Next Article in Journal
Experimental Characterization of Stress- and Strain-Dependent Stiffness in Grouted Rock Masses
Previous Article in Journal
Advances in Materials for Recent Low-Profile Implantable Bioelectronics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Reversible Bis(Salamo)-Based Fluorescence Sensor for Selective Detection of Cd2+ in Water-Containing Systems and Food Samples

School of Chemical and Biological Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(4), 523; https://doi.org/10.3390/ma11040523
Submission received: 27 February 2018 / Revised: 22 March 2018 / Accepted: 28 March 2018 / Published: 29 March 2018

Abstract

:
A novel, simple, highly selective, and sensitive fluorescence chemosensor for detecting Cd2+ that was constructed from a bis(salamo)-type compound (H4L) with two N2O2 chelating moieties as ionophore was successfully developed. Sensor H4L could show fluorescence turn-on response rapidly and significant selectivity to Cd2+ over many other metallic ions (Cu2+, Ba2+, Ca2+, K+, Cr3+, Mn2+, Sr2+, Co2+, Na+, Li+, Ni2+, Ag+, and Zn2+), and a clear change in color from colorless to yellow that can be very easily observed via the naked eyes in the existence of Cd2+, while other metallic ions do not induce such a change. Interestingly, its fluorescent intensity was increased sharply with the increased concentration of Cd2+. The detection limit of sensor H4L towards Cd2+ was down to 8.61 × 10−7 M.

Graphical Abstract

1. Introduction

Metallic ions play a key role in daily life [1,2,3,4]. Cadmium, which is an essential resource in the earth, is widely used in all kinds of agricultural processes and industry containing chemical industry, electronics industry, nuclear industry semiconducting, quantum dots, phosphate fertilizers, and other fields [5]. However, it should be vigilant that Cd2+ is a heavy metallic ion with highly toxic [6]. Some new studies reveal that a considerable amount of cadmium go into the environment from waste water, waste residue, and exhaust gas not only damages the environment, but also endangers human health [7]. Hence, with ever-increasing concern for environment and human health, there is now a much greater demand for the development of a rapid and convenient detection method for Cd2+ ion [8,9].
Up until now, with the development of optical sensors for recognizing heavy and transition metal ions in living organisms [10,11], intense efforts have been devoted to the design and synthesis of high sensitivity fluorescence sensors due to the lower cost and rapid response, as well as the easy operability of the fluorescent technique [12,13,14,15,16,17]. According to the relevant literatures, the metallic coordination compounds with salen-type N2O2 ligands and their corresponding analogues could be used to catalysis [18], nonlinear optical materials and magnetic materials [19,20,21,22,23,24,25,26,27,28], supramolecular architecture [29,30], ions recognition [31,32,33,34,35,36,37,38,39,40], and biological fields, and so forth [41,42,43,44,45,46,47]. Today, researches on the participation of bis(salamo)-type compounds in ion recognition are not to be explored [48,49,50,51,52,53,54,55,56,57,58]. Notably, when compared with most of known fluorescence probes for Zn2+ and Cu2+, there are relatively few reports on fluorescent probes for Cd2+.
As part of our ongoing interest in the development of new optical (both colorimetric and fluorescence) chemosensor, in the present paper, we report a bis(salamo)-type tetraoxime sensor H4L for detecting Cd2+ by turn-on response in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solution. The sensor H4L has higher sensitivity for Cd2+ than other metallic ions that are based upon change in color by naked eyes. The mechanism of fluorescence change has been well demonstrated. The sensor H4L as a reliable fluorescence probe displayed high sensitivity toward Cd2+ in water-containing systems, and was able to detect Cd2+ in food samples.

2. Experimental

2.1. Materials and Methods

All of the reactions were performed under an air atmosphere. Boron tribromide (99.9%), methyl trioctyl ammonium chloride (90%), 2-hydroxy-3-methoxybenzaldehyde (99%), and pyridinium chlorochromate (98%) were gained from Alfa Aesar. Hydrobromic acid 33 wt % acetic acid solution was gained from J & K Scientific Ltd. (Beijing, China). Other solvents and reagents that were used in this work were analytical grade from Tianjin Chemical Reagent Factory (Tianjin, China). Melting points were measured using a microscopic melting point apparatus made in Beijing Taike Instrument Limited Company (Beijing, China) and were uncorrected. 1H NMR spectra were made via German Bruker AVANCEDRX-400 spectrophotometer (Karlsruhe, Germany). All of the UV–vis and fluorescence spectra tests were measured on a Shimadzu UV-2550 (Kyoto, Japan) and Perkin-Elmer LS-55 spectrometer (Waltham, MA, USA). The solvent also has a great influence on the fluorescence of the complex. Through a series of experiments, we finally obtained the best selectivity for cadmium ions in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) [59,60].

2.2. Preperation of Sensor H4L

The bis(salamo)-type tetraoxime sensor H4L was prepared on the basis of the reported methods [61,62,63,64,65,66,67,68]. The 1H NMR, IR, and UV-vis spectra of H4L are nearly consistent with the previous result (Figure 1). The major reaction steps of sensor H4L are demonstrated in Scheme 1.

3. Results and Discussion

3.1. pH Effect of Sensor H4L

For the sake of avoid the disturbance by protons in the recognition process of heavy metallic ions and to get optimum conditions, we concentrated on the pH influence on the fluorescence intensity. As depicted in Figure 2, the sensor H4L have barely changed in the fluorescent intensity from pH = 3.0 to 11.0, indicating that sensor H4L was consistent. The weak fluorescent sensor H4L may be owing to intra-molecular photo-induced electron transfer. However, H4L-Cd2+ showed strong fluorescence, on account of the bonding of H4L with Cd2+ lead to the inhibition of intra-molecular photo-induced electron transfer process. The results indicated that sensor H4L can be used to detect Cd2+ and the process of detection was not affected greatly by pH values.

3.2. General UV–vis Measurements

UV–vis spectra of sensor H4L in the existence of 14 metallic ions (Cd2+, Cu2+, Ba2+, Ca2+, K+, Cr3+, Mn2+, Sr2+, Co2+, Na+, Li+, Ni2+, Ag+, and Zn2+ with nitrate anions) were investigated. As shown in Figure 3a, sensor H4L have two intense absorption bands at 313 and 358 nm, which could be attributed to π → π* transition and reveals that sensor H4L includes a larger conjugation system. However, when 3.0 equiv. Cd2+ were added into the mixed solution, a new absorption peak emerged at 375 nm (Figure 3a). Meanwhile, the change in absorbance is almost same for Cd2+, Ni2+, Zn2+, and Mn2+ ions, indicating that the H4L were also involved in coordination to the Cd2+, Ni2+, Zn2+, and Mn2+ ions, respectively. However, when Cd2+ ion was added, sensor H4L could show a clear change in color from faint yellow to green that can be very easily observed via the naked eyes under UV light at 365 nm, while other metallic ions do not induce such a change. The colour change that is observed is mainly due to the charge transfer transition of the complex. The results revealed sensor H4L could be applied to selective identification of Cd2+ according to the color changes.
The responses of sensor H4L to Cd2+ (1 × 10−3 M) were studied further via UV–vis titration experiments, as depicted in Figure 3b. With the addition of Cd2+ from 0.0–3.0 equiv. in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions, two peaks that were observed at 342 and 358 nm were clearly declined and a new band at 375 nm emerged. Meanwhile, one definitive isosbestic point was also noted at 326 nm, indicating that a new species is produced. When 3.0 equiv. Cd2+ was added, the absorption no longer change, which showed 1:3 stoichiometry between Cd2+ and sensor H4L (Figure 3b).

3.3. General Fluorescence Measurements

For the sake of evaluating the selectivity of sensor H4L, original screening of H4L to the bonding ability of metallic ions was performed in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions. The fluorescent spectra of sensor H4L in the existence of a series of metallic ions (Cd2+, Cu2+, Ba2+, Ca2+, K+, Cr3+, Mn2+, Sr2+, Co2+, Na+, Li+, Ni2+, Ag+, and Zn2+ with nitrate anions) were gained, followed via excitation at 323 nm. As shown in Figure 4a, when Cd2+ ion was added, the position of the emission peak was red-shifted from 412 to 486 nm, the fluorescence intensity increased from 125 to 712, and the fluorescence quantum yield increased from 2.6 to 13.8%. Significantly, H4L showed excellent selectivity for Cd2+ in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions with a strong fluorescence response, with either a very weak or no fluorescence response to other metallic ions (Figure 4a). It is different from the previously reported sensors for Cd2+ showing a wide range of response to Zn2+, sensor H4L nearly non-fluorescent respond to Zn2+ could be owing to its rigid cavity with bigger size. Therefore, this rigid cavity could be definitely suitable for bonding with Cd2+, but not suitable for bonding with Zn2+ that possess smaller ionic radius. Furthermore, according to the corrected Benesi-Hildebrand formula, the bonding constant for the bonding of Cd2+ to sensor H4L was estimated as 4.98 × 104 M−1 and Zn2+ to sensor H4L was estimated as 3.89 × 104 M−1 [69]. The result suggested that sensor H4L displays outstanding selectivity for Cd2+ than other metallic ions.
For the sake of the practicability of sensor H4L as a receptor of Cd2+ selective probe, competitive experiments were performed with various metallic ions in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions. As shown in Figure 4b, there are no other ions can result any clear changes in the fluorescent spectrum of sensor. The selectivity of sensor H4L for Cd2+ and other metallic ions was measured. These results also demonstrated that other metallic ions could not interfere with the detection of Cd2+.
To further understand the coordination of sensor H4L with the Cd2+, fluorescent responses of sensor H4L to changing concentrations of Cd2+ in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions at room temperature were investigated in Figure 5. With the addition of increasing Cd2+ at an excitation wavelength of 323 nm, the fluorescent emission intensity at 485 nm gradually raised while the intensity at 407 nm reduced. Furthermore, after the addition of Cd2+, the absorbance at 485 nm exhibited a sharp increase when the ratio of [Cd2+]/[H4L] is below 3:1, and no longer change when the ratio reaches 3:1. In addition, the fitting curve of fluorescence emission intensity with Cd2+ concentrations was obtained by the data obtained of the fluorescence titration experiment (Figure 5). We have performed 1H NMR spectra of sensor H4L and in which the presence of 3.0 equiv. Cd2+ (Figure S1). Moreover, the detection limit is a considerably significant parameter in molecular recognition, the LOD and LOQ parameters were estimated to be 8.61 × 10−7 M and 2.87 × 10−6 M, respectively [61,69]. The LOD and LOQ were calculated based on the following equations:
LOD = 3 × δ/S; LOQ = 10 × δ/S.
where S represents the standard deviation of the blank measurements, and δ is the slope of the intensity versus sample concentration curve.
At the same time, the Hill equation is used in determining the binding constant of ions and H4L, the bonding constant for the bonding of Cd2+ to sensor H4L was estimated as 4.98 × 104 M−1 [48,59]. These results indicate that probe H4L displays satisfactory Cd2+ detecting ability.
log(F-Fmin)/(Fmax-F) = log K + n log [Cd2+]; n = 3.
where Fmin, Fmax, and F are the emission intensities in the absence, presence of saturated Cd2+, and the addition of a given amount of Cd2+ concentration, respectively. [Cd2+] is the concentration of free Cd2+.
To know the stoichiometry between the sensor H4L and Cd2+ in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions, job’s plot has been performed (Figure 6). When the molar fraction of Cd2+ was 0.75, the intensity at 495 nm reached an extreme value, indicating the formation of a 3:1 complex between Cd2+ and H4L.
Furthermore, according to the reversible fluorescent switch of the sensor H4L to Cd2+ in a coordination compound solution, we regard it as a two-input molecular logic gate, while the emission at 373 nm serves as the output. As depicted in Figure 7 and Table 1, when the output was zero (both the Cd2+ and EDTA are absent), corresponding the gate being closed and this system shows weak fluorescence. When Cd2+ alone was existent, the output is one and the relevant to the gate being open, so intense fluorescence was observed. Thus, the sensor H4L is able to serve as a logic gate. This result demonstrates that the sensor H4L as a reversible fluorescence probe.

3.4. The Detecting Mechanism of H4L for Cd2+

According to the fluorescent spectra, the detecting mechanism of the sensor H4L for Cd2+ was suggested, as follows (Figure 8). The fluorescent intensity of the sensor H4L response to Cd2+ may be assigned to CHEF and PET. Before being coordinated with Cd2+, sensor H4L displayed a weaker fluorescence due to the lone pair of electrons of nitrogen atoms, which gives rise to an intra-molecular PET. Furthermore, the lone electron pairs of the nitrogen atoms give rise to a nonradiative process by the n-π* state, which also led to a wide degree of fluorescent quenching. Conversely, after H4L was coordinated to Cd2+, the radiation process was primarily via the π-π* state and the coordination compound was more rigid [48,59]. In addition, the PET process was restrained by the addition of Cd2+ analyte at the receptor site. It is obvious that the appearance of the ICT process influenced PET in this system, but fluorophore and receptor distances and the orientation between them can also contribute in the overall PET process.

3.5. Time and Temperature Effect of Probe H4L

The influences of time and temperature on the fluorescent intensity of sensor H4L toward Cd2+ were also studied in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions. Figure 9 illustrates that fluorescence intensity of sensor H4L did not vary with further prolong the reaction time. Besides, the fluorescent intensity of sensor H4L almost constant in the temperature range of 0–90 °C. Therefore, sensor H4L could be used for rapid response to Cd2+ in room temperature, which is of significant practicability for the detection of Cd2+.

3.6. Test Strips Measurements

Currently, test strips analytical equipment have gained momentous attention on account of their high sensitivity, low-cost, and quick response. It is perhaps the most convenient of modern detection tools, as a change in color could be seen via the naked eyes. Hence, in this study, filter papers were performed via soaking filter papers into Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) mixed solutions of probe H4L (1 × 10−5 M), and then drying by exposure to air. The filter papers, including sensor H4L, were applied to detect Cd2+ and other metallic ions. As depicted in Figure 10, after Cd2+ and the other metallic ions were added on the test tools, respectively, the distinct color changes were seen merely with Cd2+ solution under UV light at 365 nm, and potentially competitive metallic ions have no work on the detection of Cd2+ via the filter papers (Figure S2). It is a really actual convenient method that quickly measures the Cd2+, and it could be able to be satisfactorily used in the fields of food security and environmental surveillance.

3.7. Application in Food Samples

Furthermore, we investigated the applicability of sensor H4L in food samples. 500 mg of crushed naturally containing cadmium rice were put into the PTFE microwave digestion tank by the addition of 3 mL of HNO3 and 4 mL of H2O2, and the mixture solution containing cadmium(II) needs be digested after keeping 30–60 min. Then, the mixture solution was dried in vacuo and was added into the mixed solution of DMF and H2O (c = 1 × 10−3 M, DMF/H2O = 9:1, v/v, pH = 7). As shown in Figure 11, after added the sample solution to sensor H4L, the fluorescence intensity is on the increase.

4. Conclusions

In this paper, we presented a new bis(salamo)-type fluorescence sensor H4L, which could serve as a great promising analytical kit for measuring Cd2+ via different fluorescence changes and changes in color from light yellow to green that can be measured via naked eyes. The detection limit about fluorescent response of probe H4L to Cd2+ was down to 8.61 × 10−7 M. As designed, the filter papers could conveniently, efficiently, and simply detect Cd2+ in solution. Moreover, Cd2+ in the food samples was detected by the sensor H4L simply and effectively. We believe this study will inspire the development of bis(salamo)-based chemosensor by optimizing unsaturated metal coordinating sites for practicability to many other metallic ions detecting in analytical chemistry, medical treatment, biological, and environmental fields.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/11/4/523/s1, Figure S1: 1H NMR titration in upon addition of 3.0 equiv. Cd2+, Figure S2: The result of colorimetric measured photographs with sensor H4L for detecting Cd2+ under irradiation at 365 nm.

Acknowledgments

This project was supported by the National Natural Science Foundation of China (grant 21761018), the Program for Excellent Team of Scientific Research in Lanzhou Jiaotong University (201706), which is gratefully acknowledged.

Author Contributions

Jing Hao performed most of the experiments. Xiao-Yan Li and Yang Zhang contributed to the writing of the manuscript. Wen-Kui Dong designed the project. All authors reviewed the manuscript.

Conflicts of Interest

There is no conflict of interest among all authors.

References

  1. Chai, L.Q.; Tang, L.J.; Chen, L.C.; Huang, J.J. Structural, spectral, electrochemical and DFT studies of two mononuclear manganese(II) and zinc(II) complexes. Polyhedron 2017, 122, 228–240. [Google Scholar] [CrossRef]
  2. Chai, L.Q.; Zhang, K.Y.; Tang, L.J.; Zhang, J.Y.; Zhang, H.S. Two mono- and dinuclear Ni(II) complexes constructed from quinazoline-type ligands: Synthesis, X-ray structures, spectroscopic, electrochemical, thermal, and antimicrobial studies. Polyhedron 2017, 130, 100–107. [Google Scholar] [CrossRef]
  3. Chai, L.Q.; Zhang, J.Y.; Chen, L.C.; Li, Y.X.; Tang, L.J. Synthesis, crystal structure, spectroscopic properties and DFT calculations of a new schiff base-type zinc(II) complex. Res. Chem. Intermed. 2016, 42, 3473–3488. [Google Scholar] [CrossRef]
  4. Chai, L.Q.; Li, Y.X.; Chen, L.C.; Zhang, J.Y.; Huang, J.J. Synthesis, X-ray structure, spectroscopic, electrochemical properties and DFT calculation of a bridged dinuclear copper(II) complex. Inorg. Chim. Acta 2016, 444, 193–201. [Google Scholar] [CrossRef]
  5. Liu, X.; Wang, P.; Fu, J.; Yao, K.; Xue, K.; Xu, K. Turn-on fluorescent sensor for zinc and cadmium ions based on quinolone and its sequential response to phosphate. J. Lumin. 2017, 186, 16–22. [Google Scholar] [CrossRef]
  6. Marettová, E.; Maretta, M.; Legáth, J. Toxic effects of cadmium on testis of birds and mammals: A review. Anim. Reprod. Sci. 2015, 155, 1–10. [Google Scholar] [CrossRef] [PubMed]
  7. Al-Khaldi, F.A.; Abusharkh, B.; Khaled, M.; Atieh, M.A.; Nasser, M.S. Adsorptive removal of cadmium(II) ions from liquid phase using acid modified carbon-based adsorbents. J. Mol. Liq. 2015, 204, 255–263. [Google Scholar]
  8. Akesson, A.; Julin, B.; Wolk, A. Long-term dietary cadmium intake and postmenopausal endometrial cancer incidence: A population-based prospective cohort study. Cancer Res. 2008, 68, 6435–6441. [Google Scholar] [CrossRef] [PubMed]
  9. Yang, L.L.; Liu, X.M.; Liu, K.; Liu, H.; Zhao, F.Y.; Ruan, W.J. A polypyridyl-pyrene based off-on Cd2+ fluorescent sensor for aqueous phase analysis and living cell imaging. Talanta 2014, 128, 278–283. [Google Scholar] [CrossRef] [PubMed]
  10. Vinod, K.; Gupta, M.R.; Ganjali, P.; Norouzi, H.; Khani, A.N.; Shilpi, A. Electrochemical analysis of some toxic metals by ion–selective electrodes. Crit. Rev. Anal. Chem. 2011, 41, 282–313. [Google Scholar]
  11. Gupta, V.K.; Jain, A.K.; Agarwal, S.; Maheshwari, G. An iron(III) ion-selective sensor based on a mu-bis(tridentate) ligand. Talanta 2007, 71, 1964–1968. [Google Scholar] [CrossRef] [PubMed]
  12. Gupta, V.K.; Chandra, S.; Lang, H. A highly selective mercury electrode based on a diamine donor ligand. Talanta 2005, 66, 575–580. [Google Scholar] [CrossRef] [PubMed]
  13. Gupta, V.K.; Sethi, B.; Sharma, R.A.; Agarwal, S.; Bharti, A. Mercury selective potentiometric sensor based on low rim functionalized thiacalix-arene as a cationic receptor. J. Mol. Liq. 2013, 177, 114–118. [Google Scholar] [CrossRef]
  14. Xu, Z.; Zhang, L.; Guo, R.; Xiang, T.; Wu, C.; Zheng, Z. A highly sensitive and selective colorimetric and off–on fluorescent chemosensor for Cu2+, based on rhodamine derivative. Sens. Actuators B 2011, 156, 546–552. [Google Scholar] [CrossRef]
  15. Jun, M.E.; Roy, B.; Ahn, K.H. “Turn-On” fluorescent sensing with “Reactive” probes. Chem. Commun. 2011, 47, 7583–7601. [Google Scholar] [CrossRef] [PubMed]
  16. Du, J.; Hu, M.; Fan, J.; Peng, X. Cheminform abstract: Fuorescent chemodosimeters using “mild” chemical events for the detection of small anions and cations in biological and environmental media. Chem. Soc. Rev. 2012, 41, 4511–4535. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, X.; Pradhan, T.; Wang, F.; Kim, J.S.; Yoon, J. Fluorescent chemosensors based on spiroring-opening of xanthenes and related derivatives. Chem. Rev. 2012, 112, 1910–1956. [Google Scholar] [CrossRef] [PubMed]
  18. Li, X.Y.; Chen, L.; Gao, L.; Zhang, Y.; Akogun, S.F.; Dong, W.K. Syntheses, crystal structures and catalytic activities of two solvent-induced homotrinuclear Co(II) complexes with a naphthalenediol-based bis(Salamo)-type tetraoxime ligand. RSC Adv. 2017, 7, 35905–35916. [Google Scholar] [CrossRef]
  19. Wang, L.; Ma, J.C.; Dong, W.K.; Zhu, L.C.; Zhang, Y. A novel Self–assembled nickel(II)–cerium(III) heterotetranuclear dimer constructed from N2O2-type bisoxime and terephthalic acid: Synthesis, structure and photophysical properties. Z. Anorg. Allg. Chem. 2016, 642, 834–839. [Google Scholar] [CrossRef]
  20. Ma, J.C.; Dong, X.Y.; Dong, W.K.; Zhang, Y.; Zhu, L.C.; Zhang, J.T. An unexpected dinuclear Cu(II) complex with a bis(Salamo) chelating ligand: Synthesis, crystal structure, and photophysical properties. J. Coord. Chem. 2016, 69, 149–159. [Google Scholar] [CrossRef]
  21. Tao, C.H.; Ma, J.C.; Zhu, L.C.; Zhang, Y.; Dong, W.K. Heterobimetallic 3d–4f Zn(II)–Ln(III) (Ln = Sm, Eu, Tb and Dy) complexes with a N2O4 bisoxime chelate ligand and a simple auxiliary ligand Py: Syntheses, structures and luminescence properties. Polyhedron 2017, 128, 38–45. [Google Scholar] [CrossRef]
  22. Dong, Y.J.; Dong, X.Y.; Dong, W.K.; Zhang, Y.; Zhang, L.S. Three asymmetric Salamo-type copper(II) and cobalt(II) complexes: Syntheses, structures, fluorescent properties. Polyhedron 2017, 123, 305–315. [Google Scholar] [CrossRef]
  23. Dong, Y.J.; Ma, J.C.; Zhu, L.C.; Dong, W.K.; Zhang, Y. Four 3d–4f heteromultinuclear zinc(II)–lanthanide(III) complexes constructed from a distinct hexadentate N2O2-type ligand: Syntheses, structures and photophysical properties. J. Coord. Chem. 2017, 70, 103–115. [Google Scholar] [CrossRef]
  24. Song, X.Q.; Liu, P.P.; Liu, Y.A.; Zhou, J.J.; Wang, X.L. Two dodecanuclear heterometallic [Zn6Ln6] clusters constructed by a multidentate salicylamide salen-like ligand: Synthesis, structure, luminescence and magnetic properties. Dalton Trans. 2016, 45, 8154–8163. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, P.P.; Sheng, L.; Song, X.Q.; Xu, W.Y.; Liu, Y.A. Synthesis, structure and magnetic properties of a new one dimensional manganese coordination polymer constructed by a new asymmetrical ligand. Inorg. Chim. Acta 2015, 434, 252–257. [Google Scholar] [CrossRef]
  26. Song, X.Q.; Liu, P.P.; Xiao, Z.R.; Li, X.; Liu, Y.A. Four polynuclear complexes based on a versatile salicylamide salen-like ligand: Synthesis, structural variations and magnetic properties. Inorg. Chim. Acta 2015, 438, 232–244. [Google Scholar] [CrossRef]
  27. Song, X.Q.; Peng, Y.J.; Chen, G.Q.; Wang, X.R.; Liu, P.P.; Xu, W.Y. Substituted group-directed assembly of Zn(II) coordination complexes based on two new structural related pyrazolone based Salen ligands: Syntheses, structures and fluorescence properties. Inorg. Chim. Acta 2015, 427, 13–21. [Google Scholar] [CrossRef]
  28. Liu, P.P.; Wang, C.Y.; Zhang, M.; Song, X.Q. Pentanuclear sandwich-type ZnII-LnIII clusters based on a new Salen-like salicylamide ligand: Structure, near-infrared emission and magnetic properties. Polyhedron 2017, 129, 133–140. [Google Scholar] [CrossRef]
  29. Wang, P.; Zhao, L. Synthesis, structure and spectroscopic properties of the trinuclear cobalt(II) and nickel(II) complexes based on 2-hydroxynaphthaldehyde and bis(aminooxy)alkane. Spectrochim. Acta Part A 2015, 135, 342–350. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, P.; Zhao, L. An infinite 2D supramolecular cobalt(II) complex based on an asymmetric Salamo-type ligand: Synthesis, crystal structure, and spectral properties. Synth. React. Inorg. Met-Org. Nano-MetChem. 2016, 46, 1095–1101. [Google Scholar] [CrossRef]
  31. Hu, J.H.; Li, J.B.; Qi, J.; Sun, Y. Selective colorimetric and “turn-on” fluorimetric detection of cyanide using an acylhydrazone sensor in aqueous media. New J. Chem. 2015, 39, 4041–4046. [Google Scholar] [CrossRef]
  32. Hu, J.H.; Li, J.B.; Qi, J.; Chen, J.J. Highly selective and effective mercury(II) fluorescent sensor. New J. Chem. 2015, 39, 843–848. [Google Scholar] [CrossRef]
  33. Li, J.B.; Hu, J.H.; Chen, J.J.; Qi, J. Cyanide detection using a benzimidazole derivative in aqueous media. Spectrochim. Acta A 2014, 133, 773–777. [Google Scholar] [CrossRef] [PubMed]
  34. Hu, J.H.; Li, J.B.; Qi, J.; Sun, Y. Acylhydrazone based fluorescent chemosensor for zinc in aqueous solution with high selectivity and sensitivity. Sens. Actuators B Chem. 2015, 208, 581–587. [Google Scholar] [CrossRef]
  35. Sun, Y.; Hu, J.H.; Qi, J.; Li, J.B. A highly selective colorimetric and “turn-on” fluorimetric chemosensor for detecting CN—Based on unsymmetrical azine derivatives in aqueous media. Spectrochim. Acta A 2016, 167, 101–105. [Google Scholar] [CrossRef] [PubMed]
  36. Hu, J.H.; Chen, J.J.; Li, J.B.; Qi, J. A cyanide ion probe based on azosalicylic aldehyde of benzoyl hydrazone. Chin. J. Inorg. Chem. 2014, 30, 2544–2548. [Google Scholar]
  37. Qi, J.; Hu, J.H.; Chen, J.J.; Sun, Y.; Li, J.B. Cyanide Detection using Azo-acylhydrazone in Aqueous Media with High Sensitivity and Selectivity. Curr. Anal. Chem. 2016, 12, 119–123. [Google Scholar] [CrossRef]
  38. Hu, J.H.; Li, J.B.; Qi, J.; Sun, Y. Studies on the crystal structure and characterization of n-(4-acetylphenyl)-na-(2-nitrobenzoyl)-thiourea. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 984–987. [Google Scholar] [CrossRef]
  39. Hu, J.H.; Sun, Y.; Qi, J.; Pei, P.X.; Lin, Q.; Zhang, Y.M. A colorimetric and “turn-on” fluorimetric chemosensor for the selective detection of cyanide and its application in food samples. RSC Adv. 2016, 6, 100401–100406. [Google Scholar] [CrossRef]
  40. Hu, J.H.; Sun, Y.; Qi, J.; Li, Q.; Wei, T.B. A new unsymmetrical azine derivative based on coumarin group as dual-modal sensor for CN, and fluorescent “off–on” for Zn2+. Spectrochim. Acta A 2017, 175, 125–133. [Google Scholar] [CrossRef] [PubMed]
  41. Wu, H.L.; Pan, G.L.; Wang, H.; Wang, X.L.; Bai, Y.C.; Zhang, Y.H. Study on synthesis, crystal structure, antioxidant and DNA-binding of mono-, di- and poly-nuclear lanthanides complexes with bis(N-salicylidene)-3-oxapentane-1,5-diamine. J. Photochem. Photobiol. B Biol. 2014, 135, 33–43. [Google Scholar] [CrossRef] [PubMed]
  42. Wu, H.L.; Bai, Y.C.; Zhang, Y.H.; Li, Z.; Wu, M.C.; Chen, C.Y.; Zhang, J.W. Synthesis, crystal structure, antioxidation and DNA-binding properties of a dinuclear copper(II) complex with bis(N-salicylidene)-3-oxapentane-1,5-diamine. J. Coord. Chem. 2014, 67, 3054–3066. [Google Scholar] [CrossRef]
  43. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J. Synthesis, structure, antioxidation, and DNA-bindingstudies of a binuclear ytterbium(III) complex with bis(N-salicylidene)-3-oxapentane-1,5-diamine. Res. Chem. Intermed. 2015, 41, 3375–3388. [Google Scholar] [CrossRef]
  44. Wu, H.L.; Wang, C.P.; Wang, F.; Peng, H.P.; Zhang, H.; Bai, Y.C. A new manganese(III) complex from bis(5-methylsalicylaldehyde)-3-oxapentane-1,5-diamine: Synthesis, characterization, antioxidant activity and luminescence. J. Chin. Chem. Soc. 2015, 62, 1028–1034. [Google Scholar] [CrossRef]
  45. Chen, C.Y.; Zhang, J.W.; Zhang, Y.H.; Yang, Z.H.; Wu, H.L. Gadolinium(III) and dysprosium(III) complexes with a Schiff base bis(N-salicylidene)-3-oxapentane-1,5-diamine: Synthesis, characterization, antioxidation, and DNA-binding studies. J. Coord. Chem. 2015, 68, 1054–1071. [Google Scholar] [CrossRef]
  46. Wu, H.L.; Bai, Y.C.; Zhang, Y.H.; Pan, G.L.; Kong, J.; Shi, F.; Wang, X.L. Two lanthanide(III) complexes based on the schiff base N,N-Bis(salicylidene)-1,5-diamino-3-oxapentane: Synthesis, characterization, DNA-binding properties, and antioxidation. Z. Anorg. Allg. Chem. 2014, 640, 2062–2071. [Google Scholar] [CrossRef]
  47. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J.; Shi, F.; Zhang, Y.H.; Wang, X.L. Preparation, structure, DNA-binding properties, and antioxidant activities of a homodinuclear erbium(III) complex with a pentadentate Schiff base ligand. J. Chem. Res. 2014, 38, 211–217. [Google Scholar] [CrossRef]
  48. Wang, B.J.; Dong, W.K.; Zhang, Y.; Akogun, S.F. A novel relay-sensor for highly sensitive and selective detection of Zn2+/Pic- and fluorescence on/off switch response of H+/OH. Sens. Actuators B 2017, 247, 254–264. [Google Scholar] [CrossRef]
  49. Dong, Y.J.; Li, X.L.; Zhang, Y.; Dong, W.K. A highly selective visual and fluorescent sensor for Pb2+ and Zn2+ and crystal structure of Cu2+ complex based-on a novel single-armed Salamo-type bisoxime. Supramol. Chem. 2017, 29, 518–527. [Google Scholar] [CrossRef]
  50. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y.; Li, X.L. Four new nickel(II) complexes based on an asymmetric Salamo-type ligand: Synthesis, structure, solvent effect and electrochemical property. Inorg. Chim. Acta 2016, 445, 140–148. [Google Scholar] [CrossRef]
  51. Dong, W.K.; Lan, P.F.; Zhou, W.M.; Zhang, Y. Salamo-type trinuclear and tetranuclear cobalt(II) complexes based on a new asymmetry salamo-type ligand: Syntheses, crystal structures and fluorescence properties. J. Coord. Chem. 2016, 65, 1272–1283. [Google Scholar] [CrossRef]
  52. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y. Self-assembled zinc(II)-lanthanide(III) heteromultinuclear complexes constructed from 3-MeOsalamo ligand: Syntheses, structures and luminescent properties. Cryst. Growth Des. 2016, 16, 6903–6914. [Google Scholar] [CrossRef]
  53. Dong, X.Y.; Kang, Q.P.; Jin, B.X.; Dong, W.K. A dinuclear nickel(II) complex derived from an asymmetric Salamo-type N2O2 chelate ligand: Synthesis, structure and optical properties. Z. Naturforsch. 2017, 72, 415–420. [Google Scholar] [CrossRef]
  54. Dong, W.K.; Zhang, J.T.; Dong, Y.J.; Zhang, Y.; Wang, Z.K. Construction of mononuclear copper(II) and trinuclear cobalt(II) complexes based on asymmetric Salamo-type ligands. Z. Anorg. Allg. Chem. 2016, 642, 189–196. [Google Scholar] [CrossRef]
  55. Dong, W.K.; Li, X.L.; Wang, L.; Zhang, Y.; Ding, Y.J. A new application of Salamo-type bisoximes: As a relay-sensor for Zn2+/Cu2+ and its novel complexes for successive sensing of H+/OH. Sens. Actuators B 2016, 229, 370–378. [Google Scholar] [CrossRef]
  56. Dong, W.K.; Zhang, J.; Zhang, Y.; Li, N. Novel multinuclear transition metal(II) complexes based on an asymmetric Salamo-type ligand: Syntheses, structure characterizations and fluorescent properties. Inorg. Chim. Acta 2016, 444, 95–102. [Google Scholar] [CrossRef]
  57. Dong, W.K.; Ma, J.C.; Dong, Y.J.; Zhu, L.C.; Zhang, Y. Di-and tetranuclear heterometallic 3d-4f cobalt(II)-lanthanide(III) complexes derived from a hexadentate bisoxime: Syntheses, structures and magnetic properties. Polyhedron 2016, 115, 228–235. [Google Scholar] [CrossRef]
  58. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y. Nine self-assembled nickel(II)-lanthanide(III) heterometallic complexes constructed from a Salamo-type bisoxime and bearing N- or O-donor auxiliary ligand: Syntheses, structures and magnetic properties. New J. Chem. 2016, 40, 6998–7010. [Google Scholar] [CrossRef]
  59. Gao, L.; Wang, F.; Zhao, Q.; Zhang, Y.; Dong, W.K. Mononuclear Zn(II) and trinuclear Ni(II) complexes derived from a coumarin-containing N2O2 ligand: Syntheses, crystal structures and fluorescence properties. Polyhedron 2018, 139, 7–16. [Google Scholar] [CrossRef]
  60. Chen, L.; Dong, W.K.; Zhang, H.; Zhang, Y.; Sun, Y.X. Structural variation and luminescence properties of tri- and dinuclear CuII and ZnII complexes constructed from a naphthalenediol-based bis(Salamo)-type ligand. Cryst. Growth Des. 2017, 17, 3636–3648. [Google Scholar] [CrossRef]
  61. Dong, W.K.; Akogun, S.F.; Zhang, Y.; Sun, Y.X.; Dong, X.Y. A reversible “turn-on” fluorescent sensor for selective detection of Zn2+. Sens. Actuators B 2017, 238, 723–734. [Google Scholar] [CrossRef]
  62. Hao, J.; Liu, L.Z.; Dong, W.K.; Zhang, J.; Zhang, Y. Three multinuclear Co.(II), Zn(II) and Cd(II) complexes based on a single-armed salamo-type bisoxime: Syntheses, structural characterizations and fluorescent properties. J. Coord. Chem. 2017, 11, 1936–1952. [Google Scholar] [CrossRef]
  63. Zheng, S.S.; Dong, W.K.; Zhang, Y.; Chen, L.; Ding, Y.J. Four Salamo-type 3d–4f hetero-bimetallic [ZnIILnIII] complexes: Syntheses, crystal structures, and luminescent and magnetic properties. New J. Chem. 2017, 41, 4966–4973. [Google Scholar] [CrossRef]
  64. Dong, W.K.; Zheng, S.S.; Zhang, J.T.; Zhang, Y.; Sun, Y.X. Luminescent properties of heterotrinuclear 3d–4f complexes constructed from a naphthalenediol-based acyclic bis(salamo)-type ligand. Spectrochim. Acta Part A 2017, 184, 141–150. [Google Scholar] [CrossRef] [PubMed]
  65. Li, G.; Hao, J.; Liu, L.Z.; Zhou, W.M.; Dong, W.K. Syntheses, crystal structures and thermal behaviors of two supramolecular salamo-type cobalt(II) and zinc(II) complexes. Crystals 2017, 7, 217. [Google Scholar]
  66. Peng, Y.D.; Li, X.Y.; Kang, Q.P.; An, G.X.; Zhang, Y.; Dong, W.K. Synthesis and fluorescence properties of asymmetrical salamo-type tetranuclear zinc(II) complex. Crystals 2018, 8, 107. [Google Scholar] [CrossRef]
  67. Hao, J.; Li, L.L.; Zhang, J.T.; Akogun, S.F.; Wang, L.; Dong, W.K. Four homo- and hetero-bismetallic 3d/3d-2s complexes constructed from a naphthalenediol-based acyclic bis(salamo)-type tetraoxime ligand. Polyhedron 2017, 134, 1–10. [Google Scholar] [CrossRef]
  68. Wang, L.; Hao, J.; Zhai, L.X.; Zhang, Y.; Dong, W.K. Synthesis, crystal structure, luminescence, electrochemical and antimicrobial properties of bis(salamo)-based Co.(II) complex. Crystals 2017, 7, 277. [Google Scholar] [CrossRef]
  69. Wang, F.; Gao, L.; Zhao, Q.; Zhang, Y.; Dong, W.K.; Ding, Y.J. A highly selective fluorescent chemosensor for CN based on a novel bis(salamo)-type tetraoxime ligand. Spectrochim. Acta Part A 2018, 190, 111–115. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthetic route to sensor H4L.
Scheme 1. Synthetic route to sensor H4L.
Materials 11 00523 sch001
Figure 1. 1HNMR spectra of sensor H4L in DMSO.
Figure 1. 1HNMR spectra of sensor H4L in DMSO.
Materials 11 00523 g001
Figure 2. Changes in fluorescence intensity of H4L (c = 1 × 10−5 M) and Cd2+ (c = 1 × 10−3 M) performed at different pH values at room temperature. (DMF/H2O = 9:1, v/v, λex = 323 nm, λem = 495 nm).
Figure 2. Changes in fluorescence intensity of H4L (c = 1 × 10−5 M) and Cd2+ (c = 1 × 10−3 M) performed at different pH values at room temperature. (DMF/H2O = 9:1, v/v, λex = 323 nm, λem = 495 nm).
Materials 11 00523 g002
Figure 3. (a) UV-vis spectra of sensor H4L (1 × 10−5 M) recorded in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions after addition of 3.0 equiv. of metallic ions (1 × 10−3 M); (b) Absorption spectra of sensor H4L in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions upon gradual addition of Cd2+ (0.0 to 3.0 equiv.).
Figure 3. (a) UV-vis spectra of sensor H4L (1 × 10−5 M) recorded in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions after addition of 3.0 equiv. of metallic ions (1 × 10−3 M); (b) Absorption spectra of sensor H4L in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions upon gradual addition of Cd2+ (0.0 to 3.0 equiv.).
Materials 11 00523 g003
Figure 4. (a) Fluorescent spectra and (b) fluorescent intensity at 323 nm of sensor H4L (1 × 10−5 M) in the existence of different metallic ions (c = 1 × 10−3 M, excess amounts Cu2+, Ba2+, Ca2+, K+, Cr3+, Mn2+, Sr2+, Co2+, Na+, Li+, Ni2+, Ag+, and Zn2+) in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0, λex = 323 nm) solution (the black bars delegate the addition of excess amounts of metallic ions to the solution of probe H4L and the pink bars delegate the subsequent addition of Cd2+ to the mixed solution).
Figure 4. (a) Fluorescent spectra and (b) fluorescent intensity at 323 nm of sensor H4L (1 × 10−5 M) in the existence of different metallic ions (c = 1 × 10−3 M, excess amounts Cu2+, Ba2+, Ca2+, K+, Cr3+, Mn2+, Sr2+, Co2+, Na+, Li+, Ni2+, Ag+, and Zn2+) in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0, λex = 323 nm) solution (the black bars delegate the addition of excess amounts of metallic ions to the solution of probe H4L and the pink bars delegate the subsequent addition of Cd2+ to the mixed solution).
Materials 11 00523 g004aMaterials 11 00523 g004b
Figure 5. (a) Fluorescent emission spectra of sensor H4L (1 × 10−5 M) with subsequent addition of Cd2+ (0–3.0 equiv. λex = 323 nm) in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions. Inset: Naked-eyes visible colour changes of sensor H4L solution before and after addition of Cd2+. (b) Linear fitting of sensor H4L to Cd2+ bonding constant.
Figure 5. (a) Fluorescent emission spectra of sensor H4L (1 × 10−5 M) with subsequent addition of Cd2+ (0–3.0 equiv. λex = 323 nm) in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0) solutions. Inset: Naked-eyes visible colour changes of sensor H4L solution before and after addition of Cd2+. (b) Linear fitting of sensor H4L to Cd2+ bonding constant.
Materials 11 00523 g005
Figure 6. The job’s plot examined between Cd2+ and H4L, indicating the 3:1 stoichiometry, which was carried out by fluorescence spectra (λex = 323 nm).
Figure 6. The job’s plot examined between Cd2+ and H4L, indicating the 3:1 stoichiometry, which was carried out by fluorescence spectra (λex = 323 nm).
Materials 11 00523 g006
Figure 7. (a) Schematic presentation of “OFF–ON” system for sensor H4L (c = 1 × 10−5 M) in the existence of Cd2+ (c = 1 × 10−3 M) and EDTA (c = 1 × 10−3 M); (b) Performance of sensor H4L in the INHIBIT gate mode in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0, λex = 373 nm). The bars display the fluorescent output of sensor H4L at 373 nm in the existence of Cd2+ and EDTA serve as inputs.
Figure 7. (a) Schematic presentation of “OFF–ON” system for sensor H4L (c = 1 × 10−5 M) in the existence of Cd2+ (c = 1 × 10−3 M) and EDTA (c = 1 × 10−3 M); (b) Performance of sensor H4L in the INHIBIT gate mode in Tris-phosphate buffer (c = 0.2 M, DMF/H2O = 9:1, v/v, pH = 7.0, λex = 373 nm). The bars display the fluorescent output of sensor H4L at 373 nm in the existence of Cd2+ and EDTA serve as inputs.
Materials 11 00523 g007aMaterials 11 00523 g007b
Figure 8. The sensing mechanism of the sensor H4L for Cd2+.
Figure 8. The sensing mechanism of the sensor H4L for Cd2+.
Materials 11 00523 g008
Figure 9. (a) Fluorescence intensity changes of the solution contain sensor H4L and 3.0 equiv. Cd2+ with the delay of time; (b) Fluorescent intensity changes of the solution contain H4L and 3.0 equiv. Cd2+ at different temperatures.
Figure 9. (a) Fluorescence intensity changes of the solution contain sensor H4L and 3.0 equiv. Cd2+ with the delay of time; (b) Fluorescent intensity changes of the solution contain H4L and 3.0 equiv. Cd2+ at different temperatures.
Materials 11 00523 g009aMaterials 11 00523 g009b
Figure 10. The result of colorimetric measured photographs with sensor H4L for detecting Cd2+ under irradiation at 365 nm.
Figure 10. The result of colorimetric measured photographs with sensor H4L for detecting Cd2+ under irradiation at 365 nm.
Materials 11 00523 g010
Figure 11. Fluorescence spectral response of sensor H4L in cadmium contaminated rice.
Figure 11. Fluorescence spectral response of sensor H4L in cadmium contaminated rice.
Materials 11 00523 g011
Table 1. The molecular logic gate and the each symbolic expression of the INHIBIT logic gate function.
Table 1. The molecular logic gate and the each symbolic expression of the INHIBIT logic gate function.
InputsOutputs
Cd2+EDTA(INHIBIT Logic Gate) Intensity λmax
00low flu.0
10high flu.1
01low flu.0
11low flu.0

Share and Cite

MDPI and ACS Style

Hao, J.; Li, X.-Y.; Zhang, Y.; Dong, W.-K. A Reversible Bis(Salamo)-Based Fluorescence Sensor for Selective Detection of Cd2+ in Water-Containing Systems and Food Samples. Materials 2018, 11, 523. https://doi.org/10.3390/ma11040523

AMA Style

Hao J, Li X-Y, Zhang Y, Dong W-K. A Reversible Bis(Salamo)-Based Fluorescence Sensor for Selective Detection of Cd2+ in Water-Containing Systems and Food Samples. Materials. 2018; 11(4):523. https://doi.org/10.3390/ma11040523

Chicago/Turabian Style

Hao, Jing, Xiao-Yan Li, Yang Zhang, and Wen-Kui Dong. 2018. "A Reversible Bis(Salamo)-Based Fluorescence Sensor for Selective Detection of Cd2+ in Water-Containing Systems and Food Samples" Materials 11, no. 4: 523. https://doi.org/10.3390/ma11040523

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop