Next Article in Journal
Bioactive Properties of a Novel Antibacterial Dye Obtained from Laccase-Mediated Oxidation of 8-Anilino-1-naphthalenesulfonic Acid
Next Article in Special Issue
Anion-Sensing Properties of Cyclopentaphenylalanine
Previous Article in Journal
Potential COVID-19 Drug Candidates Based on Diazinyl-Thiazol-Imine Moieties: Synthesis and Greener Pastures Biological Study
Previous Article in Special Issue
Ferrocene-Labelled Electroactive Aptamer-Based Sensors (Aptasensors) for Glycated Haemoglobin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Double-Site Chemodosimeter for Selective Fluorescence Detection of a Nerve Agent Mimic

1
School of Life Science and Engineering, Southwest Jiaotong University, Chengdu 610031, China
2
Chengdu Municipal Bureau of Economic and Information Technology, Chengdu 610229, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(2), 489; https://doi.org/10.3390/molecules27020489
Submission received: 18 December 2021 / Revised: 7 January 2022 / Accepted: 11 January 2022 / Published: 13 January 2022
(This article belongs to the Special Issue Recent Advances in Molecular Sensors)

Abstract

:
A novel two-site chemodosimeter (SWJT-4) based on fluorescein skeleton to detect diethyl chlorophosphate (DCP) was designed and synthesized. It is a turn-on fluorescent probe for DCP with good selectivity and obvious color change in aqueous solution. Interestingly, the two oxime groups of SWJT-4 as dual response sites initiated different reactions with DCP to form a cyano group and an isoxazole ring, respectively. The corresponding mechanism was confirmed by 1H NMR, MS and DFT calculation. Moreover, SWJT-4 could be used as a fluorescent test paper to detect DCP vapor.

1. Introduction

Organophosphate nerve agents refer to a class of chemical substances composed of organophosphorus compounds (OPs). They have been widely used in pesticides and chemical warfare agents, such as parathion, systemic phosphorus, malathion, dimethoate, DDVP, tabun, sarin and soman [1,2]. These chemicals can result in a range of neurological symptoms such as headache, dizziness and agitation [3,4,5,6]. Organophosphorus compounds were used as a chemical weapon in war to threaten the safety of human life seriously. In a word, organophosphorus nerve agents are not only a potential threat of biochemical warfare, but also a usual weapon of terrorist organizations. Therefore, the detection of these substances by a convenient method is very necessary.
At present, the fluorescence analysis methods have attracted widespread attention due to their real-time monitoring, high selectivity, high sensitivity, low detection limit and so on. To obtain specific recognition of diethyl chlorophosphate (DCP), an organophosphate nerve agent mimic, some fluorescence probes were designed as chemodosimeters based on organic reaction [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30]. Notably, DCP can phosphorylate with such probes including amino [7,8,9], hydroxyl [10,11,12,13,14,15,16], oxime [17,18,19,20,21] and pyridine [22,23,24] groups as the reaction sites. DCP can be reacted with a hydroxyl-ammonia [25,26] or carboxyl-ammonia [27] group of probes to produce cyclization products as well. The different structures of the probe and the corresponding product lead to the changes in fluorescent signals. However, most probes for DCP are fluorescence ON-OFF responses [31,32,33,34,35,36,37]. In addition, several probes detect DCP in organic solution [17,20,25,28], but a few probes detect DCP in aqueous solution (Table S1, Supplementary Materials) [18,38,39]. Therefore, the development of a turn-on fluorescent probe for DCP in an aqueous solution is still in high demand.
In our previous work, we synthesized a series of DCP probes based on ON-OFF fluorescent responses [40,41]. In the present work, a novel fluorescent probe SWJT-4 with two response sites was designed and synthesized for selective detection of DCP. It has good photochemical stability in an aqueous solution. The turn-on fluorescent responses could be achieved by tethering two oxime groups to a fluorescein skeleton. Moreover, SWJT-4 could be used to detect DCP vapor by a fluorescent test paper.

2. Results and Discussion

2.1. Design and Photoproperties of SWJT-4

According to the Duff reaction, two aldehyde groups were formed at the ortho positions of two hydroxyl groups of fluorescein [42]. Two oxime groups as the recognition sites were then added to the fluorophore to construct SWJT-4 (Scheme 1). The C = N rotation in SWJT-4 would weaken the fluorescence. With the addition of DCP, an isoxazole or cyano group would be formed, which would inhibit the rotation of C = N and therefore enhance the fluorescence. The corresponding 1H and 13C NMR spectra and ESI-MS of SWJT-4 are demonstrated in Figures S1–S3 (Supplementary Materials). In order to study its solvent effect on fluorescence properties of SWJT-4, five common organic solvents, namely acetone, methanol, tetrahydrofuran (THF), N,N-dimethyl formamide (DMF) and dimethyl sulfoxide (DMSO), were selected to test their performance under excitation at 520 nm. As shown in Figure S4 (Supplementary Materials), SWJT-4 in DMSO has a longer emission wavelength. After the addition of DCP, the fluorescence intensity was enhanced clearly. Subsequently, the impact of pH on SWJT-4 was also studied (Figure S5, Supplementary Materials). SWJT-4 did not react with DCP under acidic conditions, and the best reaction condition was pH 7.0 to 8.0. Therefore, DMSO–HEPES (1/1, v/v, pH 7.4) buffer solution was determined as the optimal condition.

Fluorescence Response of SWJT-4 to DCP

As shown in Figure 1a, the UV–visible absorption spectrum of SWJT-4 exhibited prominent absorption at 522 nm, which was attributed to the formation of intramolecular hydrogen bonds between the hydroxyl groups on fluorescein and the nitrogen atom on oxime, and the hydrogen bonds would enlarge the conjugation of the probe [43]. When DCP was added to the solution of the probe, the absorbance was blue-shifted to 508 nm, which is the absorption of the ring-opening structure of fluorescein [44]. These results showed that the reaction occurred between SWJT-4 and DCP to break the intramolecular hydrogen bonds. The color of the solution changed from pink to pale yellow (Figure 1a, inset). For the fluorescence spectrum, under the excitation of 520 nm, SWJT-4 showed weak emission at 557 nm (Φ = 5.7%) (Figure 1b). After the addition of DCP, the fluorescence was enhanced at 545 nm (Φ = 26.7%) with a slight blue-shift [45]. The fluorescence color of the solution was observed to turn chartreuse to green (Figure 1b, inset). These results showed that SWJT-4 was a turn-on fluorescent probe and could be used for the detection of DCP.
The fluorescence titration experiment of SWJT-4 was then studied. As shown in Figure 1c, the fluorescence intensity gradually increased with the increase in DCP. There was a good linear relationship between DCP concentration and the fluorescence intensity in the range of 0–140.0 μM (Figure 1d). The detection limit was calculated as 53.0 nM (Figure S6, Supplementary Materials), which was much lower than the reported lethal dose (0.01 mg/L) [46]. This result indicated that SWJT-4 had high sensitivity to DCP and could detect DCP at lower concentrations in an aqueous solution. Moreover, as shown in Figure S7 (Supplementary Materials), the recognition reaction finished within 100 min after the addition of DCP. The pseudo-first-order reaction constant (kobs) was 5.37 × 10−4 s−1. In order to verify the stoichiometry between the SWJT-4 and DCP, a Job plot analysis was performed (Figure S8, Supplementary Materials), which indicated that the ratio of SWJT-4 and DCP was 1:2. By Benesi–Hildebrand equation of this binding mode (1:2), a linear line with good linearity was obtained, and the binding constant was 1.37 × 106 M−2 (Figure S9, Supplementary Materials). These results were in good agreement with the stoichiometry between chemodosimeter and DCP.

2.2. Competition Experiments

In order to investigate the plausible interference of other organophosphorus reagents or nerve agent mimics on the detection of DCP, phosphoric acid (PA), cyanomethyl diethyl phosphate (DCMP), cyanoyl diethyl phosphate (DCNP) and ethyl dichlorophosphate (DCEP) were selected to study the selectivity of SWJT-4 (Figure 2). With the addition of other phosphate-containing substances, the fluorescence intensity of SWJT-4 changed little. However, after the addition of DCP to the solution of SWJT-4, the fluorescence at 545 nm was significantly enhanced. These results clearly showed that the fluorescent probe SWJT-4 could recognize DCP effectively even in the presence of other war agent mimics.

2.3. Response Mechanism

In order to identify the reaction mechanism between SWJT-4 and DCP, 1H NMR titration was performed. As shown in Figure 3a, the peaks at 11.90 ppm and 11.10 ppm belonged to the proton signals of two hydroxyl groups (Ha) on oxime moieties and two hydroxyl groups (Hb) on fluorescein in SWJT-4, respectively. The chemical shift of the proton (Hc) of the aldoxime group was at 8.80 ppm. After the addition of DCP, the original signals Ha and Hb in the probe disappeared, and one aldoxime proton Hc moved from 8.80 to 9.44 ppm downfield. The above results indicated that different reactions occurred at the two reaction sites of SWJT-4 to form product 2. One oxime group reacted with one DCP to form the nitrile [47], and the other oxime group reacted with another DCP to form an isoxazole ring. Firstly, the hydroxyl group in the oxime attacked the phosphorus center of DCP to form phosphate oxime. Then, another hydroxyl group in the adjacent position of the oxime group could intramolecularly attack this generated intermediate, and the isoxazole ring was then formed through a release of phosphate moiety [48]. Meanwhile, the mixture of SWJT-4 and DCP was also characterized by ESI-MS, and the peak at m/z 381.3 corresponding to product 2 was observed (Figure S10, Supplementary Materials).

2.4. Computational Studies

In order to further study the fluorescence responses of SWJT-4 with DCP and the influence of different solvent media on the absorption and emission, the B3LYP/6-31g method was conducted by Gaussian simulation (DFT) [49,50,51]. As shown in Figure 4, the main contribution for SWJT-4 was from HOMO ‒ 2 to LUMO + 1. The electron clouds of their HOMO ‒ 2 orbital in SWJT-4 were distributed on one benzene ring of the xanthrene group and the adjacent oxime group. The electron clouds of the LUMO orbital of SWJT-4 were mainly distributed on the whole molecule instead. These results suggested the weak fluorescence of SWJT-4. As for product 2, the main contribution was HOMO ‒ 2 to LUMO. The electron clouds of HOMO ‒ 2 and LUMO orbitals were all mainly distributed in the xanthrene moiety, indicating the strong fluorescence character of 2. These results were very consistent with the fluorescence turn-on change on the detection of DCP using SWJT-4. Then, the absorption and emission of SWJT-4 were calculated in DMSO or in water (Figure S11, Supplementary Materials). As shown in Figure S11a,c (Supplementary Materials), the maximum emission wavelength of SWJT-4 in DMSO or in water was about 287 nm and the maximum absorption wavelength of SWJT-4 in DMSO or in water was 280 nm (Figure S11b,d, Supplementary Materials). Although these wavelengths greatly differ from the measured results, the results indicated that the solvent effect had almost no effect on absorption and emission spectra in different solvent media.

2.5. Gas-Phase Detection of DCP

Considering its practical application, DCP vapors were used to determine the recognition ability of SWJT-4 (Figure 5a–d). The solution of SWJT-4 was placed in a glass bottle with a lid (Figure 5a,c). The color of the solution was orange-pink under visible light (Figure 5a) and chartreuse in ultraviolet light (Figure 5c). As a contrast, the solution of SWJT-4 was placed in another glass bottle with a lid, in which there was a smaller bottle containing DCP (Figure 5b,d). When the DCP vapors came into contact with the solution of SWJT-4, the color of the solution changed from orange-pink to yellow in visible light (Figure 5b), and the chartreuse color turned to green under ultraviolet light (Figure 5d). These results indicated that SWJT-4 could detect DCP vapor.
Moreover, the responses of SWJT-4 loaded on filter paper to different concentrations of DCP vapor were studied. With the increase in DCP concentration, the color of filter paper changed obviously under ultraviolet (Figure 5e) or visible light (Figure 5f), which was observed by the naked eye. The green color in paper sensors gradually increased under ultraviolet light (Figure 5e), and under visible light, the color of the paper sensors changed from pink to pale-yellow (Figure 5f). These results showed that the probe can be used as a fluorescent test paper to detect DCP vapor and has potential application in the development of detection kits for DCP.

3. Materials and Methods

3.1. Materials and Reagents

Related materials, reagents and the detail of detection are described in the Supplementary Materials.

3.2. Synthesis of Probe SWJT-4

The compound 1 was synthesized according to a known procedure [42]. Hydroxylamine hydrochloride (24.4 mg, 0.62 mmol) was dissolved in ethanol (3 mL) and stirred at room temperature for 10 min. Then compound 1 (60.4 mg, 0.16 mmol) was dispersed in 5 mL of ethanol and dripped into the above solution. The reaction mixture was stirred at room temperature for 2 h. The organic solvent was removed by rotary evaporation, and the crude product was isolated by column chromatography (dichloromethane:methanol = 80:1) on silica gel to obtain the probe SWJT-4 (54.2 mg, yield 82.9 %) as a pink solid. 1H NMR (400 MHz, DMSO-d6): δ = 11.90 (s, 2H), 11.10 (s, 2H), 8.80 (s, 2H), 8.02 (d, J = 7.6 Hz, 1H), 7.81 (t, J = 7.5 Hz, 1H), 7.74 (t, J = 7.0 Hz, 1H), 7.34 (d, J = 7.5 Hz, 1H), 6.74 (d, J = 8.8 Hz, 2H), 6.67 (d, J = 8.8 Hz, 2H). 13C NMR (100 MHz, DMSO-d6): δ = 169.0, 159.1, 152.5, 149.0, 145.6, 136.3, 130.8, 130.1, 126.4, 125.3, 124.5, 113.6, 110.2, 106.1, 82.6 ppm. ESI-MS: m/z 419.1 [M + H] +.

4. Conclusions

In summary, the fluorescence changes of a novel fluorescent probe SWJT-4 for the detection of DCP based on the dual reaction site were explored. SWJT-4 showed good selectivity for DCP and the obvious color change in an aqueous solution. Notably, the two reaction sites in the probe also triggered different reaction types. Moreover, SWJT-4 could be used for DCP vapor detection and as fluorescent test paper to detect DCP.

Supplementary Materials

The following supporting information can be downloaded online: Table S1: Some reported work for the detection of DCP. Figures S1–S3: Copies of 1H,13C NMR, ESI-MS spectra of SWJT-4. Figure S4: The changes of fluorescence in different solvents. Figure S5: The effect of pH on fluorescence intensity. Figure S6: The linear relationship between SWJT-4 and different concentrations of DCP. Figure S7: Time-dependent experiment of SWJT-4. Figures S8 and S9: Job’s and Benesi-Hildebrand plots of SWJT-4 with DCP. Figure S10: ESI-MS spectrum of SWJT-4 + DCP. Figure S11: The fluorescence spectrum and absorption spectrum of SWJT-4 calculated by DFT in DMSO and water.

Author Contributions

Y.-W.W. conceived and designed the experiments; X.G. and C.-X.L. performed the experiments; X.G., C.-X.L., Y.L., Y.-W.W. and Y.P. analyzed the data; X.G., Y.-W.W. and Y.P. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Nos. 21772078 and 22071200) and the Fundamental Research Funds for the Central Universities (Nos. 2682020CX55, 2682021ZTPY011 and XJ2021KJZK004) by MoE of China. We also thank the Science and Technology Department of Sichuan Province (2020JDRC0021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank the Analytical and Testing Center of Southwest Jiaotong University for the NMR test.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Walton, I.; Davis, M.; Munro, L.; Catalano, V.J.; Cragg, P.J.; Huggins, M.T.; Wallace, K.J. A Fluorescent Dipyrrinone Oxime for the Detection of Pesticides and Other Organophosphates. Org. Lett. 2012, 14, 2686–2689. [Google Scholar] [CrossRef]
  2. Muñoz-Quezada, M.T.; Lucero, B.A.; Iglesias, V.P.; Muñoz, M.P.; Cornejo, C.A.; Achu, E.; Baumert, B.; Hanchey, A.; Concha, C.; Brito, A.M.; et al. Chronic exposure to organophosphate (OP) pesticides and neuropsychological functioning in farm workers: A review. Int. J. Occup. Environ. Health 2016, 22, 68–79. [Google Scholar] [CrossRef] [Green Version]
  3. Stone, R. How to defeat a nerve agent. Science 2018, 359, 23. [Google Scholar] [CrossRef]
  4. Manco, G.; Porzio, E.; Suzumoto, Y. Enzymatic detoxification: A sustainable means of degrading toxic organophosphate pesticides and chemical warfare nerve agents. J. Chem. Technol. Biotechnol. 2018, 93, 2064–2082. [Google Scholar] [CrossRef]
  5. Chen, L.; Wu, D.; Yoon, J. Recent Advances in the Development of Chromophore-Based Chemosensors for Nerve Agents and Phosgene. ACS Sens. 2018, 3, 27–43. [Google Scholar] [CrossRef] [Green Version]
  6. Weis, J.G.; Swager, T.M. Thiophene-Fused Tropones as Chemical Warfare Agent-Responsive Building Blocks. ACS Macro Lett. 2015, 4, 138–142. [Google Scholar] [CrossRef] [Green Version]
  7. Gharami, S.; Aich, K.; Das, S.; Patra, L.; Mondal, T.K. Facile detection of organophosphorus nerve agent mimic (DCP) through a new quinoline-based ratiometric switch. New J. Chem. 2019, 43, 8627–8633. [Google Scholar] [CrossRef]
  8. Goswami, S.; Manna, A.; Paul, S. Rapid ‘naked eye’ response of DCP, a nerve agent simulant: From molecules to low-cost devices for both liquid and vapour phase detection. RSC Adv. 2014, 4, 21984–21988. [Google Scholar] [CrossRef]
  9. Zhou, X.; Zeng, Y.; Liyan, C. A Fluorescent Sensor for Dual-Channel Discrimination between Phosgene and a Nerve-Gas Mimic. Angew. Chem. 2016, 128, 4807–4811. [Google Scholar] [CrossRef]
  10. Xuan, W.; Cao, Y.; Zhou, J.; Wang, W. A FRET-based ratiometric fluorescent and colorimetric probe for the facile detection of organophosphonate nerve agent mimic DCP. Chem. Commun. 2013, 49, 10474–10476. [Google Scholar] [CrossRef]
  11. Lu, Z.; Fan, W.; Shi, X.; Black, C.A.; Fan, C.; Wang, F. A highly specific BODIPY-based fluorescent probe for the detection of nerve-agent simulants. Sens. Actuators B Chem. 2018, 255, 176–182. [Google Scholar] [CrossRef]
  12. Sarkar, H.S.; Ghosh, A.; Das, S.; Maiti, P.K.; Maitra, S.; Mandal, S.; Sahoo, P. Visualisation of DCP, a nerve agent mimic, in Catfish brain by a simple chemosensor. Sci. Rep. 2018, 8, 3402. [Google Scholar] [CrossRef]
  13. Singh, J.; Thakur, S.; Singh, R.; Kaur, V. Schiff base—Zn2+ ion combo as ‘pick and degrade’ probe for selected organophosphorus chemical weapon mimics and flame retardant analog: Detoxification of fruits and vegetables in aqueous media. Food Chem. 2020, 327, 127080. [Google Scholar] [CrossRef]
  14. Jang, Y.J.; Mulay, S.V.; Kim, Y.; Jorayev, P.; Churchill, D.G. Nerve agent simulant diethyl chlorophosphate detection using a cyclization reaction approach with high stokes shift system. New J. Chem. 2017, 41, 1653–1658. [Google Scholar] [CrossRef]
  15. Jang, Y.J.; Murale, D.P.; Churchill, D.G. Novel reversible and selective nerve agent simulant detection in conjunction with superoxide “turn-on” probing. Analyst 2014, 139, 1614–1617. [Google Scholar] [CrossRef] [Green Version]
  16. Kim, T.I.; Maity, S.B.; Bouffard, J.; Kim, Y. Molecular Rotors for the Detection of Chemical Warfare Agent Simulants. Anal. Chem. 2016, 88, 9259–9263. [Google Scholar] [CrossRef]
  17. Huo, B.; Du, M.; Shen, A.; Li, M.; Lai, Y.; Bai, X.; Gong, A.; Yang, Y. “Covalent-Assembly”-Based Fluorescent Probe for Detection of a Nerve-Agent Mimic (DCP) via Lossen Rearrangement. Anal. Chem. 2019, 91, 10979–10983. [Google Scholar] [CrossRef]
  18. Ali, S.S.; Gangopadhyay, A.; Pramanik, A.K.; Guria, U.N.; Samanta, S.K.; Mahapatra, A.K. Ratiometric sensing of nerve agent mimic DCP through in situ benzisoxazole formation. Dyes. Pigm. 2019, 170, 107585. [Google Scholar] [CrossRef]
  19. Cai, Y.C.; Li, C.; Song, Q.H. Selective and visual detection of a nerve agent mimic by phosphorylation and protonation of quinolin oximes. J. Mater. Chem. C 2017, 5, 7337–7343. [Google Scholar] [CrossRef]
  20. Guria, U.N.; Maiti, K.; Ali, S.S.; Gangopadhyay, A.; Samanta, S.K.; Roy, K.; Mandal, D.; Mahapatra, A.K. An Organic Nanofibrous Polymeric Composite for Ratiometric Detection of Diethyl Chlorophosphate (DCP) in Solution and Vapor. ChemistrySelect 2020, 5, 3770–3777. [Google Scholar] [CrossRef]
  21. Kim, Y.; Jang, Y.J.; Mulay, S.V.; Nguyen, T.T.T.; Churchill, D.G. Fluorescent Sensing of a Nerve Agent Simulant with Dual Emission over Wide pH Range in Aqueous Solution. Chem. Eur. J. 2017, 23, 7785–7790. [Google Scholar] [CrossRef]
  22. Kim, Y.; Jang, Y.J.; Lee, D.; Kim, B.S.; Churchill, D.G. Real nerve agent study assessing pyridyl reactivity: Selective fluorogenic and colorimetric detection of Soman and simulant. Sens. Actuators B 2017, 238, 145–149. [Google Scholar] [CrossRef]
  23. Cai, Y.C.; Li, C.; Song, Q.H. Fluorescent Chemosensors with Varying Degrees of Intramolecular Charge Transfer for Detection of a Nerve Agent Mimic in Solutions and in Vapor. ACS Sens. 2017, 2, 834–841. [Google Scholar] [CrossRef]
  24. Huang, S.; Wu, Y.; Zeng, F.; Sun, L.; Wu, S. Handy ratiometric detection of gaseous nerve agents with AIE-fluorophore-based solid test strips. J. Mater. Chem. C 2016, 4, 10105–10110. [Google Scholar] [CrossRef]
  25. Wu, W.H.; Dong, J.J.; Wang, X.; Li, J.; Sui, S.H.; Chen, G.Y.; Liu, J.W.; Zhang, M. Fluorogenic and chromogenic probe for rapid detection of a nerve agent simulant DCP. Analyst 2012, 137, 3224–3226. [Google Scholar] [CrossRef]
  26. Barba-Bon, A.; Costero, A.M.; Gil, S.; Harriman, A.; Sancenón, F. Highly Selective Detection of Nerve-Agent Simulants with BODIPY Dyes. Chem. Eur. J. 2014, 20, 6339–6347. [Google Scholar] [CrossRef]
  27. Hu, X.X.; Su, Y.T.; Ma, Y.W.; Zhan, X.Q.; Zheng, H.; Jiang, Y.-B. A near infrared colorimetric and fluorometric probe for organophosphorus nerve agent mimics by intramolecular amidation. Chem. Commun. 2015, 51, 15118–15121. [Google Scholar] [CrossRef]
  28. So, H.S.; Angupillai, S.; Son, Y.A. Prompt liquid-phase visual detection and low-cost vapor-phase detection of DCP, a chemical warfare agent mimic. Sens. Actuators B 2016, 235, 447–456. [Google Scholar] [CrossRef]
  29. Mahapatra, A.K.; Maiti, K.; Manna, S.K.; Maji, R.; Mondal, S.; Das Mukhopadhyay, C.; Sahoo, P.; Mandal, D. A cyclization-induced emission enhancement (CIEE)-based ratiometric fluorogenic and chromogenic probe for the facile detection of a nerve agent simulant DCP. Chem. Commun. 2015, 51, 9729–9732. [Google Scholar] [CrossRef] [PubMed]
  30. Goswami, S.; Das, S.; Aich, K. Fluorescent chemodosimeter based on spirobenzopyran for organophosphorus nerve agent mimics (DCP). RSC Adv. 2015, 5, 28996–29001. [Google Scholar] [CrossRef]
  31. Patra, L.; Ghosh, P.; Das, S.; Gharami, S.; Murmu, N.; Mondal, T.K. A selective fluorogenic chemosensor for visual detection of chemical warfare reagent mimic diethylchlorophosphate. J. Photochem. Photobiol. A Chem. 2020, 388, 112188. [Google Scholar] [CrossRef]
  32. Heo, G.; Manivannan, R.; Kim, H.; Son, Y.A. Liquid and gaseous state visual detection of chemical warfare agent mimic DCP by optical sensor. Dyes. Pigm. 2019, 171, 107712. [Google Scholar] [CrossRef]
  33. Ali, S.S.; Gangopadhyay, A.; Pramanik, A.K.; Samanta, S.K.; Guria, U.N.; Manna, S.; Mahapatra, A.K. Real time detection of the nerve agent simulant diethylchlorophosphate by nonfluorophoric small molecules generating a cyclization-induced fluorogenic response. Analyst 2018, 143, 4171–4179. [Google Scholar] [CrossRef] [PubMed]
  34. Fu, Y.; Yu, J.; Wang, K.; Liu, H.; Yu, Y.; Liu, A.; Peng, X.; He, Q.; Cao, H.; Cheng, J. Simple and Efficient Chromophoric-Fluorogenic Probes for Diethylchlorophosphate Vapor. ACS Sens. 2018, 3, 1445–1450. [Google Scholar] [CrossRef] [PubMed]
  35. Suhasini, R.; Karpagam, R.; Thirumoorthy, K.; Thiagarajan, V. “Turn-on” unsymmetrical azine based fluorophore for the selective detection of diethylchlorophosphate via photoinduced electron transfer to intramolecular charge transfer pathway. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2021, 263, 120206. [Google Scholar] [CrossRef]
  36. Zhang, S.; Zhou, C.; Yang, B.; Zhao, Y.; Wang, L.; Yuan, B.; Li, H. Rhodamine phenol-based fluorescent probe for the visual detection of GB and its simulant DCP. New J. Chem. 2021, 45, 7564–7570. [Google Scholar] [CrossRef]
  37. Zheng, P.; Cui, Z.; Liu, H.; Cao, W.; Li, F.; Zhang, M. Ultrafast-response, highly-sensitive and recyclable colorimetric/fluorometric dual-channel chemical warfare agent probes. J. Hazard Mater. 2021, 415, 125619. [Google Scholar] [CrossRef]
  38. Kim, H.J.; Jang, S.; Ren, W.X.; Bartsch, R.A.; Sohn, H.; Kim, J.S. Imine-functionalized, turn-on fluorophore for DCP. Sens. Actuators B 2011, 153, 266–270. [Google Scholar] [CrossRef]
  39. Ali, S.S.; Gangopadhyay, A.; Maiti, K.; Mondal, S.; Pramanik, A.K.; Guria, U.N.; Uddin, M.R.; Mandal, S.; Mandal, D.; Mahapatra, A.K. A chromogenic and ratiometric fluorogenic probe for rapid detection of a nerve agent simulant DCP based on a hybrid hydroxynaphthalene–hemicyanine dye. Org. Biomol. Chem. 2017, 15, 5959–5967. [Google Scholar] [CrossRef]
  40. Khan, M.S.J.; Wang, Y.W.; Senge, M.O.; Peng, Y. Sensitive fluorescence on-off probes for the fast detection of a chemical warfare agent mimic. J. Hazard. Mater. 2018, 342, 10–19. [Google Scholar] [CrossRef]
  41. Feng, X.; Wang, Y.; Feng, W.; Peng, Y. Development of BINOL-Si complexes with large stokes shifts and their application as chemodosimeters for nerve agent. Chin. Chem. Lett. 2020, 31, 2960–2964. [Google Scholar] [CrossRef]
  42. Xiong, X.; Song, F.; Sun, S.; Fan, J.; Peng, X. Red-Emissive Fluorescein Derivatives and Detection of Bovine Serum Albumin. Asian J. Org. Chem. 2013, 2, 145–149. [Google Scholar] [CrossRef]
  43. Zhao, L.; Dong, X.T.; Zhang, S.T.; Chen, Q. Crystal structure of 1-(4-{[(E)-5-bromo-2-hydroxybenzylidene]-amino} phenyl) ethanone oxime, C15H13BrN2O2. Z. Kristallogr. NCS 2012, 227, 101–102. [Google Scholar] [CrossRef]
  44. Sjöback, R.; Nygren, J.; Kubista, M. Absorption and fluorescence properties of fluorescein. Spectrochim Acta A Mol. Biomol. Spectrosc. 1995, 51, L7–L21. [Google Scholar] [CrossRef]
  45. Melhuish, W.H. Quantum efficiencies of fluorescence of organic substances: Effect of solvent and concentration of the fluorescent solute. J. Phys. Chem. 1961, 65, 229–235. [Google Scholar] [CrossRef]
  46. Sadik, O.A.; Land, W.H., Jr.; Wang, J. Targeting Chemical and Biological Warfare Agents at the Molecular Level. Electroanalysis 2003, 15, 1149–1159. [Google Scholar] [CrossRef]
  47. Lee, H.; Kim, H.-J. Novel fluorescent probe for the selective detection of organophosphorous nerve agents through a cascade reaction from oxime to nitrile via isoxazole. Tetrahedron 2014, 70, 2966–2970. [Google Scholar] [CrossRef]
  48. Radić, Z.; Dale, T.; Kovarik, Z.; Berend, S.; Garcia, E.; Zhang, L.; Amitai, G.; Green, C.; Radić, B.; Duggan, B.M.; et al. Catalytic detoxification of nerve agent and pesticide organophosphates by butyrylcholinesterase assisted with non-pyridinium oximes. Biochem. J. 2013, 450, 231–242. [Google Scholar] [CrossRef] [Green Version]
  49. 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] [PubMed] [Green Version]
  50. Becke, A.D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  51. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of SWJT-4.
Scheme 1. Synthesis of SWJT-4.
Molecules 27 00489 sch001
Figure 1. (a) The absorption spectrum of SWJT-4 (10.0 µM) and SWJT-4 + DCP (1.0 mM) in DMSO– HEPES (1/1, v/v, pH 7.4) buffer solution. Inset: the images of SWJT-4 and SWJT-4 after addition of DCP under visible light. (b) Fluorescence spectrum of SWJT-4 (10.0 µM) and SWJT-4 + DCP (1.0 mM) in DMSO–HEPES (1/1, v/v, pH 7.4) buffer solution. Inset: the images of SWJT-4 and SWJT-4 after addition of DCP under ultraviolet light. (c) Fluorescence titrations of SWJT-4 (10.0 µM) with different concentrations of DCP (0–1000.0 µM); (d) Linear relationship between the fluorescence intensity of SWJT-4 (10.0 µM) at 545 nm and the concentration of DCP (0–140.0 μM).
Figure 1. (a) The absorption spectrum of SWJT-4 (10.0 µM) and SWJT-4 + DCP (1.0 mM) in DMSO– HEPES (1/1, v/v, pH 7.4) buffer solution. Inset: the images of SWJT-4 and SWJT-4 after addition of DCP under visible light. (b) Fluorescence spectrum of SWJT-4 (10.0 µM) and SWJT-4 + DCP (1.0 mM) in DMSO–HEPES (1/1, v/v, pH 7.4) buffer solution. Inset: the images of SWJT-4 and SWJT-4 after addition of DCP under ultraviolet light. (c) Fluorescence titrations of SWJT-4 (10.0 µM) with different concentrations of DCP (0–1000.0 µM); (d) Linear relationship between the fluorescence intensity of SWJT-4 (10.0 µM) at 545 nm and the concentration of DCP (0–140.0 μM).
Molecules 27 00489 g001
Figure 2. The fluorescence responses of (a) SWJT-4 (10.0 μM) with different war agent simulants (1.0 mM). From left to right: SWJT-4, phosphoric acid (PA), cyanomethyl diethyl phosphate (DCMP), cyanoyl diethyl phosphate (DCNP), ethyl dichlorophosphate (DCEP) and diethyl chlorophosphate (DCP); (b) SWJT-4 (10.0 μM) with different war agent simulants (1.0 mM) and DCP (1.0 mM) in DMSO–HEPES (1/1, v/v, pH 7.4) buffer solution (1, phosphoric acid; 2, cyanomethyl diethyl phosphate; 3, cyanoyl diethyl phosphate; 4, ethyl dichlorophosphate). Black bar: SWJT-4 + war agent simulants; red bar: SWJT-4 + war agent simulants + DCP.
Figure 2. The fluorescence responses of (a) SWJT-4 (10.0 μM) with different war agent simulants (1.0 mM). From left to right: SWJT-4, phosphoric acid (PA), cyanomethyl diethyl phosphate (DCMP), cyanoyl diethyl phosphate (DCNP), ethyl dichlorophosphate (DCEP) and diethyl chlorophosphate (DCP); (b) SWJT-4 (10.0 μM) with different war agent simulants (1.0 mM) and DCP (1.0 mM) in DMSO–HEPES (1/1, v/v, pH 7.4) buffer solution (1, phosphoric acid; 2, cyanomethyl diethyl phosphate; 3, cyanoyl diethyl phosphate; 4, ethyl dichlorophosphate). Black bar: SWJT-4 + war agent simulants; red bar: SWJT-4 + war agent simulants + DCP.
Molecules 27 00489 g002
Figure 3. Partial 1H NMR spectra of (a) SWJT-4 and (b) SWJT-4 + DCP in DMSO-d6.
Figure 3. Partial 1H NMR spectra of (a) SWJT-4 and (b) SWJT-4 + DCP in DMSO-d6.
Molecules 27 00489 g003
Figure 4. The optimized structures and the molecular orbital plots of SWJT-4 and 2.
Figure 4. The optimized structures and the molecular orbital plots of SWJT-4 and 2.
Molecules 27 00489 g004
Figure 5. The color changes of (a,c) SWJT-4 (10.0 M) (DMSO–HEPES, 1/1, v/v, pH 7.4) solution and (b,d) exposure to DCP vapors under visible light or ultraviolet light. Paper sensors that visually detect different concentrations of DCP under ultraviolet light (e) and visible light (f). From left to right: [DCP] = 0 mM, 1.0 × 10−3 mM, 1.0 × 10−2 mM, 1.0 × 10−1 mM, 1.0 mM.
Figure 5. The color changes of (a,c) SWJT-4 (10.0 M) (DMSO–HEPES, 1/1, v/v, pH 7.4) solution and (b,d) exposure to DCP vapors under visible light or ultraviolet light. Paper sensors that visually detect different concentrations of DCP under ultraviolet light (e) and visible light (f). From left to right: [DCP] = 0 mM, 1.0 × 10−3 mM, 1.0 × 10−2 mM, 1.0 × 10−1 mM, 1.0 mM.
Molecules 27 00489 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, X.; Liu, C.-X.; Lu, Y.; Wang, Y.-W.; Peng, Y. A Double-Site Chemodosimeter for Selective Fluorescence Detection of a Nerve Agent Mimic. Molecules 2022, 27, 489. https://doi.org/10.3390/molecules27020489

AMA Style

Guo X, Liu C-X, Lu Y, Wang Y-W, Peng Y. A Double-Site Chemodosimeter for Selective Fluorescence Detection of a Nerve Agent Mimic. Molecules. 2022; 27(2):489. https://doi.org/10.3390/molecules27020489

Chicago/Turabian Style

Guo, Xin, Chang-Xiang Liu, Yuan Lu, Ya-Wen Wang, and Yu Peng. 2022. "A Double-Site Chemodosimeter for Selective Fluorescence Detection of a Nerve Agent Mimic" Molecules 27, no. 2: 489. https://doi.org/10.3390/molecules27020489

Article Metrics

Back to TopTop