Next Article in Journal
Kukhtin–Ramirez-Reaction-Inspired Deprotection of Sulfamidates for the Synthesis of Amino Sugars
Next Article in Special Issue
Ultrarapid Microwave-Assisted Synthesis of Fluorescent Silver Coordination Polymer Nanoparticles and Its Application in Detecting Alkaline Phosphatase Activity
Previous Article in Journal
Upcycling Rocha do Oeste Pear Pomace as a Sustainable Food Ingredient: Composition, Rheological Behavior and Microstructure Alone and Combined with Yeast Protein Extract
Previous Article in Special Issue
Recent Progress in Identifying Bacteria with Fluorescent Probes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Highly Efficient Fluorescent Sensor Based on AIEgen for Detection of Nitrophenolic Explosives

1
Henan Key Laboratory of Function-Oriented Porous Materials, College of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang 471000, China
2
AIE Research Center, College of Chemistry and Chemical Engineering, Baoji University of Arts and Sciences, Baoji 721013, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(1), 181; https://doi.org/10.3390/molecules28010181
Submission received: 17 November 2022 / Revised: 20 December 2022 / Accepted: 21 December 2022 / Published: 25 December 2022

Abstract

:
The detection of nitrophenolic explosives is important in counterterrorism and environmental protection, but it is still a challenge to identify the nitroaromatic compounds among those with a similar structure. Herein, a simple tetraphenylethene (TPE) derivative with aggregation-induced emission (AIE) characteristics was synthesized and used as a fluorescent sensor for the detection of nitrophenolic explosives (2, 4, 6-trinitrophenol, TNP and 2, 4-dinitrophenol, DNP) in water solution and in a solid state with a high selectivity. Meanwhile, it was found that only hydroxyl containing nitrophenolic explosives caused obvious fluorescence quenching. The sensing mechanism was investigated by using fluorescence titration and 1H NMR spectra. This simple AIE-active probe can potentially be applied to the construction of portable detection devices for explosives.

1. Introduction

The effective detection of nitroaromatic compounds (NACs) is of great importance in modern society owing to their threats to both national security and environmental safety. Different nitroaromatic explosives such as 2, 4, 6-Trinitrophenol (TNP, picric acid) and 2, 4-dinitrophenol (DNP) are broadly utilized in the assembling of rocket fills, firecrackers, matches, etc. Additionally, TNP and its derivatives are also used in the field of medical formulation as an antiseptic agent, as well as being used as a yellow pigment in dye and leather industries, which will cause environmental pollution and health hazards [1,2]. Consequently, the development of highly selective and sensitive sensors with a quick response for discriminating TNP and DNP is imperative for natural remediation, regular citizen security, and military tasks.
Presently, several analytical techniques have been developed for the detection of nitroaromatic explosives, such as trained canines [3], mass spectrometry [4], gas chromatography [5,6], liquid chromatography [7], ion mobility spectrometry [8], electrochemical assay [9], high performance liquid chromatography [10], X-ray imaging, Raman spectroscopy [11], and so on. In contrast to the above-mentioned methods, fluorescent probes with the merits of high sensitivity, selectivity, easy visualization, and a short response time have attracted great attention in various areas [12,13,14,15,16]. To date, numerous fluorescent sensors based on polymers [17,18], metal–organic frameworks [19], fluorescent quantum dots [20,21], and organic small molecules [13] have been designed and developed. Despite these advances, most of the traditional fluorophores have often encountered the aggregation-caused quenching (ACQ) problem in the aggregated state, which has been a hurdle to fluorescence sensing. Moreover, selective detection of nitroaromatic explosives remains difficult due to their similar electron affinity.
Since aggregation-induced emission (AIE) was termed by Tang in 2001 [22], a large number of AIE luminogens (AIEgens) have been developed for photoelectric devices, biomaterials and fluorescent probes [23,24,25,26,27]. Thanks to their high sensitivity and solid-state emission efficiency, detection of explosives based on AIE-active probes has received a lot of interest [28,29,30,31,32]. Among these AIEgens, tetraphenylethene (TPE) is the most popular building block for the construction of fluorescence sensors due to its high photostability, simple synthesis, and easy structure modification. For example, Tang and their co-workers reported that a series of TPE-based polymers can serve as highly sensitive sensors for TNP through a mode of emission quenching with this substrate [33,34,35,36]. To provide the fluorescence sensors with better selectivity, the Zheng group designed several TPE-based macrocycles showing a high selectivity for NACs, alongside the excellent sensitivity by virtue of the encapsulation [37,38,39]. Throughout the course of our continuous effort to develop an excellent fluorescence probe for explosives, herein a simple TPE-based AIEgen was synthesized to detect TNP from a series of NACs. Such material showed highly sensitive fluorescence quenching to TNP in aqueous solution and in a solid state by virtue of photo-induced electron transfer. This simple AIE-active probe can potentially be applied to the construction of portable detection devices for explosives.

2. Results and Discussion

The synthetic procedure of the target compound 3 was designed and presented in Scheme 1. The known tetraphenylethene derivative 1 was utilized as the starting material. It was then nitrated by concentrated nitric acid and acetic acid in dichloromethane to afford molecule 2. By a reduction of 2 with hydrazine hydrate and Pd/C in ethanol, the target molecule 3 was obtained in a good yield of 87%. The molecule 3 was fully characterized by 1H NMR, 13C NMR, high resolution mass spectra (HRMS), and infrared (IR) spectra in the Supporting Information (Figures S1–S4).
The photophysical properties of compound 3 were then investigated. The UV/Vis absorption spectra of 3 responding to TNP were measured in dimethylsulfoxide (DMSO). The molecule 3 showed absorption bands at 342, 292 and 261 nm. When adding TNP to 3, the absorption intensity largely increased, but the algebraic sum of spectrum of 3 and the spectrum of TNP is similar to the spectrum of the mixture of 3 + TNP, implying that there were weak interactions in the solution state (Figure S5). From the photoluminescence (PL) spectra we can see that compound 3 was non-emissive in THF (Figure 1). Upon the addition of poor solvent water to THF, the solution remained weak in fluorescence from 0 to 80% water fraction. Upon further enhancing the water fraction to 92%, compound 3 emitted bright blue fluorescence at 478 nm. The 58-fold enhancement of PL intensity indicated that compound 3 is an AIE-active compound. The scanning electron microscope (SEM) images revealed that the turbid solution of 3 in 92% water fraction was composed of many rod-like aggregates with the length of micrometers (Figure S6).
The fluorescence response of molecule 3 to NACs was then studied in H2O/THF (9: 1, v/v). As shown in Figure 2A, molecule 3 (20 μM) fluoresced strongly in 90% water fraction without NACs. When different nitrophenolic explosives (40 μM) were added into the solution, such as 2,4,6-trinitrophenol (TNP), 2,4-dinitrophenol (DNP), p-nitrophenol (PNP), o-nitrophenol (ONP), 2,4,6-trinitromethylbenzene (TNT), 2,4-dinitrotoluene (DNT), p-nitrotoluene (PNT), o-nitrotoluene (ONT), nitrobenzene (NB), 1,3-dinitrobenzene (DNB), 1-fluoro-2,4-dinitrobenzene (DNFB), 4-hydroxyisophthalonitrile (HPN), 2,4-dinitrochlorobenzene (DNCB), 3-cyanopheno (CP), phenol, 3,5-dinitrobenzoic acid (DNA), and 2-Hydroxybenzonitrile (HBN), the sharp quenching caused by TNP and DNP was observed under a 365 nm UV light. Meanwhile, it was found that only hydroxyl containing nitrophenolic explosives caused obvious fluorescence quenching, while other NACs showed a minor influence on the emission of molecule 3. The quenching efficiency of PNP, ONP, DNP, and TNP increased with the enhancement of acidity. Therefore, it is inferred that the existence of electrostatic interactions led to the high selectivity of molecule 3 to TNP and DNP. The quenching efficiency ((1 − I/I0) × 100%) of compound 3 for TNP and DNP were 95% and 79%, respectively (Figure 2B and Figure S7).
The fluorescence titration of 3 (20 μM) with different equivalents of TNP was measured to verify their complexation ratio. As shown in Figure 3A, the PL intensity of molecule 3 at 460 nm gradually decreased when the TNP concentration was increased to 40 μM. The stoichiometry between compound 3 and TNP was calculated from Job’s plot to be 1: 2 (Figure 3B). Based on the data presented in Figure S8, the association constant was estimated to be 3.4 × 108 from linear curve fitting using Origin software [40,41]. Similar results were obtained from the fluorescence titrations of 3 with DNP (Figure 3C). The stoichiometry from Job’s plot was estimated to be 1:2 also (Figure 3D) and the association constant was calculated to be 7.4 × 107 (Figure S9). From the association constants we can conclude that molecule 3 has much a stronger affinity to TNP than DNP. Due to the higher number of electron-withdrawing nitro groups in TNP, it exhibited much more acidity than other nitrophenolic compounds. Thus, it is easy to form complexes with molecule 3 to occur photo-induced electron transfer within the complex resulting in fluorescence quenching. Then fluorescence lifetimes of compound 3, TNP, and the mixture of compound 3 + TNP were measured in an aqueous solution to verify their interactions. As illustrated in Figure S10, compound 3 gave a lifetime of 4.52 ns. Upon the addition of TNP to the system, the lifetime reduced to 0.35 ns, indicating the existence of static fluorescence quenching.
1H NMR titration in d6-DMSO was performed again to gain a deeper understanding of the bond formation between compound 3 and TNP. In fact, we made an attempt to record the NMR titration in the mixture of deuterated water and THF with 90% deuterated water fraction. Unfortunately, the compound 3 and TNP could not dissolve in the mixed solvents. Thus, we carried out 1H NMR titration in d6-DMSO. As shown in Figure 4A, when two equivalents of TNP were added, the chemical shift Ha of compound 3 showed a downfield shift from 6.26 to 7.01 ppm. Upon further addition of TNP, the chemical shift was only slightly altered. Additionally, there was also an obvious downfield shift of Hb in compound 3 from 6.58 to 7.02 ppm. These results proved that the electrostatic interaction between amino groups and hydroxyl group mainly contributes to the binding between TNP and compound 3. 1H NMR titration indicated that a 1:2 binding ratio was obtained between compound 3 and TNP as shown in Figure S11. It is worth noting that this binding ratio was obtained from a d6-DMSO concentrated solution (5 mM) in 1H NMR titration and it is not clear if this will also be true with its diluted solution in H2O: THF (90% water fraction), with concentration of the order 10−5 M. However, from the results of 1H NMR and fluorescence titration we can see that they obtained the same complexation ratio. Additionally, Figure 4B illustrates a possible binding mode between compound 3 and TNP.
Furthermore, fluorescence detection of TNP was also estimated in the solid state (Figure 5). Compound 3 was dripped onto the TLC plate and dried under a vacuum. The TLC plate with pure compound 3 showed bright fluorescence. When different concentrations of TNP solution were dripped on the spot of compound 3, the fluorescence was quickly quenched. From the picture we can see that 0.2 μM of compound 3 can cause obvious fluorescence quenching on the TLC plate. The data verified that compound 3 is an excellent sensor for TNP not only in the solution, but also in the solid state.

3. Materials and Methods

Materials: All reagents and solvents were chemical pure (CP) grade or analytical reagent (AR) grade and were used as received.
Measurements: 1H and 13C NMR were measured on 400 MHz Bruker Advanced III (Bremen, Germany). Mass spectrum was measured on Waters instrument (Framingham, MA, USA). IR was measured on Bruker VERTEX70 (Bremen, Germany). Fluorescent spectra were collected on Hitachi F-4500 spectrophotometer (Tokyo, Japan).

3.1. Synthesis of 2

The mixture of 30 mL dichloromethane, 2.85 mL (40.9 mmol) concentrated nitric acid, and 2.34 mL (20.8 mmol) acetic acid was added to a 100 mL drying flask. The flask was placed in a low temperature reaction bath at –15 °C, 2 g 4, 4’-dimethoxytetraphenylethene (5 mmol) was added to the mixture under stirring for 0.5 h. The reaction mixture was taken out of the low-temperature reaction bath and returned to room temperature. The product was quenched by adding 20 mL of water. The organic phase was separated and washed with water several times until the pH was 7. Compound 2 was obtained as a yellow solid by evaporating the solvent using a rotary evaporator (1.95 g, 81%).

3.2. Synthesis of 3

Compound 2 (1 g, 2.1 mmol), 30 mL of absolute ethanol, hydrazine hydrate (1 mL, 21 mmol), and Pd/C (87 mg, 0.42), respectively, were added to the flask. After two hours of refluxing, the mixture was cooled to room temperature. The organic phase was obtained after filtration and then concentrated with a rotary evaporator. An ethyl acetate/petroleum ether 1/1 eluent was used to purify the crude product with a silica gel column. Compound 3 was obtained as a yellow powder with an 87% yield (0.77 g). Mp 216.4–218.6 °C; IR (KBr) ν 3460, 3414, 3363, 3024, 2835, 1605, 1508, 1443, 1238, 1172, 1026, 833 cm−1; 1H NMR (400 MHz, DMSO) δ 6.46 (d, J = 8.4 Hz, 4H), 6.63 (d, J = 8.4 Hz, 4H), 6.56 (d, J = 8.0 Hz, 4H), 6.24 (d, J = 8.0 Hz, 4H), 4.92 (s, 4H), 3.64 (s, 6H) ppm; 13CNMR (100 MHz, DMSO) δ: 156.9, 146.7, 139.9, 137.4, 134.6, 132.0, 131.7, 131.6, 113.1, 113.0 ppm; MS m/z calcd for C28H26N2O2 422.5 [M], found 423.2055 [M+].

4. Conclusions

In conclusion, this paper presented the design and synthesis of an aggregation-induced emission fluorescent sensor for a nitrophenolic explosives in aqueous media, which was highly effective. The fluorescence quenching of the sensor was observed due to the formation of the 1: 2 complex with TNP/DNP. The binding constants of compound 3 with TNP and DNP were calculated to be 3.4 × 108 and 7.4 × 107, respectively. This simple sensor exhibited high potential for the detection of TNP and DNP both in water and in the solid state.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28010181/s1, Figures S1–S4: characteristic spectra; Figure S5: UV/vis spectra; Figure S6: SEM images; Figures S7–S11: fluorescence spectra.

Author Contributions

Conceptualization, D.L. and H.-T.F.; methodology, D.L. and P.L.; validation, X.-W.H., Z.J. and M.Z.; writing—original draft preparation, D.L.; writing—review and editing, H.-T.F. 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 (52173152, 21805002), Guangdong Basic and Applied Basic Research Foundation (2020A1515110476), the Fund of the Rising Stars of Shaanxi Province (2021KJXX-48), Scientific Research Program Funded by Shaanxi Provincial Education Department (22JK0247), Scientific and Technological Innovation Team of Shaanxi Province (2022TD-36).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Patil, G.; Dongre, S.-D.; Das, T. Sukumaran. S.-B. Dual Mode Selective Detection and Differentiation of TNT from Other Nitroaromatic Compounds. J. Mater. Chem. A 2020, 8, 10767–10771. [Google Scholar]
  2. Shanmugaraju, S.; Dabadie, C.; Byrne, K.; Savyasachi, A.-J.; Umadevi, D.; Schmitt, W.; Kitchen, J.-A.; Gunnlaugsson, T. A supramolecular Tröger’s base derived coordination zinc polymer for fluorescent sensing of phenolic-nitroaromatic explosives in water. Chem. Sci. 2017, 8, 1535–1546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Danquah, M.-K.; Wang, S.; Wang, Q.-Y.; Wang, B.; Wilson, L.-D. A porous b-cyclodextrin-based terpolymer fluorescence sensor for in situ trinitrophenol detection. RSC. Adv. 2019, 9, 8073–8080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Forbes, T.-P.; Sisco, E. Recent advances in ambient mass spectrometry of trace explosives. Analyst 2018, 143, 1948–1969. [Google Scholar] [CrossRef]
  5. Abhiram, P.; Basanta, P.-S.; Sonam, M.; Debasis, N.; Santanab, G.; Tridib, K.-S. AIE active fluorescent organic nanoaggregates for selective detection of phenolic-nitroaromatic explosives and cell imaging. J. Photoch. Photobio. A 2019, 374, 194–205. [Google Scholar]
  6. Marder, D.; Tzanani, N.; Prihed, H.; Gura, S. Trace detection of explosives with a unique large volume injection gas chromatography-mass spectrometry (LVI-GC-MS) method. Anal. Methods 2018, 10, 2712–2721. [Google Scholar] [CrossRef]
  7. Mu, R.; Shi, H.; Yuan, Y.; Karnjanapiboonwong, A.; Burken, J.-G.; Ma, Y. Fast separation and quantification method for nitroguanidine and 2,4-dinitroanisole by ultrafast liquid chromatography–tandem mass spectrometry. Anal. Chem. 2012, 84, 3427–3432. [Google Scholar] [CrossRef]
  8. Najarro, M.; Dávila Morris, M.-E.; Staymates, M.-E.; Fletcher, R.; Gillen, G. Optimized thermal desorption for improved sensitivity in trace explosives detection by ion mobility spectrometry. Analyst 2012, 137, 2614–2622. [Google Scholar] [CrossRef] [PubMed]
  9. Singh, S.; Meena, V.-K.; Mizaikoff, B.; Singh, S.-P.; Suri, C.-R. Electrochemical sensing of nitro-aromatic explosive compounds using silver nanoparticles modified electrochips. Anal. Methods 2016, 8, 7158–7169. [Google Scholar] [CrossRef]
  10. Babaee, S.; Beiraghi, A. Micellar extraction and high performance liquid chromatography-ultra violet determination of some explosives in water samples. Anal. Chim. Acta 2010, 662, 9–13. [Google Scholar] [CrossRef]
  11. Jha, S.-K.; Ekinci, Y.; Agio, M.; Löffler, J.-F. Towards deep-UV surface-enhanced resonance Raman spectroscopy of explosives: Ultrasensitive, real-time and reproducible detection of TNT. Analyst 2015, 140, 5671–5677. [Google Scholar] [CrossRef]
  12. Salinas, Y.; Martinez-Manez, R.; Marcos, M.-D.; Sancenon, F.; Costero, A.-M.; Parra, M.; Gil, S. Optical chemosensors and reagents to detect explosives. Chem. Soc. Rev. 2012, 41, 1261–1296. [Google Scholar] [CrossRef] [PubMed]
  13. Shanmugaraju, S.; Mukherjee, P.-S. π-Electron rich small molecule sensors for the recognition of nitroaromatics. Chem. Commun. 2015, 51, 16014–16032. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, X.; Wang, Y.; Lei, Y. Fluorescence based explosive detection: From mechanisms to sensory materials. Chem. Soc. Rev. 2015, 44, 8019–8061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wu, X.-X.; Fu, H.-R.; Han, M.-L.; Zhou, Z.; Ma, L.-F. Tetraphenylethylene immobilized metal-organic frameworks: Highly sensitive fluorescent sensor for the detection of Cr2O72− and nitroaromatic explosives. Cryst. Growth Des. 2017, 17, 6041–6048. [Google Scholar] [CrossRef]
  16. Dang, R.; Ma, X.-R.; Zhao, Y.-P.; Ren, M.; Guo, W.; Kang, Y.-H.; Gao, Y.; Bi, S.; Gao, W.; Hao, H.-R.; et al. Construction of 3D Ni2+-Al3+-LDH/γ-Fe2O3-Cd2+-Ni2+-Fe3+ -LDH structures and multistage recycling treatment of dye and fluoride-containing wastewater. Chem. Eng. J. 2023, 451, 138499. [Google Scholar] [CrossRef]
  17. Nabeel, F.; Rasheed, T.; Mahmood, M.-F.; Khan, S.-U.-D. Hyperbranched copolymer based photoluminescent vesicular probe conjugated with tetraphenylethene: Synthesis, aggregation-induced emission and explosive detection. J. Mol. Liq. 2020, 308, 113034–113042. [Google Scholar] [CrossRef]
  18. Liu, S.-G.; Luo, D.; Li, N.; Zhang, W.; Lei, J.-L.; Li, N.-B.; Luo, H.-Q. Water-Soluble Nonconjugated Polymer Nanoparticles with Strong Fluorescence Emission for Selective and Sensitive Detection of Nitro-Explosive Picric Acid in Aqueous Medium. ACS Appl. Mater. Interfaces 2016, 8, 21700–21709. [Google Scholar] [CrossRef]
  19. Hu, Z.; Deibert, B.-J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815–5840. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, Y.; Ma, Y.; Li, H.; Wang, L. Composite QDS@MIP nanospheres for specific recognition and direct fluorescent quantification of pesticides in aqueous media. Anal. Chem. 2012, 84, 386–395. [Google Scholar] [CrossRef]
  21. Liu, B.; Tong, C.; Feng, L.; Wang, C.; He, Y.; Lü, C. Water-soluble polymer functionalized CdTe/ZnS quantum dots: A facile ratiometric fluorescent probe for sensitive and selective detection of nitroaromatic explosives. Chem. Eur. J. 2014, 20, 2132–2137. [Google Scholar] [CrossRef] [PubMed]
  22. Luo, J.; Xie, Z.; Lam, J.W.-Y.; Cheng, L.; Chen, H.; Qiu, C.; Kwok, H.-S.; Zhan, X.; Liu, Y.; Zhu, D.; et al. Aggregation-induced emission of 1-methyl-1,2,3,4,5-pentaphenylsilole. Chem. Commun. 2001, 18, 1740–1741. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, E.-G.; Chen, Y.-L.; Chen, S.-J.; Deng, H.-Q.; Gui, C.; Leung, C.-W.-T.; Hong, Y.-N.; Lam, J.-W.-Y.; Tang, B.-Z. A luminogen with aggregation-induced emission characteristics for wash-free bacterial imaging, high-throughput antibiotics screening and bacterial susceptibility evaluation. Adv. Mater. 2015, 27, 4931–4937. [Google Scholar] [CrossRef] [PubMed]
  24. Niu, G.-L.; Zheng, X.-L.; Zhao, Z.; Zhang, H.-K.; Wang, J.-G.; He, X.-W.; Chen, Y.-C.; Shi, X.-J.; Ma, C.; Kwok, R.-T.-K.; et al. Functionalized acrylonitriles with aggregation-induced emission: Structure tuning by simple reaction-condition variation, efficient red emission, and two-photon bioimaging. J. Am. Chem. Soc. 2019, 141, 15111–15120. [Google Scholar] [CrossRef] [PubMed]
  25. Gao, M.; Tang, B.-Z. Fluorescent sensors based on aggregation-induced emission: Recent advances and perspectives. ACS Sens. 2017, 2, 1382–1399. [Google Scholar] [CrossRef]
  26. Zhang, J.; Liu, Q.-M.; Wu, W.-J.; Peng, J.-H.; Zhang, H.-K.; Song, F.-Y.; He, B.-Z.; Wang, X.-Y.; Sung, H.H.-Y.; Chen, M.; et al. Real-time monitoring of hierarchical self-assembly and induction of circularly polarized luminescence from achiral luminogens. ACS Nano 2019, 13, 3618–3628. [Google Scholar] [CrossRef]
  27. Kwok, R.T.K.; Leung, C.W.T.; Lam, J.W.Y.; Tang, B.Z. Biosensing by luminogens with aggregation- induced emission characteristics. Chem. Soc. Rev. 2015, 44, 4228–4238. [Google Scholar] [CrossRef]
  28. Wang, J.-H.; Feng, H.-T.; Zheng, Y.-S. Synthesis of tetraphenylethylene pillar[6]arenes and the selective fast quenching of their AIE fluorescence by TNT. Chem. Commun. 2014, 50, 11407–11410. [Google Scholar] [CrossRef]
  29. Che, W.; Li, G.; Liu, X.; Shao, K.; Zhu, D.; Su, Z.; Bryce, M.-R. Selective sensing of 2,4,6-trinitrophenol (TNP) in aqueous media with “aggregation-induced emission enhancement” (AIEE)-active iridium(iii) complexes. Chem. Commun. 2018, 54, 1730–1733. [Google Scholar] [CrossRef] [Green Version]
  30. Zhou, H.; Chua, M.-H.; Tang, B.-Z.; Xu, J. Aggregation-induced emission (AIE)-active polymers for explosive detection. Polym. Chem. 2019, 10, 3822–3840. [Google Scholar] [CrossRef]
  31. Delente, J.M.; Umadevi, D.; Shanmugaraju, S.; Kotova, O.; Watson, G.W.; Gunnlaugsson, T. Aggregation induced emission (AIE) active 4-amino-1,8-naphthalimide-Tröger’s base for the selective sensing of chemical explosives in competitive aqueous media. Chem. Commun. 2020, 56, 2562–2565. [Google Scholar] [CrossRef] [PubMed]
  32. Prusti, B.; Chakravarty, M. An electron-rich small AIEgen as a solid platform for the selective and ultrasensitive on-site visual detection of TNT in the solid, solution and vapor states. Analyst 2020, 145, 1687–1694. [Google Scholar] [CrossRef] [PubMed]
  33. Yuan, W.-Z.; Zhao, H.; Shen, X.-Y.; Mahtab, F.; Lam, J.-W.-Y.; Sun, J.-Z.; Tang, B.-Z. Luminogenic Polyacetylenes and Conjugated Polyelectrolytes: Synthesis, Hybridization with Carbon Nanotubes, Aggregation-Induced Emission, Superamplification in Emission Quenching by Explosives, and Fluorescent Assay for Protein Quantitation. Macromolecules 2009, 42, 9400–9411. [Google Scholar] [CrossRef]
  34. Hu, R.; Maldonado, J.-L.; Rodriguez, M.; Deng, C.; Jim, C.-K.-W.; Lam, J.-W.-Y.; Yuen, M.-M.-F.; Ramos-Ortiz, G.; Tang, B.-Z. Luminogenic materials constructed from tetraphenylethene building blocks: Synthesis, aggregation-induced emission, two-photon absorption, light refraction, and explosive detection. J. Mater. Chem. 2012, 22, 232–240. [Google Scholar] [CrossRef] [Green Version]
  35. Qin, A.; Lam, J.-W.-Y.; Tang, L.; Jim, C.-K.-W.; Zhao, H.; Sun, J.; Tang, B.-Z. Polytriazoles with Aggregation-Induced Emission Characteristics: Synthesis by Click Polymerization and Application as Explosive Chemosensors. Macromolecules 2009, 42, 1421–1424. [Google Scholar] [CrossRef]
  36. Liu, J.; Zhong, Y.; Lam, J.-W.-Y.; Lu, P.; Hong, Y.; Yu, Y.; Yue, Y.; Faisal, M.; Sung, H.-H.-Y.; Williams, I.-D.; et al. Hyperbranched Conjugated Polysiloles: Synthesis, Structure, Aggregation-Enhanced Emission, Multicolor Fluorescent Photopatterning, and Superamplified Detection of Explosives. Macromolecules 2010, 43, 4921–4936. [Google Scholar] [CrossRef]
  37. Feng, H.-T.; Zheng, Y.-S. Highly Sensitive and Selective Detection of Nitrophenolic Explosives by Using Nanospheres of a Tetraphenylethylene Macrocycle Displaying Aggregation-Induced Emission. Chem. Eur. J. 2014, 20, 195–201. [Google Scholar] [CrossRef] [PubMed]
  38. Feng, H.-T.; Wang, J.-H.; Zheng, Y.-S. CH3−π Interaction of Explosives with Cavity of a TPE Macrocycle: The Key Cause for Highly Selective Detection of TNT. ACS Appl. Mater. Interfaces 2014, 6, 20067–20074. [Google Scholar] [CrossRef]
  39. Xiong, J.-B.; Wang, J.-H.; Li, B.; Zhang, C.; Tan, B.; Zheng, Y.-S. Porous Interdigitation Molecular Cage from Tetraphenylethylene Trimeric Macrocycles That Showed Highly Selective Adsorption of CO2 and TNT Vapor in Air. Org. Lett. 2018, 20, 321–324. [Google Scholar] [CrossRef]
  40. Li, D.; Zhang, T.; Ji, B. Influences of pH, urea and metal ions on the interaction of sinomenine with Lysozyme by steady state fluorescence spectroscopy. Spectrochim. Acta Part A 2014, 130, 440–446. [Google Scholar] [CrossRef]
  41. Li, D.; Yang, Y.; Cao, X.; Xu, C.; Ji, B. Investigation on the pH-dependent binding of vitamin B12 and lysozyme by fluorescence and absorbance. J. Mol. Struct. 2012, 1007, 102–112. [Google Scholar] [CrossRef]
Scheme 1. Synthetic route of compound 3.
Scheme 1. Synthetic route of compound 3.
Molecules 28 00181 sch001
Figure 1. (A) Fluorescence spectra of compound 3 (5.0 × 10−5 M) in THF with different water fractions. λex = 340 nm, ex/em slit widths = 5/5 nm. (B) PL intensity ratio of I/I0 in different THF/H2O mixtures. (C) Fluorescent images of compound 3 (5.0 × 10−5 M) with changing water fractions from 0 to 92% in THF under a 365 nm UV lamp.
Figure 1. (A) Fluorescence spectra of compound 3 (5.0 × 10−5 M) in THF with different water fractions. λex = 340 nm, ex/em slit widths = 5/5 nm. (B) PL intensity ratio of I/I0 in different THF/H2O mixtures. (C) Fluorescent images of compound 3 (5.0 × 10−5 M) with changing water fractions from 0 to 92% in THF under a 365 nm UV lamp.
Molecules 28 00181 g001
Figure 2. (A) Fluorescent images of compound 3 with various nitrophenolic explosives under a 365 nm UV lamp. (B) The Fluorescence quenching efficiencies ((1 − I/I0) × 100%), where I and I0 denote the fluorescence intensity of compound 3 with and without analytes, respectively. 1: TNP, 2: DNP, 3: PNP, 4: ONP, 5: TNT, 6: DNT, 7: PNT, 8: ONT, 9: NB, 10: DNB, 11: DNFB, 12: HPN, 13: DNCB, 14: 3−Cyanopheno, 15: Phenol, 16: DNA, 17: HBN. Solvent: H2O: THF = 9: 1, [3] = [explosives]/2 = 2 × 10−5 M. λex = 340 nm, ex/em slit widths = 5/5 nm.
Figure 2. (A) Fluorescent images of compound 3 with various nitrophenolic explosives under a 365 nm UV lamp. (B) The Fluorescence quenching efficiencies ((1 − I/I0) × 100%), where I and I0 denote the fluorescence intensity of compound 3 with and without analytes, respectively. 1: TNP, 2: DNP, 3: PNP, 4: ONP, 5: TNT, 6: DNT, 7: PNT, 8: ONT, 9: NB, 10: DNB, 11: DNFB, 12: HPN, 13: DNCB, 14: 3−Cyanopheno, 15: Phenol, 16: DNA, 17: HBN. Solvent: H2O: THF = 9: 1, [3] = [explosives]/2 = 2 × 10−5 M. λex = 340 nm, ex/em slit widths = 5/5 nm.
Molecules 28 00181 g002
Figure 3. (A) The PL spectra of compound 3 with different amounts of TNP. Solvent: H2O: THF = 9: 1, [3] = 2 × 10−5 M, [TNP] = 0–1.6 × 10−4 M. Inset: the PL change was observed at 460 nm when compound 3 was combined with TNP. (B) The Job’s plot of compound 3 (2 × 10−5 M) with TNP at 460 nm. (C) The PL spectra of compound 3 with different amounts of DNP. Solvent: H2O: THF = 9: 1, [3] = 2 × 10−5 M, [DNP] = 0–3 ×1 0−4 M. Inset: compound 3 fluorescence intensity variation with TNP content at 460 nm. (D) The Job’s plot of compound 3 (2 × 10−5 M) with DNP at 460 nm. λex = 340 nm, ex/em slit widths = 5/5 nm.
Figure 3. (A) The PL spectra of compound 3 with different amounts of TNP. Solvent: H2O: THF = 9: 1, [3] = 2 × 10−5 M, [TNP] = 0–1.6 × 10−4 M. Inset: the PL change was observed at 460 nm when compound 3 was combined with TNP. (B) The Job’s plot of compound 3 (2 × 10−5 M) with TNP at 460 nm. (C) The PL spectra of compound 3 with different amounts of DNP. Solvent: H2O: THF = 9: 1, [3] = 2 × 10−5 M, [DNP] = 0–3 ×1 0−4 M. Inset: compound 3 fluorescence intensity variation with TNP content at 460 nm. (D) The Job’s plot of compound 3 (2 × 10−5 M) with DNP at 460 nm. λex = 340 nm, ex/em slit widths = 5/5 nm.
Molecules 28 00181 g003
Figure 4. (A) Changes of 1H NMR spectra of compound 3 (5 mM) in d6-DMSO with the addition of TNP. (B) The possible binding mode of compound 3 with TNP.
Figure 4. (A) Changes of 1H NMR spectra of compound 3 (5 mM) in d6-DMSO with the addition of TNP. (B) The possible binding mode of compound 3 with TNP.
Molecules 28 00181 g004
Figure 5. Fluorescent photographs of compound 3 with and without TNP solution on the TLC plate under 365 nm UV light.
Figure 5. Fluorescent photographs of compound 3 with and without TNP solution on the TLC plate under 365 nm UV light.
Molecules 28 00181 g005
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, D.; Lv, P.; Han, X.-W.; Jia, Z.; Zheng, M.; Feng, H.-T. A Highly Efficient Fluorescent Sensor Based on AIEgen for Detection of Nitrophenolic Explosives. Molecules 2023, 28, 181. https://doi.org/10.3390/molecules28010181

AMA Style

Li D, Lv P, Han X-W, Jia Z, Zheng M, Feng H-T. A Highly Efficient Fluorescent Sensor Based on AIEgen for Detection of Nitrophenolic Explosives. Molecules. 2023; 28(1):181. https://doi.org/10.3390/molecules28010181

Chicago/Turabian Style

Li, Dongmi, Panpan Lv, Xiao-Wen Han, Zhilei Jia, Min Zheng, and Hai-Tao Feng. 2023. "A Highly Efficient Fluorescent Sensor Based on AIEgen for Detection of Nitrophenolic Explosives" Molecules 28, no. 1: 181. https://doi.org/10.3390/molecules28010181

Article Metrics

Back to TopTop