 1 features an infinite 2D-layered structure, displaying strong luminescence emission in solid state at room temperature. It was found that the fluorescence intensity of CP 1 could be markedly quenched by Fe3+, Cr2O72− and MnO4 in the presence of interfering ions, respectively. CP 1 can behave as a chemical sensor based on the fluorescence turn-off effect, with the characteristics of selective and sensitive detection. Furthermore, it also has a relatively low LODs of 3.66 × 10−6, 2.38 × 10−5 and 2.94 × 10−6 M for Fe3+, Cr2O72− and MnO4, respectively.

2. Results and Discussion

2.1. Crystal Structure of CP 1

Single crystal X-ray diffraction analysis reveals that CP 1 crystallizes in the orthorhombic unit cell with space group Pbcn (Table 1). The asymmetric unit consists of one Zn center and two atyha ligands (Figure 1a), with the chemical formula of [Zn(atyha)2]n. Each Zn atom is six-coordinated by two thiazole N atoms and two oxime N atoms from two atyha ligands and two carboxylate O atoms of two atyha ligands, showing a distorted ZnN4O2 octahedral geometry (Figure 1b). The Zn-N distances are between 2.169(2) Å and 2.050(2) Å, the Zn-O lengths are 2.2504(17) Å, the O-Zn-N and N-Zn-N bond angles are in the ranges of 83.84(8)°–162.22(8)° and 76.20(9)°–169.18(13)° (Table 2), respectively, which are similar to the values reported in the references for other Zn complexes [25]. The atyha ligand adopts the carboxylate group to bind to Zn center in a monodentate coordination fashion, as well as thiazole N atom and oxime N atom chelation with Zn center, while the amino group remains uncoordinated. The atyha anions are employed as tridentate ligands to link with adjacent Zn centers, generating a 2D-layered architecture through carboxylate, thiazole and oxime groups (Figure 1c). The hydrogen bonds formed between the amino N atom and the carboxylate O atom (N1-H1B···O2ii) can expand the layered structure into a 3D supramolecular structure (Figure 1d). The hydrogen bonds are formed between oxime O atoms and carboxylate O atoms (O3-H3···O2), and between amino N atoms and carboxylate O atoms (N1-H1A···O1i and N1-H1A···O2) (Table S1), which make the structure more stable.
From a topological perspective, each atyha ligand coordinates with two Zn centers, which can be considered as a bridging ligand-based node (Figure 1e). Each Zn center is surrounded by four atyha ligands, which can be regarded as a 4-c node (Figure 1f). Therefore, CP 1 features a 4, 4-c network structure with a Schläfli symbol of {44·62} (Figure 1g) [26].

2.2. TG Analysis

The TG analysis was performed to investigate the thermal stability of CP 1 (Figure S1). The framework of CP 1 can remain stable before the temperature reaches 196 °C. As the temperature continues to rise, the framework structure of CP 1 undergoes structural collapse and thermal decomposition at 290 °C, resulting in abrupt weight loss, and then it tends to slow down, due to the disintegration of atyha ligand. At 900 °C, CP 1 has not fully decomposed, and the TGA curve still follows a downward trend. The residue may be a mixture of ZnO and ZnS.

2.3. PXRD of CP 1

The powder X-ray diffraction (PXRD) measurement was carried out to confirm the bulk phase purity of CP 1 (Figure S2). The measured pattern of CP 1 was in agreement with the simulated one generated from single crystal X-ray diffraction, revealing that the obtained bulk samples were pure phase. The sample powder of CP 1 was immersed in deionized water for 12 hours and 7 days, respectively, and after centrifugation and natural drying, the PXRD analyses were performed (Figure S3). The results revealed that the experimental patterns of the soaked samples were consistent with the simulated one, manifesting that the framework of CP 1 was intact and possessed high stability in water.

2.4. Fluorescence Spectrum

The solid-state luminescent spectra of CP 1 and free Hatyha ligand were investigated at room temperature (Figure 2). The free Hatyha ligand demonstrates the maximum emission centered at 382 nm when excited at 282 nm. The band may be originated from ππ* and/or nπ* transitions [27]. The luminescent band of CP 1 displays maximum emission at 390 nm (λex = 292 nm). Considering the d10 electron configuration of Zn(II), it is difficult to be oxidized or reduced. Therefore, the emission may result from neither ligand-to-metal charge transfer nor metal-to-ligand charge transfer, but the fluorescence emission of CP 1 can be attributed to intraligand ππ* and/or nπ* charge transfer [27]. In addition, the emission band of CP 1 indicates red-shift of 8 nm in comparison with those of free Hatyha ligand. This perturbation may result from the coordination interactions of atyha ligands with central metal ions.

2.5. Selective Sensing of Fe3+

Considering the excellent fluorescent property and water stability of CP 1, it can be utilized as a fluorescent sensor. From the perspective of environmental protection, water can be used as the medium; moreover, the aqueous suspension of CP 1 shows strong fluorescence intensity. Therefore, the powder sample of CP 1 was ultrasonically dispersed in water as a blank sample to investigate its fluorescent sensing behaviors toward different metal ions. Upon addition of MClx solutions (M = Na+, Mg2+, Ba2+, Sr2+, Mn2+, Zn2+, Ca2+, K+, Cd2+, Pb2+, Al3+, Cr3+, Co2+, Cu2+, Ni2+, Fe3+) to the aqueous suspensions of CP 1 (0.5 × 10−5 M) with metal ion concentration of 0.01 M in the mixture, the emission spectra of the suspensions were measured at the excitation wavelength of 334 nm (Figure 3a). In comparation with the blank sample, Na+ can cause fluorescence enhancement of CP 1; nevertheless, the addition of other metal ions can lead to significant decreases in the fluorescence intensities of CP 1, especially Fe3+ with quenching efficiency of 99.5%, indicating that CP 1 possesses efficient fluorescent turn-off sensing of Fe3+ (Figure 3b).
In order to evaluate the anti-interference ability of CP 1 for detecting Fe3+, we explored the selective detection ability of CP 1 towards Fe3+ with competitive experiments (Figure 4), and further demonstrated that CP 1 can serve as a fluorescent turn-off sensor for detecting Fe3+. The fluorescence response of CP 1 towards Fe3+ was investigated in the presence of interfering metal ions. A 1.5 mL Fe3+ solution (0.01 M) was slowly dripped into a suspension of the powder sample of CP 1 with 0.01 M interfering metal ions (Na+, Mg2+, Ba2+, Sr2+, Mn2+, Zn2+, Ca2+, K+, Cd2+, Pb2+, Al3+, Cr3+, Co2+, Cu2+, and Ni2+), respectively. The fluorescence intensities of CP 1 decreased with a significant fluorescent turn-off effect after adding Fe3+. The measured result indicates that CP 1 can selectively sense Fe3+ without interference from other metal ions.
In order to further evaluate the detection sensitivity of CP 1 toward Fe3+ in detail, a titration experiment of quantitative fluorescence quenching was performed at the excitation wavelength of 334 nm (Figure 5a). With the addition of Fe3+, the fluorescence quenching efficiency sequentially increased. The relationship between the concentration of Fe3+ and the fluorescence intensity of the suspension of powder sample of CP 1 can be analyzed with the Stern–Volmer equation: I0/I = Ksv[M] + 1 in the range of low concentration, where I0 and I are fluorescence intensities of the suspension of powder sample of CP 1 before and after the addition of Fe3+, respectively, Ksv is the slope of the linear curve (quenching coefficient) and [M] is the molar concentration of Fe3+ (Figure 5b). The result indicates that the relationship conforms to the linear equation of I0/I = 0.123[Fe3+] + 0.966 with the linear correlation (R2) of 0.990 and Ksv of 1.23 × 104 M−1. Meanwhile, the LOD of CP 1 toward Fe3+ was further calculated to be 3.66 × 10−6 M by using the equation of LOD = 3σ/Ksv, where σ is the standard deviation. We performed 11 consecutive measurements on the blank sample of CP 1 to obtain 11 fluorescence intensity values, and then calculated the σ value to be 0.015 [28].
As the recyclability of CP 1 for sensing of Fe3+ can increase its potential application, recycling experiments were performed (Figure 6). After the first quenching induced by Fe3+, it was regenerated by centrifugation and washing several times with deionized water, and its fluorescence quenching effect towards Fe3+ was examined again. The result revealed that the fluorescence intensity and quenching efficiency of CP 1 remained almost unchanged through at least three cycles of use, which indicates that CP 1 can be reused for detecting Fe3+ in water.

2.6. Mechanism of Fluorescence Response to Fe3+

In order to investigate the mechanism of fluorescence quenching of CP 1 induced by Fe3+, further investigations were conducted. After immersion in a solution of Fe3+ for 12 h, XRD measurement of the sample of CP 1 was performed (Figure 7a). A similar comparison of PXRD patterns revealed that the crystalline framework of CP 1 was not destroyed. The substitution of the central metal ion of CP 1 with the added Fe3+ will take a long time; however, the fluorescence of CP 1 is quenched relatively quickly by Fe3+. Therefore, ion substitution is not the main reason for fluorescence quenching. The IR spectrum of the sample of CP 1 after immersion in Fe3+ solution essentially matched that of the pristine sample of CP 1, indicating the non-coordination of N, O donors of functional groups in CP 1 with metal ions added (Figure S4). There was an overlap between the fluorescence emission band of CP 1ex = 292 nm) and the UV-Vis absorption spectrum of Fe3+ (Figure 7b), which indicates that fluorescence resonance energy transfer occurs in the sensing process. In addition, the UV-Vis absorption spectrum of Fe3+ presented large overlap with the excitation band of CP 1em = 390 nm) (Figure 7b), which indicated that the competitive absorption might be the main cause of fluorescence quenching of CP 1 by Fe3+. In summary, based on the above results, the plausible mechanism of the quenching phenomenon can be attributed to the combined effect of resonance energy transfer and competitive energy absorption [28].

2.7. Selective Sensing of Cr2O72− and MnO4

Industrial wastewater is usually composed of coexisting metal cations and inorganic anions, and thus, its analysis poses multiple challenges. Sensing measurements were further performed at the excitation wavelength of 334 nm to explore the fluorescence responses of CP 1 to different inorganic anions (Figure 8a). A powder sample of CP 1 was ultrasonically dispersed in 0.01 M aqueous solutions of various potassium salts KMx (M = NO3, CO32−, PO43−, H2PO4, F, Cl, SO42−, Br, Cr2O72− and MnO4), generating a suspension solution, respectively. It was found that inorganic anions induced different fluorescence responses on CP 1 (Figure 8b), especially Cr2O72− and MnO4, which could cause obvious fluorescent turn-off effects with quenching efficiencies of 99.4% and 99.5%, respectively. The results indicate that Cr2O72− and MnO4 can be detected by CP 1 in aqueous solution. In addition, interference experiments were performed to detect Cr2O72− and MnO4 in the presence of other inorganic anions, respectively (Figure 9). The fluorescence quenching intensity of CP 1 towards Cr2O72− and MnO4 remained constant as the interfering anion was changed, respectively. It was evident that the fluorescent turn-off effects of Cr2O72− and MnO4 on CP 1 were almost unaffected by the interfering inorganic anions, respectively.
A titration experiment showed that the fluorescence intensity of CP 1 was gradually quenched at the excitation wavelength of 334 nm with the addition of Cr2O72− and MnO4, respectively (Figure 10 and Figure 11). Moreover, the Ksv curves of Cr2O72− and MnO4 showed good linear correlation, while the Ksv values for Cr2O72− and MnO4 were 1.89 × 103 M−1 and 1.53 × 104 M−1, respectively. Meanwhile, the LODs of Cr2O72− and MnO4 were 2.38 × 10−5 and 2.94 × 10−6 M, respectively. The experimental results indicate that CP 1 can sensitively detect Cr2O72− and MnO4 in water.
To evaluate the recyclability of CP 1 as a fluorescent sensor, the fluorescence of CP 1 was repeatedly quenched by Cr2O72− and MnO4, respectively (Figure 12). After each quenching, the powder sample of CP 1 was recovered through centrifugation, water washing and drying. The experimental results indicated that the fluorescent intensity of CP 1 can almost be restored after exposure to Cr2O72− through at least four cycles; furthermore, the fluorescence quenching efficiency also remains unchanged, indicating that CP 1 has good recyclability for detecting Cr2O72−. However, the fluorescent intensity of CP 1 could not be restored for sensing of MnO4, revealing its poor reversibility.

2.8. Mechanism of Fluorescence Response to Cr2O72−/MnO4

In order to elucidate the mechanism of the fluorescence quenching of CP 1 induced by Cr2O72−/MnO4, additional measurements were performed. The PXRD pattern of a sample of CP 1 immersed in the solution of Cr2O72−/MnO4 was consistent with the simulated one (Figure 13a), indicating that the fluorescence quenching was not caused by the collapse of the structure. The IR spectrum of the sample of CP 1 treated by the solution of Cr2O72−/MnO4 essentially matched with that of the pristine sample, indicating that there was no weak interaction between CP 1 and the inorganic anions added (Figure S4). In addition, the UV-Vis absorption spectra of Cr2O72−/MnO4 overlapped the emission band of CP 1ex = 292 nm) (Figure 13b), which indicated that the fluorescence quenching caused by Cr2O72−/MnO4 could be attributed to the resonance energy transfer. Moreover, a partial overlap existed between the absorption spectra of Cr2O72−/MnO4 and the excitation band of CP 1em = 390 nm) (Figure 13b), which hindered the absorption of CP 1 and caused photoluminescence attenuation. Therefore, there is clear evidence for the competitive adsorption between each analyte and CP 1. In summary, the fluorescence quenching mechanism of Cr2O72−/MnO4 on CP 1 is mainly attributed to the resonance energy transfer, as well as the competitive energy absorption [28].

2.9. Comparison with Other, Related Sensors

Zn(II) CPs have been proven to be excellent sensing materials for metal cations and inorganic anions, etc. Some of them can only be utilized for the detection of a single analyte. For instance, a 2D Zn(II) CP [Zn(4-PP)(1,4-BDC)∙(H2O)]n (1,4-PP = 4-(1H-pyrazol-3-yl)pyridine, 1,4-H2BDC = 1,4-benzenedicarboxylic acid) synthesized under hydrothermal conditions exhibits selective and sensitive detection of Fe3+ in water medium [29]. However, the title Zn(II) CP displays multi-functional fluorescence responses towards Fe3+, Cr2O72− and MnO4. The title CP was constructed by atyha anions and Zn cations, and shows excellent sensing performances towards Fe3+, Cr2O72− and MnO4, which are mainly due to the fluorescent turn-off effect. According to relevant reports in the literatures, it can be seen that these are comparable to or even better than those of other Zn(II) CPs [30,31,32,33,34]. Zhang et al. designed and prepared a Zn(II) CP as a dual-responsive luminescent sensor for Fe3+ and MnO4 in water, with LODs of 5.0 × 10−6 M and 8.86 × 10−6 M, respectively [30]. Another Zn-MOF with a 1D chain structure was confirmed as a multi-functional luminescence sensor for the detection of Fe3+ and Cr2O72− in water through fluorescent quenching, and the LODs were 1.172 × 10−5 M and 2.465 × 10−4 M, respectively [31].

3. Experimental Section

3.1. Materials and Methods

All reagents and solvents were of reagent grade and purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. Elemental analysis was performed with a Perkin-Elmer 240CHN analyzer (Perkin-Elmer Corporation, Norwalk CT, USA). FT-infrared spectra were collected by a Magna FT-IR 750 spectrometer (Nicolet, Tokyo, Japan) in the range of 4000 cm−1–400 cm−1 (KBr pellet). Thermogravimetric data were collected on a NETZSCH STA 449C (Netzsch Corporation, Bavaria, Germany) unit between 25 °C and 900 °C with a heating rate of 10 °C·min−1 under a nitrogen atmosphere, with an Al2O3 crucible used to hold the solid sample. Fluorescence excitation and emission spectra were measured by a Perkin-Elmer LS55 fluorescence spectrophotometer (Perkin-Elmer Corporation, Norwalk, CT, USA) using a 75W Xenon arc-lamp and an R928 photomultiplier tube as a detector for solid samples. Emission intensity measurements were carried out using the adapter and holder supplied by the manufacturer of thespectrophotometer. Excitation and emission spectra were corrected for the instrumental response. UV-Vis absorption spectra were measured by a UV-1800 spectrophotometer (ShiMadzu Corporation, Kyoto, Japan). Powder X-ray diffraction (PXRD) patterns were recorded on a Bruker D8 Advanced XRD diffractometer (Bruker, Rheinstetten, Germany) with Cu-Kα monochromator.

3.2. Synthesis of CP 1

A mixture of Hatyha ligand (0.037 g, 0.2 mmol) and ZnCl2 (0.014 g, 0.1 mmol) in deionized water (20 mL), with its pH value adjusted to about 5 with NaOH (0.1 M) solution, was transferred to a 25 mL Teflon-lined autoclave and heated at 120 °C for 48 h. After cooling to room temperature, colorless blocked crystals of 1 were collected with a yield of 58% based on Zn(II). Elemental analysis (%) found: C, 27.46; H, 1.92; N, 19.26. Calcd for C10H8N6O6S2Zn: C, 27.42; H, 1.83; N, 19.19%. IR (KBr, cm−1): 3414, 3269 (νNH2), 3124(νO-H), 1632, 1347(νCOO), 1512, 1424(νC=C, C=N), 1194, 1083, 1028(νC-O), 800(δC-H), 726, 641(νC-S), 535(νM-N) (Figure S4).

3.3. X-ray Crystallography

Single-crystal X-ray diffraction data for CP 1 were recorded using a Bruker D8 VENTURE diffractometer with graphite-monochromatized Mo-Kα radiation (λ = 0.71073 Å) by φ-ω scan mode. The structure was solved by direct methods and refined by full-matrix least-squares techniques using the SHELXS and SHELXL programs, respectively [35,36]. For CP 1, crystallographic data and structure refinement details are summarized in Table 1. Selected bond lengths and bond angles are given in Table 2.

4. Conclusions

In conclusion, a novel 2D Zn-based coordination polymer was hydrothermally synthesized by introducing Hatyha ligand. The atyha linkers adopted the carboxylate, thiazole and oxime groups to bridge with Zn2+, generating a 2D framework with Schläfli symbol {44·62} topology. Notably, CP 1 displays strong fluorescence emission, which is derived from intraligand transition. It exhibits high structural stabilities towards aqueous solutions. It was found that CP 1 could not only detect Fe3+ and Cr2O72− with high selectivity, sensitivity and recyclability, but also serve as an excellent candidate for the selective and sensitive sensing of MnO4, indicating low LODs of 3.66 × 10−6, 2.38 × 10−5 and 2.94 × 10−6 M, respectively. The mechanism of the fluorescence turn-off sensing can be attributed to the synergistic effect of resonance energy transfer and competitive energy absorption.

Supplementary Materials

The following supporting information can be downloaded at: https://mdpi.longhoe.net/article/10.3390/molecules29122943/s1, Figure S1: TGA curve of CP 1; Figure S2: Measured and simulated PXRD patterns of CP 1; Figure S3: PXRD patterns of CP 1 soaked in water for 12 h and 7 d; Figure S4: IR spectra of sample of CP 1, as well as after immersion in Fe3+, Cr2O72− and MnO4 solutions, Table S1: Hydrogen bond lengths (Å) and bond angles (°) for CP 1.

Author Contributions

Conceptualization, L.G.; methodology, L.G.; software, M.Z.; validation, Y.L., D.Z., X.H. and Y.G.; formal analysis, M.Z.; investigation, M.Z., L.G. and Y.W.; resources, L.G.; data curation, L.G.; writing—original draft preparation, Y.L. and M.Z.; writing—review and editing, L.G. and Y.W.; supervision, L.G.; project administration, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Basic Scientific Research Project of Universities from Liaoning Provincial Education Department, grant number LJKMZ20220738.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhao, X.Y.; Wu, J.F.; Tian, W. Terbium(III)-based coordination polymer with millimeter-size single crystals and high selectivity and sensitivity for folic acid. CrystEngComm 2023, 25, 945–952. [Google Scholar] [CrossRef]
  2. Liu, L.J.; Zhu, H.; Han, C.; Cui, G.H.; Fu, L.S. Luminescent detecting of Fe3+ and Cr2O72− ions by three ternary 2D coordination polymers. Polyhedron 2021, 198, 115074. [Google Scholar] [CrossRef]
  3. Zhou, H.C.; Long, J.R.; Yaghi, O.M. Introduction to metal-organic frameworks. Chem. Rev. 2012, 112, 673–674. [Google Scholar] [CrossRef] [PubMed]
  4. Barnett, B.R.; Gonzalez, M.I.; Long, J.R. Recent progress towards light hydrocarbon separations using metal-organic frameworks. Trends Chem. 2019, 1, 159–171. [Google Scholar] [CrossRef]
  5. Zhao, Q.; Li, Q.Y.; Li, J. Structures, fluorescence and magnetism of a series of coordination polymers driven by a tricarboxypyridine ligand. CrystEngComm 2022, 24, 6751–6761. [Google Scholar] [CrossRef]
  6. Andac, O.; Gorduk, S.; Yilmaz, H. Synthesis, characterization and H2 adsorption performances of polymeric Co(II) and Ni(II) complexes of pyrazine-2,3-dicarboxylic acid and 1-vinylimidazole. J. Iran. Chem. Soc. 2018, 15, 1699–1708. [Google Scholar] [CrossRef]
  7. Liu, H.J.; Yi, R.; Chen, D.M.; Huang, C.; Zhu, B.X. Self-assembly by tridentate or bidentate ligand: Synthesis and vapor adsorption properties of Cu(II), Zn(II), Hg(II) and Cd(II) complexes derived from a bis(pyridylhydrazone) compound. Molecules 2021, 26, 109. [Google Scholar] [CrossRef] [PubMed]
  8. Kuwamura, N.; Konno, T. Heterometallic coordination polymers as heterogeneous electrocatalysts. Inorg. Chem. Front. 2021, 8, 2634–2649. [Google Scholar] [CrossRef]
  9. Sun, Y.Q.; Cheng, Y.; Yin, X.B. Dual-ligand lanthanide metal-organic framework for sensitive ratiometric fluorescence detection of hypochlorous acid. Anal. Chem. 2021, 93, 3559–3566. [Google Scholar] [CrossRef]
  10. Zhang, D.C.; Li, X. A Zn(II) complex with large channels based on 3′-nitro-biphenyl-3,5,4′-tricarboxylic acid: Synthesis, crystal structure, fluorescence sensing of ATP, ADP, GTP, and UTP in aqueous solution and drug delivery. CrystEngComm 2017, 19, 6673–6680. [Google Scholar] [CrossRef]
  11. Wen, M.Y.; Ren, L.; Cui, G.H. Two Co(II) complexes containing pyridylbenzimidazole ligands as chemosensors for the sensing of levofloxacin, acetylacetone, and Ni2+ with high selectivity and sensitivity. CrystEngComm 2021, 23, 8563–8571. [Google Scholar] [CrossRef]
  12. Xu, N.; Zhang, Q.H.; Zhang, G.A. A carbazole-functionalized metal-organic framework for efficient detection of antibiotics, pesticides and nitroaromatic compounds. Dalton Trans. 2019, 48, 2683–2691. [Google Scholar] [CrossRef] [PubMed]
  13. Arici, M. Multifunctional luminescent coordination polymers based on tricarboxylic acid for the detection of 2,4-dinitrophenol and iron(III) and aluminum(III) ions. New J. Chem. 2019, 43, 3690–3697. [Google Scholar] [CrossRef]
  14. Goshisht, M.K.; Tripathi, N. Fluorescence-based sensors as an emerging tool for anion detection: Mechanism, sensory materials and applications. J. Mater. Chem. C 2021, 9, 9820–9850. [Google Scholar] [CrossRef]
  15. Jana, A.K.; Natarajan, S. Fluorescent metal-organic frameworks for selective sensing of toxic cations (Tl3+, Hg2+) and highly oxidizing anions [(CrO4)2−, (Cr2O7)2−, (MnO4)]. ChemPlusChem 2017, 82, 1153–1163. [Google Scholar] [CrossRef] [PubMed]
  16. Rao, P.C.; Mandal, S. Europium-based metal-organic framework as a dual luminescence sensor for the selective detection of the phosphate anion and Fe3+ ion in aqueous media. Inorg. Chem. 2018, 57, 11855–11858. [Google Scholar]
  17. Alexandrov, E.V.; Shevchenko, A.P.; Nekrasova, N.A.; Blatov, V.A. Topological methods for analysis and design of coordination polymers. Russ. Chem. Rev. 2022, 91, RCR5032. [Google Scholar] [CrossRef]
  18. Sokolov, A.V.; Vologzhanina, A.V.; Sudakova, T.V.; Popova, Y.V.; Alexandrov, E.V. Design and synthesis of coordination polymers with Cu(II) and heterocyclic N-oxides. CrystEngComm 2022, 24, 2505–2515. [Google Scholar] [CrossRef]
  19. Tran, M.; Kline, K.; Qin, Y.; Shen, Y.X.; Green, M.D.; Tongay, S. 2D coordination polymers: Design guidelines and materials perspective. Appl. Phys. Rev. 2019, 6, 041311. [Google Scholar] [CrossRef]
  20. Fan, Y.R.; Li, H.B.; Ji, Z.Y.; Liu, J.Y.; Wu, M.Y. Syntheses, structures and photoluminescence of three Zn(II) coordination polymers based on N-containing heterocyclic ligand and varied auxiliary ligands. Inorg. Chem. Commun. 2019, 102, 229–232. [Google Scholar] [CrossRef]
  21. Li, J.D.; Bai, C.; Hu, H.M.; Yang, Z.H.; Xue, G.L. Solvothermal syntheses, crystal structures and luminescence properties of Zn(II) coordination compounds based on imidazophenanthroline carboxylate derivative ligand. J. Solid State Chem. 2019, 227, 1–8. [Google Scholar] [CrossRef]
  22. Wang, Y.N.; Xu, H.; Wang, S.D.; Chang, X.P.; Wang, Y.T.; Qiu, Q.C.; Bai, J.T.; Mo, Y.; Feng, W.Y.; Zhang, M.H.; et al. Efficient antibiotics, small organic molecules and inorganic anions detection with a fluorescent Ni(II) coordination polymer based multiple sensor system. J. Mol. Struct. 2024, 1296, 136829. [Google Scholar] [CrossRef]
  23. Saini, K.; Singh, J.; Malik, S.; Saharan, Y.; Goyat, R.; Umar, A.; Akbar, S.; Ibrahim, A.A.; Baskoutas, S. Metal-Organic Frameworks: A promising solution for efficient removal of heavy metal ions and organic pollutants from industrial wastewater. J. Mol. Liquids 2024, 399, 124365. [Google Scholar] [CrossRef]
  24. Devi, M.K.; Yaashikaa, P.R.; Kumar, P.S.; Oviyapriya, M.; Varshika, V.; Rangasamy, G. Recent advances in carbon-based nanomaterials for the treatment of toxic inorganic pollutants in wastewater. New J. Chem. 2023, 47, 7655–7667. [Google Scholar] [CrossRef]
  25. Agarwal, R.A. One Dimensional Coordination Polymer of Zn(II) for Develo** Multifunctional Nanoparticles. Sci. Rep. 2017, 7, 13212. [Google Scholar] [CrossRef] [PubMed]
  26. Blatov, V.A.; Shevchenko, A.P.; Proserpio, D.M. Applied topological analysis of crystal structures with the program package ToposPro. Cryst. Growth Des. 2014, 14, 3576–3586. [Google Scholar] [CrossRef]
  27. Cao, H.Y.; Liu, Q.Y.; Li, L.Q.; Wang, Y.L.; Chen, L.L.; Yao, Y. Two cadmium coordination compounds with 5-sulfonyl-1,2,4- benzenetricarboxylate ligand: Syntheses, structures, and photoluminescence. Z. Anorg. Allg. Chem. 2014, 640, 1420–1425. [Google Scholar] [CrossRef]
  28. Zhu, C.Y.; Wang, C.L.; Chen, L.; Gao, W.; Li, P.; Zhang, X.M. A water-stable Zn(II) coordination polymer for a high sensitivity detection of Fe3+ and 2,4,6-trinitrophenol. J. Solid State Chem. 2022, 310, 123079. [Google Scholar] [CrossRef]
  29. Liu, N.; **ng, G.E.; Huang, X.X.; Guo, J. A new Zn(II) coordination polymer constructed from 4-(1H-pyrazol-3-yl)pyridine as fluorescent sensor for Fe3+. Chin. J. Struct. Chem. 2019, 38, 660–666. [Google Scholar]
  30. Zhang, X.Q.; Chen, F.M.; Wen, Q.; Zhou, C.C.; He, X.; Li, Y.; Liu, H.F. Zn-based coordination polymers with tricarboxylic acid ligand: Fluorescence sensor toward Fe3+ and MnO4. J. Struct. Chem. 2022, 1252, 132183. [Google Scholar] [CrossRef]
  31. Wang, Y.N.; Ma, Y.L.; Zhang, S.S.; Li, S.F.; Du, L.; Zhao, Q.H. A multifunctional Zn-based coordination polymer showing luminescence detection toward multiple pollutants in water. Inorg. Chem. Commun. 2021, 126, 108476. [Google Scholar] [CrossRef]
  32. Wang, Y.; Liu, Y.; Zhang, X.S.; Luan, J.; Yang, A.A.; Li, W.Z. Effect of secondary bis-pyridine-bis-amide ligand on the construction of Zn-based coordination polymers and the enhancement of ultrasensitive luminescent sensing properties. J. Solid State Chem. 2022, 315, 123516. [Google Scholar] [CrossRef]
  33. Liu, G.C.; Han, S.W.; Gao, Y.; Xu, N.; Wang, X.L.; Chen, B.K. Multifunctional fluorescence responses of phenylamide-bridged d10 coordination polymers structurally regulated by dicarboxylates and metal ions. CrystEngComm 2020, 22, 7952–7961. [Google Scholar] [CrossRef]
  34. Chai, Y.H.; Liu, X.Y.; Cui, Z.Y.; Zhao, Y.; Ma, L.F.; Zhao, B.T. Design and syntheses of two luminescent metal-organic frameworks for detecting nitro-antibiotic, Fe3+ and Cr2O72−. J. Solid State Chem. 2022, 312, 123211. [Google Scholar] [CrossRef]
  35. Sheldrick, G.M. SHELXT-integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  36. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
Figure 1. (a) A drawing showing the coordination environment about Zn2+; (b) octahedral coordination configuration of Zn2+; (c) 2D-layered structure; (d) 3D supramolecular structure; (e) bridged atyha ligand-based node; (f) 4-c node of Zn2+; (g) topological structure.
Figure 1. (a) A drawing showing the coordination environment about Zn2+; (b) octahedral coordination configuration of Zn2+; (c) 2D-layered structure; (d) 3D supramolecular structure; (e) bridged atyha ligand-based node; (f) 4-c node of Zn2+; (g) topological structure.
Molecules 29 02943 g001aMolecules 29 02943 g001b
Figure 2. Fluorescence excitation and emission spectra of (a) free Hatyha ligand and (b) CP 1.
Figure 2. Fluorescence excitation and emission spectra of (a) free Hatyha ligand and (b) CP 1.
Molecules 29 02943 g002
Figure 3. (a) Fluorescence spectra of CP 1 and (b) fluorescence intensities in different solutions of metal ions.
Figure 3. (a) Fluorescence spectra of CP 1 and (b) fluorescence intensities in different solutions of metal ions.
Molecules 29 02943 g003
Figure 4. Fluorescence intensities of CP 1 in solutions with different interfering metal ions before and after addition of Fe3+.
Figure 4. Fluorescence intensities of CP 1 in solutions with different interfering metal ions before and after addition of Fe3+.
Molecules 29 02943 g004
Figure 5. (a) Fluorescence responses of CP 1 in solutions with different concentrations of Fe3+ and (b) Stern–Volmer plot.
Figure 5. (a) Fluorescence responses of CP 1 in solutions with different concentrations of Fe3+ and (b) Stern–Volmer plot.
Molecules 29 02943 g005
Figure 6. Recyclability of CP 1 for sensing of Fe3+.
Figure 6. Recyclability of CP 1 for sensing of Fe3+.
Molecules 29 02943 g006
Figure 7. (a) PXRD pattern of CP 1 soaked in Fe3+ solution; (b) UV-Vis absorption spectrum of Fe3+, fluorescence excitation and emission spectra of CP 1.
Figure 7. (a) PXRD pattern of CP 1 soaked in Fe3+ solution; (b) UV-Vis absorption spectrum of Fe3+, fluorescence excitation and emission spectra of CP 1.
Molecules 29 02943 g007
Figure 8. (a) Fluorescence spectra and (b) fluorescence intensities of CP 1 in different anionic solutions.
Figure 8. (a) Fluorescence spectra and (b) fluorescence intensities of CP 1 in different anionic solutions.
Molecules 29 02943 g008
Figure 9. Fluorescence intensities of CP 1 in solutions with different interfering anions before and after addition of (a) Cr2O72− and (b) MnO4.
Figure 9. Fluorescence intensities of CP 1 in solutions with different interfering anions before and after addition of (a) Cr2O72− and (b) MnO4.
Molecules 29 02943 g009
Figure 10. (a) Fluorescence responses of CP 1 in solutions with different concentrations of Cr2O72− and (b) Stern–Volmer plot.
Figure 10. (a) Fluorescence responses of CP 1 in solutions with different concentrations of Cr2O72− and (b) Stern–Volmer plot.
Molecules 29 02943 g010
Figure 11. (a) Fluorescence responses of CP 1 in solutions with different concentrations of MnO4 and (b) Stern–Volmer plot.
Figure 11. (a) Fluorescence responses of CP 1 in solutions with different concentrations of MnO4 and (b) Stern–Volmer plot.
Molecules 29 02943 g011
Figure 12. Cyclic experiments using CP 1 for detection of (a) Cr2O72− and (b) MnO4.
Figure 12. Cyclic experiments using CP 1 for detection of (a) Cr2O72− and (b) MnO4.
Molecules 29 02943 g012
Figure 13. (a) PXRD patterns of CP 1 soaked in Cr2O72− and MnO4 solution, respectively; (b) UV-Vis absorption spectra of Cr2O72− and MnO4, fluorescence excitation and emission spectra of CP 1.
Figure 13. (a) PXRD patterns of CP 1 soaked in Cr2O72− and MnO4 solution, respectively; (b) UV-Vis absorption spectra of Cr2O72− and MnO4, fluorescence excitation and emission spectra of CP 1.
Molecules 29 02943 g013
Table 1. Crystallographic data and structure refinement details for CP 1.
Table 1. Crystallographic data and structure refinement details for CP 1.
1
Chemical formulaC10H8N6O6S2Zn
Mr437.71
Crystal system,Orthorhombic
Space groupPbcn
Temperature(K)253
a, b, c (Å)11.6451(16), 7.2951(8), 16.761(2)
Z4
V3)1423.9(3)
µ (mm−1)2.07
No. of measured, independent and obverted [I > 2σ(I)] reflection32,248, 1631, 1500
Rint0.057
(sinθ/λ)max−1)0.650
R[F2 > 2σ(F2)], ωR(F2), S0.035, 0.098, 1.08
No. of reflections1631
No. of parameters115
No. of restraints1
R1, ωR2 [I ≥ 2σ(I)]0.0349, 0.0956
R1, ωR2 (all data)0.0378, 0.0978
Δρmax, Δρmin (e·Å−3)1.63, −1.37
Table 2. Selected bond lengths (Å) and angles (°) for CP 1.
Table 2. Selected bond lengths (Å) and angles (°) for CP 1.
Atom1-Atom2DistanceAtom1-Atom2DistanceAtom1-Atom2Distance
Zn1-O1i2.2504(17)Zn1-N22.050(2)Zn1-N32.169(2)
Zn1-O1ii2.2504(17)Zn1-N2iii2.050(2)Zn1-N3iii2.169(2)
Atom1-Atom2-Atom3AngleAtom1-Atom2-Atom3AngleAtom1-Atom2-Atom3Angle
O1i-Zn1-O1ii79.32(7)N2iii-Zn1-N2169.18(13)N3iii-Zn1-O1ii162.22(8)
N2-Zn1-O1ii90.24(8)N2-Zn1-N376.20(9)N3-Zn1-O1i162.22(8)
N2iii-Zn1-O1i90.24(8)N2iii-Zn1-N397.77(9)N3-Zn1-O1ii83.84(8)
N2iii-Zn1-O1ii98.10(8)N2iii-Zn1-N3iii76.20(9)N3iii-Zn1-O1i83.84(8)
N2-Zn1-O1i98.10(8)N2-Zn1-N3iii97.77(9)N3iii-Zn1-N3113.44(14)
Symmetry codes: (i) x+1/2, y+1/2, −z+3/2; (ii) − x+1/2, y+1/2, z; (iii) −x+1, y, −z+3/2.
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

Li, Y.; Zhang, M.; Wang, Y.; Guan, L.; Zhao, D.; Hao, X.; Guo, Y. A Zn(II) Coordination Polymer for Fluorescent Turn-Off Selective Sensing of Heavy Metal Cation and Toxic Inorganic Anions. Molecules 2024, 29, 2943. https://doi.org/10.3390/molecules29122943

AMA Style

Li Y, Zhang M, Wang Y, Guan L, Zhao D, Hao X, Guo Y. A Zn(II) Coordination Polymer for Fluorescent Turn-Off Selective Sensing of Heavy Metal Cation and Toxic Inorganic Anions. Molecules. 2024; 29(12):2943. https://doi.org/10.3390/molecules29122943

Chicago/Turabian Style

Li, Yaxin, Mouyi Zhang, Ying Wang, Lei Guan, Di Zhao, **nyu Hao, and Yuting Guo. 2024. "A Zn(II) Coordination Polymer for Fluorescent Turn-Off Selective Sensing of Heavy Metal Cation and Toxic Inorganic Anions" Molecules 29, no. 12: 2943. https://doi.org/10.3390/molecules29122943

Article Metrics

Back to TopTop