Next Article in Journal
Transcriptomics and Proteomics Characterizing the Anticancer Mechanisms of Natural Rebeccamycin Analog Loonamycin in Breast Cancer Cells
Next Article in Special Issue
Multifaceted Structurally Coloured Materials: Diffraction and Total Internal Reflection (TIR) from Nanoscale Surface Wrinkling
Previous Article in Journal
rGO-WO3 Heterostructure: Synthesis, Characterization and Utilization as an Efficient Adsorbent for the Removal of Fluoroquinolone Antibiotic Levofloxacin in an Aqueous Phase
Previous Article in Special Issue
Transparent Polymer Opal Thin Films with Intense UV Structural Color
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Random Copolymers of Styrene with Pendant Fluorophore Moieties: Synthesis and Applications as Fluorescence Sensors for Nitroaromatics

by
Mohamad Zen Eddin
1,2,3,*,
Ekaterina F. Zhilina
1,
Roman D. Chuvashov
4,
Alyona I. Dubovik
2,
Alexandr V. Mekhaev
1,
Konstantin A. Chistyakov
1,
Anna A. Baranova
4,
Konstantin O. Khokhlov
4,
Gennady L. Rusinov
1,2,
Egor V. Verbitskiy
1,2,* and
Valery N. Charushin
1,2
1
I. Postovsky Institute of Organic Synthesis, Ural Branch of the Russian Academy of Sciences, S. Kovalevskaya Str. 22, Ekaterinburg 620108, Russia
2
Chemical Engineering Institute, Ural Federal University, Mira Str. 19, Ekaterinburg 620002, Russia
3
College of Science, University of Aleppo, Mouhafaza Str., Aleppo 12212, Syria
4
Institute of Physics and Technology, Ural Federal University, Mira Str. 19, Ekaterinburg 620002, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(20), 6957; https://doi.org/10.3390/molecules27206957
Submission received: 28 September 2022 / Revised: 8 October 2022 / Accepted: 13 October 2022 / Published: 17 October 2022
(This article belongs to the Special Issue Polymeric Photonic Materials)

Abstract

:
Five random copolymers comprising styrene and styrene with pendant fluorophore moieties, namely pyrene, naphthalene, phenanthrene, and triphenylamine, in molar ratios of 10:1, were synthesized and employed as fluorescent sensors. Their photophysical properties were investigated using absorption and emission spectral analyses in dichloromethane solution and in solid state. All copolymers possessed relative quantum yields up to 0.3 in solution and absolute quantum yields up to 0.93 in solid state, depending on their fluorophore components. Fluorescence studies showed that the emission of these copolymers is highly sensitive towards various nitroaromatic compounds, both in solution and in the vapor phase. The detection limits of these fluorophores for nitroaromatic compounds in dichloromethane solution proved to be in the range of 10−6 to 10−7 mol/L. The sensor materials for new hand-made sniffers based on these fluorophores were prepared by electrospinning and applied for the reliable detection of nitrobenzene vapors at 1 ppm in less than 5 min.

1. Introduction

The development of new chemosensors and applications thereof for the trace probing of various explosives is an important task for researchers, considering their potential for improving anti-terrorism measures and homeland security [1,2,3]. In addition, these technologies could help to enhance the environmental and forensic investigations associated with explosive substances [4,5]. Accordingly, the design and synthesis of prompt and highly sensitive devices for the detection of explosive compounds is crucial for safeguarding global security. Nitroaromatic compounds (NACs) are the principal ingredients in many explosive blends, and they are also widely exploited by the agrochemical, dye, and pharmaceutical industries. In particular, 2,4,6-trinitrotoluene (TNT); 2,4-dinitrotoluene (2,4-DNT); and 2,4,6-trinitrophenol (picric acid, PA) are widely used in military ammunition, being present in unexploded landmines around the world [6,7,8].
Although many precise techniques exist for NAC recognition and quantification (e.g., surface-enhanced Raman spectroscopy, ion mobility spectrometry, gas/liquid chromatography/mass spectrometry), they are generally quite expensive, time-consuming, and/or lack the portability necessary for real-time application [9,10].
On the other hand, various types of sensors [11], such as biosensors, electrochemical sensors, colorimetric sensors [12], and fluorescent sensors, have been developed and applied for detection purposes. The fluorescence sensor has received the most attention due to its low cost, high response speed, contactless detection capability, and high detection sensitivity [6,7,8,13].
Thanks to their unique optical and electrochemical properties, various polymers have attracted considerable attention from researchers over the last two decades, which has resulted in numerous technological innovations [14,15]. In particular, their implementation in devices for detecting nitroaromatic explosives should be mentioned [16,17,18,19].
Polycyclic aromatic hydrocarbons, for instance, pyrene, can be considered preferred fluorophore units due to their chemical stability and high quantum yields [20,21,22].
Notably, pyrene was efficiently utilized as a fluorescent indicator dopant in polystyrene to detect vapors of nitro explosives at the nanomolecular level [23]. Recently, Akkoc and Karagoz developed pyrene-based polymeric microbeads as fluorescence chemosensors for the highly sensitive and selective detection of 2,4-dinitrotoluene (DNT) up to 1.39 ppb in aqueous media [24].
This paper deals with a further extension of the research focused on the design of novel fluorophores for chemosensors. Herein, we report the synthesis of new random copolymers of styrene with pendant fluorophore moieties, such as pyrene, naphthalene, phenanthrene, and triphenylamine (Figure 1), and the systematic investigation of their photophysical properties and application as fluorescent sensors for nitroaromatic explosives.

2. Results and Discussion

2.1. Synthesis and Thermal Properties

The monomers with pendant fluorophore moieties, 4-arylstyrenes (IV), were synthesized by reacting 4-bromostyrene with the corresponding arylboronic acid through the microwave-assisted Suzuki coupling reaction (Scheme 1) [25].
The polymerization of vinyl monomers is usually carried out in a radical manner by the action of a thermal initiator. As is well-known, 2,2′-azobis(isobutyronitrile) (AIBN) is the key initiator of thermal radicals, and it can be introduced into the reaction in rather small quantities [26]. The polymerization of styrene with 4-(9-phenanthrenyl)styrene (III) by the action of AIBN, as the initiator, was recently described for the manufacture of organic field-effect transistor memory materials [27]. We chose this reaction as an appropriate method to prepare the sensitive fluorophores (Scheme 2). The conditions of radical polymerization in THF were optimized to obtain the highest yield of the copolymer and to reach the lowest polydispersity (PDI).
The impact of changing the amount of AIBN and THF, the temperature, and the reaction time was studied (Table 1). When the polymerization took place in toluene, the yield of compound P3 was two-times lower than that in THF (Table 1, Entry 2). Unfortunately, polymerization under microwave irradiation did not improve the yield of the desired product P3 (Table 1, Entry 9). The analysis of the results revealed that 1.5 mol.% initiator in 0.5 mL THF at 80 °C for 3 h proved to be the optimal conditions, providing the best yield of polymer P3 and the lowest polydispersity PDI (see Table 1, Entry 7).
The optimum reaction conditions determined above were applied to carry out the copolymerization of styrene with 4-(2-naphthyl)styrene (I), 4-(1-naphthyl)styrene (II), 4-(1-pyrenyl)styrene (III), and 4-(4′-(N,N-diphenylamino)phenyl)styrene (V) to obtain the corresponding copolymers (P1P5) in high yields (Scheme 3, Figures S1–S5). The polymerization results and thermal properties are summarized in Table 2.
The thermal stability of the synthesized copolymers was investigated by thermogravimetric analysis (TGA) under argon flow. The Td for all synthesized polymers proved to be higher than 385 °C (Figure 2, Table 2). These data show the good thermal stability of the obtained polymers for applications as sensor materials. According to the shape of the thermogravimetric curves, it is most likely that the thermal decomposition of all the copolymers P1P5 proceeded along the path of polymer depolymerization, similarly to styrene.
The structures of the obtained vinyl monomers (IV) and copolymers (P1P5) were also proved by Fourier transform infrared spectroscopy (FTIR) (Table S1, Figures S6–S16). All IR spectra of IV contained the bands that are characteristic for vibrations of vinyl groups at ~1626 cm−1. In addition, the out-of-plane =CH2 wagging vibrations gave rise to a strong band at ~900 cm−1. These bands were logically absent in the FTIR spectra of P1P5, in which vibrational modes of methylene groups were observed. The FTIR spectra for copolymers P1P5 were similar to the spectrum of polystyrene [28]. The main difference concerned the presence of CH wagging vibrations among adjacent hydrogens of the aryl substituents in the range of 890–760 cm−1.

2.2. Photophysical Studies of the Obtained Fluorescent Polymers

The photophysical properties of the prepared polymers P1P5 were investigated at room temperature using UV–vis and photoluminescence (PL) spectroscopy in dichloromethane (DCM) solutions and in solid state (Table 3, Figure 3 and Figure 4, and Figures S17–S27).
The UV–vis spectra of polymers P1P5 contained maximum absorption bands at 350–270 nm (ε = 132,600–858,800 M−1·cm−1), corresponding to the absorption to 4-arylphenyl substituents. An intensive band at 260–240 nm (332,700–1,318,300 M−1·cm−1) could be attributed to the absorption of the phenyl ring in polystyrene [29]. The long-wavelength absorption maximum was a bathochromic shift in the series P1P2 < P3 < P5, P4 due to the increasing conjunction of the aromatic system.
Figure 4 and Figures S17–S21 show the steady-state fluorescence spectra of P1P5 in DCM solution. All polymers had a fluorescence band in the range of 350–450 nm, which was most likely associated with emission from the monomeric 4-arylphenyl moiety in the polymer chain [25]. The PL spectra of P1, P3, P3*, and P4 (though not P2 and P5) had a vibrational structure in this fluorescence band, probably due to steric factors. Figure 4 demonstrates that the PL spectrum of P4, in addition to the monomeric structured bands (III, V), showed a longer-wavelength structureless fluorescence band centered at 478 nm, which was attributed to emission from the excimer of pyrene molecules [30].
At the transition from solution to solid powder, the emissions changed noticeably for the P4 and P5 polymers. The emission spectrum of each polymer consisted of one band at 457 and 420 nm, respectively.
Tables S2 and S3, as well as Figures S28–S48, demonstrate the fluorescence decay time for all polymers in DCM and solid powders. The experimental results show that the decay process of P2 and P5 in DCM consisted predominantly of one fast component, 1.80 ns (92.2%) and 1.09 ns (98.0%), respectively, which could be assigned to 4-arylphenyl substituents. This was also observed in the solid state: τ1 = 2.04 ns (97.5%) for P2 and τ1 = 1.08 ns (77.3%) for P5. In addition, a small percentage of the short-lived component was observed for P4: 1.7–2.2 ns (~8%). The well-known excimer conformation of pyrene in P4 at 478 nm showed 100% content with τ = 25.23 ns. An increasing average lifetime was observed for the solid powders of P1, P3, P3*, and P4.

2.3. Detection of Nitroaromatic Compounds in Dichloromethane Solution

Various nitroaromatic compounds, such as nitrobenzene, 2,4-dinitrotoluene, 2,4,6-trinitrotoluene, and picric acid, are not only components of explosives and industrial reagents, but also hazardous ecotoxicants that can cause serious harm to human health (Figure 5). Therefore, we must pay great attention to detecting these dangerous chemicals both in solution and in the vapor phase [31,32,33]. To estimate the ability of the copolymers P1P5 to detect nitroaromatic analytes, the fluorescence titration of P1P5 was performed in dichloromethane solution using a method similar to that previously described in [34]. As depicted in Figure 6, all fluorophores P1P5 acted as efficient detectors of the considered nitroaromatics in solution.
The fluorescence quenching efficiency was quantified, as is typical, by the Stern−Volmer constant (KSV); the fluorophores showed good linear quenching responses in the low concentration range of NB, DNT, TNT, and PA, as indicated by the regression factors (R2) being close to 0.99 in most cases (Figures S49–S54). The Stern–Volmer equation is defined as I0/I = 1 + Ksv × [Q], where I0 and I are the fluorescence intensities in the absence and presence of a quencher and [Q] represents the quencher concentration. The detection limits (DLs) were determined according to the well-known equation DL = 3σ/k, where σ is the standard deviation of the fluorophore intensity in the absence of an analyte and k is the slope of the calibration curve [8]. The data on the obtained Stern–Volmer constants and the detection limits for P1P5 copolymers are summarized in the Table 4.
The linear relationship of the Stern–Volmer plots (in the range of concentrations from 0 to 1 × 10−4 M) (Figures S49–S54) and the lifetime measurements for P5 upon the addition of DNT suggested the predominant role of static interactions in the fluorescence quenching (Figure 7 and Figures S55–S58 and Table S4 in Supplementary Materials). To confirm this suggestion, the quenching rate constant (kq) value was calculated using the Stern–Volmer equation:
I0/I = 1 + Ksv × [Q] = 1 + kqτ0[Q].
Accordingly, the quenching rate constant value could was determined as kq = Ksv/τ0, where τ0 is the average lifetime of the copolymer in the absence of an analyte. For the DNT titrated dichloromethane solution of P5, the Ksv value was found to be 4.66 × 104 M−1, and the average lifetime value (τ0) was calculated as 1.12 × 10−9 s. Evidently, the calculated value of kq = 4.16 × 1013 was much higher than the maximum scatter collision quenching constant, which is 2 × 1010 M−1 × s−1 for dynamic quenching [35]. Therefore, the detection mechanism was undoubtedly static quenching. Thus, we believe that the formation of the non-fluorescent complex P5–DNT took place for the pendant fluorophore moiety due to the π–π interaction, as was previously proved for pyrimidine polymers [33] and push–pull systems [8]. The proposed mechanism of complexation is shown in Figure 8.

2.4. Application of Polymers to Detect Nitroaromatic Compounds in Vapor Phase

2.4.1. Preparation and Morphology of Electrospun Sensing Materials

The on-site detection of nitroaromatic vapors is critical for security providers and for the protection of human health. The electrospinning technique was used to produce sensing materials based on the synthesized polymers. As is common, polystyrene (PS) was used for the electrospinning, and the process parameters were selected according to the literature [36,37,38]. Melamine formaldehyde foam cut into 2 mm sheets was selected as the substrate for material deposition due to its permeability, low fluorescence under 365 nm UV excitation, chemical inertness, availability, and ease of manipulation. The electrospinning solution was prepared by dissolving the sensor polymer in THF to obtain approximately 500 µL of 5% w/v solution. The molecular weight of the polymers was not sufficient to form fibers [36,37,38], and the deposition was achieved by the spraying of polystyrene beads. Sensing materials M1, M2, and M3 were prepared from polymers P4, P1, and P5, respectively. All electrospun polymers produced similar structures, featuring beads approximately 10 µm in diameter (8.65 ± 1.99 µm for P4 (M1), 9.48 ± 1.07 µm for P1 (M2), and 10.10 ± 2.72 µm for P5 (M3), see Figure 9 and Figure S59). The shape of the beads resembled structures obtained for 5% solution w/v in the literature [36,37,38]. Electrospun copolymers P3, P3*, and P2 emitted negligible fluorescence intensities in a solid state under 365 nm UV irradiation and were not included for further examination. Out of the remaining materials, M1 was the brightest and M2 was the dimmest, which was in good agreement with the higher absolute quantum yield in the solid state for copolymer P4 and the much lower yield for copolymer P1 (see Table 3). Figure 9 shows the obtained M1 material under magnification.

2.4.2. Sensor Cartridge and Fluorescence Recorder

To adapt the sensor materials for gas-phase sampling, a cartridge was manufactured from ABS plastic via 3D printing (Figure 10). The cartridge featured radially positioned sensor material slots with through-holes and was installed perpendicular to the airflow. The remainder of the cartridge’s surface blocked the passage of air to force the airflow through the slots. Foam sheets carrying electrospun sensor materials were cut into fragments and installed into the slots. The cartridge design enabled the housing of an array of materials for the parallel evaluation of their responses to gas-phase analytes.
For the detection of vapors, the cartridge was put into an original fluorescent recorder (Figure 11). The main design idea of this portable fluorescence recorder was the use of a compact camera for the parallel recording of the fluorescence intensities of sensing materials in the visible band (Figure S60). A UV-light-emitting diode (λem = 365 nm) was applied for the fluorescence excitation of materials on the cartridge (Figure S61). The camera was fitted with a UV-suppressing optical filter to prevent the interference of the excitation light in the fluorescence recording. During the measurement, the video stream was transmitted for processing to the PC. Custom software was used to operate the device and extract fluorescence intensity data from the video stream.

2.4.3. The Sensing Performance of Materials for Various Analytes in Vapor Phase

The evaluation of the materials was performed as follows. Two measurement scenarios imitating possible sampling cases were formulated: (1) the detection of saturated vapors accumulated in a limited volume (saturated vapor scenario), and (2) the detection of diluted vapors emanating from a vapor source (diluted vapor scenario). Accordingly, the exposure to analytes was achieved either by inserting the vessel with saturated vapors into the airflow, imitating sampling inside a limited volume, or by dosing saturated vapors into the airflow to produce a constant analyte vapor concentration. The duration of exposure was set to 50 s and 100 s for the two scenarios, respectively. The UV illumination intensity was at a constant level during the test. Figure 12 shows an example of a saturated vapor case record (scenario 1); the 50 s exposure interval is marked by dashed lines. According to the Stern–Volmer equation, the fluorescence attenuation was described by the fluorescence intensity before and during exposure, defined as I and I0, respectively. During the exposure to vapors, I was the observed intensity; the base level I0 had to be predicted. In the experiment, a linear model trained on pre-exposure intensity data was used for I0 prediction. The primary task of the prediction was to account for the loss of intensity due to the increase in photobleaching over time. Differences between I and I0 at the end of exposure to vapors Iexp and at the end of post-exposure recovery under clean air Irecov were taken as metrics of the sensor response to vapors and are shown in Figure 12.
The fluorescence intensity responses to saturated vapors of interferents and nitroaromatic analytes are presented in Figure 13. The amplitude of the quenching by the nitroaromatics reflected the saturated vapor pressures of the analytes (372 ppm for NB, 411 ppb for DNT, 9.15 ppb for TNT, and 971 ppt for PA at 25 °C) [39,40]. In the case of analyte vapors, the fluorescence response of the polymers was determined not only by the electron affinity of the analytes towards the fluorophores but also by the permeation dynamics of the vapor molecules into the depths of the polymer. For glassy polymers such as PS, the analyte permeated into the polymer as a saturated front, and the higher vapor pressure allowed for a faster front advance into the interior of the polymer [41]. No significant signal was produced upon exposure to picric acid, presumably due to the saturated vapor pressure at the level below 1 ppb. The material M1 was capable of detecting saturated TNT vapors at 5.8 ppb from a 160 mL sample volume, and saturated DNT vapors were detected at 282.4 ppb and 386.1 ppb from a 160 mL sample volume by M1 and M3, respectively. All materials produced a signal of at least 30% fluorescence quenching upon exposure to NB. A summary of the detection times for the analytes is provided in Table S5.
To study the selectivity of the sensors M1M3, the fluorescence responses towards interferent vapors were evaluated (see columns 1–10 in Figure 13). Sensor materials M1M3 demonstrated good selectivity for NB compared to the other analytes and NACs (see column 11 in Figure 13), since their quenching sensitivities were much weaker in these vapors than in NB.
To determine the sensing performance for diluted nitro compound vapors, the materials were exposed to NB and DNT according to the diluted vapor scenario 2 (Figure S46). Picric acid and 2,4,6-TNT were not selected for the test regarding the magnitude of the response to saturated vapors because their saturated vapor levels were too low. At this dilution level, 2,4-DNT was registered by M3, and NB was registered by M1 and M3. These measurements demonstrated the applicability of the obtained sensory materials for the direct vapor-phase sensing of explosives with the use of a pre-concentrator or via closed packages allowing vapor accumulation.
The lack of fluorescence recovery under clean air after exposure to nitroaromatics was common to all sensor materials. PS is known to have a sorption capacity linked to the π–π interactions of its benzene rings with aromatic analytes, which contributes to the retention of analytes in the polymer, especially those that are weakly volatile and have an electron affinity towards PS moieties [42]. The tendency to accumulate analytes can be utilized for the detection of low concentrations of vapor contaminants within long exposure intervals. In order to simulate the NB detection at the maximum permissible level [43], a 1 ppm concentration was created in the airflow via a syringe driver and a 20 mL syringe containing saturated NB vapors. The sensor cartridge carrying materials M1 and M3 was used in the fluorescence recorder to sample the contaminated air; M2 was ruled out due to low brightness.
The accumulation of nitroaromatic analytes implied that the resulting fluorescence attenuation would grow as the exposure interval duration increased. Therefore, the detection times were calculated as the times required to produce the 3σ threshold signal at a certain analyte concentration so that it would be registered by the fluorescence recorder. The signal value was calculated as the difference between I and I0 within the exposure intervals, and the noise was estimated as the difference between I and I0 for the separate measurement under clean air. The brightness of the material dictated the gain used in the video capture, and the 3σ threshold was higher for dimmer materials due to noise amplification.
The 100 s interval before exposure to vapors was used to estimate the expected fluorescence intensity and the noise at the 3σ level (Figure 14). The results showed that both materials could detect NB at 1 ppm in less than 5 min and could be applied for air contamination monitoring.

3. Experimental Materials and Methods

The detailed specifications of the chemical materials and the methods used for their characterization are provided in the electronic Supplementary Materials.

3.1. Materials

Since we have already described the general details of the analytical equipment and methods [8,22,25] used, they are presented in the “General Information” section (see Supplementary Materials).

3.2. Synthesis of Polymers

The 4-arylstyrene (IV) (0.356 mmol), styrene (372 mg, 3.57 mmol), and AIBN (0.9 mg, 1.5 mol.%) were dissolved in 0.5 mL THF. The resulting reaction mixture was stirred under an argon atmosphere and refluxed for 3 h. After cooling to room temperature, the viscous polymer solution was reprecipitated by pouring the solution into a large excess of ethanol three times. The products, poly[styrene-co-(4-arylstyrene)] (P1P5), were dried under vacuum overnight.
Poly[styrene-co-4-(2-naphthyl)styrene] (P1) was synthesized by the reaction of styrene with 4-(2-naphthyl)styrene (I). Yield 420 mg (76%), white powder; 1H NMR (500 MHz, CDCl3, δ): 7.4–7.9 (CH2CHC6H4C10H7), 6.4–7.4 (CH2CHC6H5, CH2CHC6H4C10H7), 1.2–2.1 (CH2CHC6H5, CH2CHC6H4C10H7); Mw (g∙mol−1)/PDI (GPC): 23 × 103/1.7; Td = 387 °C.
Poly[styrene-co-4-(1-naphthyl)styrene] (P2) was synthesized by the reaction of styrene with 4-(1-naphthyl)styrene (II). Yield 450 mg (82%), white powder; 1H NMR (500 MHz, CDCl3, δ): 7.3–7.9 (CH2CHC6H4C10H7), 6.4–7.3 (CH2CHC6H5, CH2CHC6H4C10H7), 1.3–2.1 (CH2CHC6H5, CH2CHC6H4C10H7); Mw (g∙mol−1)/PDI (GPC): 25 × 103/1.6; Td = 387 °C.
Poly[styrene-co-4-(9-phenanthrenyl)styrene] (P3) (and P3*) was synthesized by the reaction of styrene with 4-(9-phenanthrenyl)styrene (III). Yield 395 mg (83%) for P3, 280 mg (58%) for P3*, white powder; 1H NMR (500 MHz, CDCl3, δ): 7.4–7.9, 8.6–8.8 (CH2CHC6H4C14H9), 6.3–7.2 (CH2CHC6H5, CH2CHC6H4C14H9), 1.2–2.1 (CH2CHC6H5, CH2CHC6H4C14H9); Mw (g∙mol−1)/PDI (GPC): 24 × 103/1.6 for P3, 9×103/1.8 for P3*; Td = 392 °C for P3, 385 °C for P3*.
Poly[styrene-co-4-(1-pyrenyl)styrene] (P4) was synthesized by the reaction of styrene with 4-(1-pyrenyl)styrene (IV). Yield 321 mg (78%), white powder; 1H NMR (500 MHz, CDCl3, δ): 7.8–8.2 (CH2CHC6H4C16H9), 6.3–7.2 (CH2CHC6H5, CH2CHC6H4C14H9), 1.2–2.1 (CH2CHC6H5, CH2CHC6H4C14H9); Mw (g∙mol−1)/PDI (GPC): 24 × 103/1.7; Td = 395 °C.
Poly[styrene-co-4-(4′-(N,N-diphenylamino)phenyl)styrene] (P5) was synthesized by the reaction of styrene with 4-(4’-(N,N-diphenylamino)phenyl)styrene (V). Yield 320 mg (80%), white powder; 1H NMR (500 MHz, CDCl3, δ): 7.2–7.5 (CH2CHC6H4C18H14N), 6.3–7.2 (CH2CHC6H5, CH2CHC6H4 C18H14N), 1.2–2.1 (CH2CHC6H5, CH2CHC6H4C18H14N); Mw (g∙mol−1)/PDI (GPC): 26 × 103/1.6; Td = 391 °C.

4. Conclusions

We described the synthesis and photophysical properties of five random copolymers comprising styrene and styrene with pendant fluorophore moieties, namely pyrene, naphthalene, phenanthrene, and triphenylamine, in molar ratios of 10:1. It was shown that the emission of these copolymers was highly sensitive towards various nitroaromatic compounds, both in solution and in the vapor phase. The detection limits of these fluorophores for nitroaromatic compounds in dichloromethane solution reached 10−7 mol/L. The sensor materials for new hand-made sniffers based on these fluorophores were prepared by electrospinning and applied for the reliable detection of nitrobenzene vapors at 1 ppm in less than 5 min.

Supplementary Materials

The following supporting information can be downloaded at: https://mdpi.longhoe.net/article/10.3390/molecules27206957/s1. General Information; Figure S1: 1H NMR (500 MHz, CDCl3) spectrum of P1; Figure S2: 1H NMR (500 MHz, CDCl3) spectrum of P2; Figure S3: 1H NMR (500 MHz, CDCl3) spectrum of P3; Figure S4: 1H NMR (500 MHz, CDCl3) spectrum of P4; Figure S5: 1H NMR (500 MHz, CDCl3) spectrum of P5; Figure S6: IR spectrum of I; Figure S7: IR spectrum of II; Figure S8: IR spectrum of III; Figure S9. IR spectrum of IV; Figure S10: IR spectrum of V; Figure S11: IR spectrum of P1; Figure S12: IR spectrum of P2; Figure S13: IR spectrum of P3; Figure S14: IR spectrum of P3*; Figure S15: IR spectrum of P4; Figure S16: IR spectrum of P5; Table S1: IR characteristic frequencies of I–V and P1–P5; Figure S17: Absorption (blue), fluorescence emission (green) and fluorescence excitation (blue dashed) spectra of compound P1 in CH2Cl2; Figure S18: Absorption (blue), fluorescence emission (green) and fluorescence excitation (blue dashed) spectra of compound P2 in CH2Cl2; Figure S19: Absorption (blue), fluorescence emission (green) and fluorescence excitation (blue dashed) spectra of compound P3* in CH2Cl2; Figure S20: Absorption (blue), fluorescence emission (green) and fluorescence excitation (blue dashed) spectra of compound P3 in CH2Cl2; Figure S21: Absorption (blue), fluorescence emission (green) and fluorescence excitation (blue dashed) spectra of compound P5 in CH2Cl2; Figure S22: Excitation (green) and emission (blue) spectra of solid powder P1; Figure S23: Excitation (green) and emission (blue) spectra of solid powder P2; Figure S24: Excitation (green) and emission (blue) spectra of solid powder P3; Figure S25: Excitation (blue) and emission (green) spectra of solid powder P3*; Figure S26: Excitation (green) and emission (blue) spectra of solid powder P4; Figure S27: Excitation (green) and emission (blue) spectra of solid powder P5; Table S2: Detailed data of the fluorescence lifetime measurements of P1-P5 in DCM: τ – lifetime, f - fractional contribution, τavg – average lifetime, χ2 - chi-squared distribution; Figure S28: Time-resolved fluorescence lifetime decay profile of P1 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 362 nm; Figure S29: Time-resolved fluorescence lifetime decay profile of P1 (green), IRF (blue). λex = 300 nm, λex = 345 nm; Figure S30: Time-resolved fluorescence lifetime decay profile of P2 (green), IRF (blue). λex = 300 nm, λex = 358 nm; Figure S31: Time-resolved fluorescence lifetime decay profile of P3 (green), IRF (blue). λex = 300 nm, λex = 376 nm; Figure S32: Time-resolved fluorescence lifetime decay profile of P3 (green), IRF (blue). λex = 300 nm, λex = 359 nm; Figure S33: Time-resolved fluorescence lifetime decay profile of P3* (green), IRF (blue). λex = 300 nm, λex = 376 nm; Figure S34: Time-resolved fluorescence lifetime decay profile of P3* (green), IRF (blue). λex = 300 nm, λex = 359 nm; Figure S35: Time-resolved fluorescence lifetime decay profile of P4 (green), IRF (blue). λex = 300 nm, λex = 478 nm; Figure S36: Time-resolved fluorescence lifetime decay profile of P4 (green), IRF (blue). λex = 300 nm, λex = 401 nm; Figure S37: Time-resolved fluorescence lifetime decay profile of P4 (green), IRF (blue). λex = 300 nm, λex = 383 nm; Figure S38: Time-resolved fluorescence lifetime decay profile of P5 (green), IRF (blue). λex = 300 nm, λex = 392 nm; Table S3: Detailed data of the fluorescence lifetime measurements of solid powders P1-P5: τ – lifetime, f - fractional contribution, τavg – average lifetime, χ2 - chi-squared distribution; Figure S39: Time-resolved fluorescence lifetime decay profile of solid P1 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 362 nm; Figure S40: Time-resolved fluorescence lifetime decay profile of solid P1 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 355 nm; Figure S41: Time-resolved fluorescence lifetime decay profile of solid P1 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 346 nm; Figure S42: Time-resolved fluorescence lifetime decay profile of solid P2 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 359 nm; Figure S43: Time-resolved fluorescence lifetime decay profile of solid P3 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 377 nm; Figure S44: Time-resolved fluorescence lifetime decay profile of solid P3 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 364 nm; Figure S45: Time-resolved fluorescence lifetime decay profile of solid P3* (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 377 nm; Figure S46: Time-resolved fluorescence lifetime decay profile of solid P3* (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 364 nm; Figure S47: Time-resolved fluorescence lifetime decay profile of solid P4 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 457 nm; Figure S48: Time-resolved fluorescence lifetime decay profile of solid P5 (green), instrumental response function (IRF, blue). λex = 300 nm, λex = 420 nm; Figure S49: Fluorescence quenching studies of P1 (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 290 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 290 nm for P1 solution; Figure S50: Fluorescence quenching studies of P2 (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 294 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 294 nm for P2 solution; Figure S51: Fluorescence quenching studies of P3 (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 301 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 301 nm for P3 solution; Figure S52: Fluorescence quenching studies of P3* (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 301 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 301 nm for P3* solution; Figure S53: Fluorescence quenching studies of P4 (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 345 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 345 nm for P4 solution; Figure S54: Fluorescence quenching studies of P5 (1.0 × 10−5 mol/L) recorded in the presence of various amounts of NB (a), DNT (b), TNT (c), and PA (d), of which 325 nm was taken as the excitation wavelength. The Stern–Volmer plots as function of NB (e), DNT (f), TNT (g), and PA (h) concentration in CH2Cl2, with an excitation wavelength of 325 nm for P5 solution; Table S4: Detailed data of the fluorescence lifetime measurements of P1-P5 in DCM: τ – lifetime, f - fractional contribution, τavg – average lifetime, χ2 - chi-squared distribution; Figure S55: Time-resolved fluorescence lifetime decay profile of P5 with DNT (1×10−4 M) (blue), IRF (green). λex = 300 nm, λex = 392 nm; Figure S56: Time-resolved fluorescence lifetime decay profile of P5 with DNT (2×10-4 M) (blue), IRF (green). λex = 300 nm, λex = 392 nm; Figure S57: Time-resolved fluorescence lifetime decay profile of P5 with DNT (3×10−4 M) (blue), IRF (green). λex = 300 nm, λex = 392 nm; Figure S58: Time-resolved fluorescence lifetime decay profile of P5 with DNT (4×10−4 M) (blue), IRF (green). λex = 300 nm, λex = 392 nm; Figure S59: Photos of sensor materials M1–M3 obtained by electrospinning of fluorescent polymers onto melamine-formaldehyde substrate. For shots under UV illumination, the source of illumination was fixed relative to the material; Figure S60: Photos of the typical sensor cartridge loadout used for measurements under UV illumination (365 nm) shot by the phone camera (left) and the camera of the fluorescence recorder (right); Figure S61: The optical scheme of fluorescence recorder. The device has disassembled for the sensor installation; Figure S62: Fluorescence responses of sensor materials to NB and DNT vapors at a concentration of 10% from saturated one in a 100 sec exposure interval. The 3σ noise levels of registered fluorescence intensity difference are marked by horizontal dashed lines; Table S5: Detection times of obtained sensor materials towards nitroaromatic vapours when applied in the fluorescence recorder. Saturated vapor concentrations at 25 °C; Figure S63: The scheme of vapor measurements setup; Figure S64: Vapor vessels, the 3 L glass vessel on the left, and the 160 mL syringe vessel installed onto the syringe driver on the right. References [44,45,46,47] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, E.V.V.; Data curation, A.A.B. and K.O.K.; Formal analysis, E.F.Z. and R.D.C.; Funding acquisition, G.L.R. and V.N.C.; Investigation, A.V.M., M.Z.E., E.F.Z., A.I.D. and R.D.C.; Methodology, E.F.Z., K.A.C. and R.D.C.; Project administration, V.N.C.; Validation, E.V.V., K.A.C. and E.F.Z.; Visualization, E.F.Z. and R.D.C.; Writing—original draft, R.D.C., M.Z.E., E.V.V.; Writing—review and editing, E.V.V. and V.N.C. All authors have read and agreed to the published version of the manuscript.

Funding

The synthetic part of this work was supported by the Ministry of Science and Higher Education of the Russian Federation within the framework of the State Assignment for Research (Project No. AAAA-A19-119011790132-7). E.V.V. is grateful for the financial support for the photophysical and sensory investigation from the Ministry of Science and Higher Education of the Russian Federation (Agreement with Zelinsky Institute of Organic Chemistry RAS No 075-15-2020-803).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author and co-authors.

Acknowledgments

Analytical studies were carried out using equipment from the Center for Joint Use “Spectroscopy and Analysis of Organic Compounds” at the Postovsky Institute of Organic Synthesis of the Ural Branch of the Russian Academy of Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds P1–P5 are available from the authors.

References

  1. Yinon, J. (Ed.) Counterterrorist Detection Techniques of Explosives; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  2. Giannoukos, S.; Brkić, B.; Taylor, S.; Marshall, T.; Verbeck, G.F. Chemical Sniffing Instrumentation for Security Applications. Chem. Rev. 2016, 116, 8146–8172. [Google Scholar] [CrossRef] [PubMed]
  3. Salinas, Y.; Martínez-Máñez, R.; Marcos, M.D.; Sancenón, 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]
  4. Sunahara, G.I.; Lotufo, G.; Kuperman, R.G.; Hawari, J. Ecotoxicology of Explosives; CRC Press: Boca Raton, FL, USA; Taylor & Francis: London, UK; New York, NY, USA, 2009. [Google Scholar]
  5. Klapec, D.J.; Czarnopys, G.; Pannuto, J. Interpol review of detection and characterization of explosives and explosives residues 2016–2019. Forensic Sci. Int. Synerg. 2020, 2, 670–700. [Google Scholar] [CrossRef]
  6. La Grone, M.J.; Cumming, C.J.; Fisher, M.E.; Fox, M.J.; Jacob, S.; Reust, D.; Rockley, M.G.; Towers, E. Detection of land mines by amplified fluorescence quenching of polymer films: A man-portable chemical sniffer for detection of ultratrace concentrations of explosives emanating from land mines. In Detection and Remediation Technologies for Mines and Minelike Targets V; SPIE: Bellingham, WA, USA, 2000; Volume 4038. [Google Scholar] [CrossRef]
  7. Zyryanov, G.V.; Kopchuk, D.S.; Kovalev, I.S.; Nosova, E.V.; Rusinov, V.L.; Chupakhin, O.N. Chemosensors for detection of nitroaromatic compounds (explosives). Russ. Chem. Rev. 2014, 83, 783–819. [Google Scholar] [CrossRef]
  8. Verbitskiy, E.V.; Baranova, A.A.; Lugovik, K.I.; Shafikov, M.Z.; Khokhlov, K.O.; Cheprakova, E.M.; Rusinov, G.L.; Chupakhin, O.N.; Charushin, V.N. Detection of nitroaromatic explosives by new D–π–A sensing fluorophores on the basis of the pyrimidine scaffold. Anal. Bioanal. Chem. 2016, 408, 4093–4101. [Google Scholar] [CrossRef]
  9. Moore, D.S. Instrumentation for trace detection of high explosives. Rev. Sci. Instum. 2004, 75, 2499–2512. [Google Scholar] [CrossRef]
  10. Moore, D.S. Recent advances in trace explosives detection instrumentation. Sens. Imaging 2007, 8, 9–38. [Google Scholar] [CrossRef]
  11. Singh, S. Sensors—An effective approach for the detection of explosives. J. Hazard Mater. 2007, 144, 15–28. [Google Scholar] [CrossRef]
  12. Wu, D.; Sedgwick, A.C.; Gunnlaugsson, T.; Akkaya, E.U.; Yoon, J.; James, T.D. Fluorescent chemosensors: The past, present and future. Chem. Soc. Rev. 2017, 46, 7105–7123. [Google Scholar] [CrossRef]
  13. Verbitskiy, E.V.; Rusinov, G.L.; Chupakhin, O.N.; Charushin, V.N. Design of fluorescent sensors based on azaheterocyclic push-pull systems towards nitroaromatic explosives and related compounds: A review. Dyes Pigment. 2020, 180, 108414. [Google Scholar] [CrossRef]
  14. Kumar, V.; Maiti, B.; Chini, M.K.; De, P.; Satapathi, S. Multimodal Fluorescent Polymer Sensor for Highly Sensitive Detection of Nitroaromatics. Sci. Rep. 2019, 9, 7269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kumar, V.; Saini, S.K.; Choudhury, N.; Kumar, A.; Maiti, B.; De, P.; Kumar, M.; Satapathi, S. Highly Sensitive Detection of Nitro Compounds Using a Fluorescent Copolymer-Based FRET System. ACS Appl. Polym. Mater. 2021, 3, 4017–4026. [Google Scholar] [CrossRef]
  16. Li, J.; Kendig, C.E.; Nesterov, E.E. Chemosensory Performance of Molecularly Imprinted Fluorescent Conjugated Polymer Materials. J. Am. Chem. Soc. 2007, 129, 15911–15918. [Google Scholar] [CrossRef] [PubMed]
  17. Nie, H.; Sun, G.; Zhang, M.; Baumgarten, M.; Müllen, K. Fluorescent conjugated polycarbazoles for explosives detection: Side chain effects on TNT sensor sensitivity. J. Mater. Chem. 2012, 22, 2129–2132. [Google Scholar] [CrossRef]
  18. Rochat, S.; Swager, T.M. Conjugated Amplifying Polymers for Optical Sensing Applications. ACS Appl. Mater. Interfaces 2013, 5, 4488–4502. [Google Scholar] [CrossRef]
  19. McQuade, D.T.; Pullen, A.P.; Swager, T.M. Conjugated Polymer-Based Chemical Sensors. Chem. Rev. 2000, 100, 2537–2574. [Google Scholar] [CrossRef] [PubMed]
  20. Turhan, H.; Tukenmez, E.; Karagoz, B.; Bicak, N. Highly fluorescent sensing of nitroaromatic explosives in aqueous media using pyrene-linked PBEMA microspheres. Talanta 2018, 179, 107–114. [Google Scholar] [CrossRef] [PubMed]
  21. Qazi, F.; Shahsavari, E.; Prawer, S.; Ball, A.S.; Tomljenovic-Hanic, S. Detection and identification of polyaromatic hydrocarbons (PAHs) contamination in soil using intrinsic fluorescence. Environ. Pollut. 2021, 272, 116010. [Google Scholar] [CrossRef]
  22. Verbitskiy, E.V.; Baranova, A.A.; Lugovik, K.I.; Khokhlov, K.O.; Chuvashov, R.D.; Dinastiya, E.M.; Rusinov, G.L.; Chupakhin, O.N.; Charushin, V.N. Linear and V-shaped push–pull systems on a base of pyrimidine scaffold with a pyrene–donative fragment for detection of nitroaromatic compounds. J. Iran Chem. Soc. 2018, 15, 787–797. [Google Scholar] [CrossRef]
  23. Wang, Y.; La, A.; Ding, Y.; Liu, Y.; Lei, Y. Novel Signal-Amplifying Fluorescent Nanofibers for Naked-Eye-Based Ultrasensitive Detection of Buried Explosives and Explosive Vapors. Adv. Funct. Mater. 2012, 22, 3547–3555. [Google Scholar] [CrossRef]
  24. Akkoc, E.; Karagoz, B. One step synthesis of crosslinked fluorescent microspheres for the effective and selective sensing of explosives in aqueous media. Eur. Polym. J. 2022, 172, 111238. [Google Scholar] [CrossRef]
  25. Eddin, M.Z.; Pervova, M.G.; Zhilina, E.F.; Chistyakov, K.A.; Verbitskiy, E.V.; Rusinov, G.L.; Charushin, V.N. A new approach to 4-arylstyrenes: Microwave-assisted synthesis and photophysical properties. Russ. Chem. Bull. 2021, 70, 2139–2144. [Google Scholar] [CrossRef]
  26. Moad, G. A Critical Assessment of the Kinetics and Mechanism of Initiation of Radical Polymerization with Commercially Available Dialkyldiazene Initiators. Prog. Polym. Sci. 2019, 88, 130–188. [Google Scholar] [CrossRef]
  27. Chung, F.-J.; Liu, H.-Y.; Jiang, B.-Y.; He, G.-Y.; Wang, S.-H.; Wu, W.-C.; Liu, C.-L. Random Styrenic Copolymers with Pendant Pyrene Moieties: Synthesis and Applications in Organic Field-Effect Transistor Memory. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 910–917. [Google Scholar] [CrossRef]
  28. Liang, C.Y.; Krimm, S. Infared spectra of high polymers. VI. Polystyrene. J. Polym. Sci. 1958, 27, 241–254. [Google Scholar] [CrossRef]
  29. Fedorenko, E.V.; Mirochnik, A.G.; Beloliptsev, A.Y. New polymers containing BF2-benzoylacetonate groups. Synthesis, luminescence, excimer and exciplex formation. J. Lumin. 2017, 185, 23–33. [Google Scholar] [CrossRef]
  30. Bains, G.K.; Kim, S.H.; Sorin, E.J.; Narayanaswami, V. The Extent of Pyrene Excimer Fluorescence Emission Is a Reflector of Distance and Flexibility: Analysis of the Segment Linking the LDL Receptor-Binding and Tetramerization Domains of Apolipoprotein E3. Biochemistry 2012, 51, 6207–6219. [Google Scholar] [CrossRef] [Green Version]
  31. US EPA. Provisional Peer-Reviewed Toxicity Values for Picric Acid (2,4,6-Trinitrophenol); CASRN 88-89-1; United States Environmental Protection Agency: Washington, DC, USA, 2015.
  32. ATSDR. 2,4,6-Trinitrotoluene (TNT); Agency for Toxic Substances and Disease Registry (ATSDR), 1996. Available online: https://wwwn.cdc.gov/TSP/substances/ToxSubstance.aspx?toxid=125 (accessed on 28 September 2022).
  33. Svalova, T.S.; Saigushkina, A.A.; Verbitskiy, E.V.; Chistyakov, K.A.; Varaksin, M.V.; Rusinov, G.L.; Charushin, V.N.; Kozitsina, A.N. Rapid and sensitive determination of nitrobenzene in solutions and commercial honey samples using a screen-printed electrode modified by 1,3-/1,4-diazines. Food Chem. 2022, 372, 131279. [Google Scholar] [CrossRef] [PubMed]
  34. Verbitskiy, E.V.; Baranova, A.A.; Lugovik, K.I.; Khokhlov, K.O.; Cheprakova, E.M.; Rusinov, G.L.; Chupakhin, O.N.; Charushin, V.N. New V-shaped push-pull systems based on 4,5-di(hetero)aryl substituted pyrimidines: Their synthesis and application to the detection of nitroaromatic explosives. ARKIVOC 2016, 2016, 360–373. [Google Scholar] [CrossRef]
  35. Ware, W.R. Oxygen quenching of fluorescence in solution: An experimental study of diffusion process. J. Phys. Chem. 1962, 66, 455–458. [Google Scholar] [CrossRef]
  36. Eda, G.; Shivkumar, S. Bead-to-fiber transition in electrospun polystyrene. J. Appl. Polym. Sci. 2007, 106, 475–487. [Google Scholar] [CrossRef]
  37. Zheng, J.; Zhang, H.; Zhao, Z.; Han, C.C. Construction of hierarchical structures by electrospinning or electrospraying. Polymer 2012, 53, 546–554. [Google Scholar] [CrossRef]
  38. Casper, C.L.; Stephens, J.S.; Tassi, N.G.; Chase, D.B.; Rabolt, J.F. Controlling Surface Morphology of Electrospun Polystyrene Fibers:  Effect of Humidity and Molecular Weight in the Electrospinning Process. Macromolecules 2004, 37, 573–578. [Google Scholar] [CrossRef]
  39. Lynch, E.J.; Wilke, C.R. Vapor Pressure of Nitrobenzene at Low Temperatures. J. Chem. Eng. Data 1960, 5, 300. [Google Scholar] [CrossRef] [Green Version]
  40. Ewing, R.G.; Waltman, M.J.; Atkinson, D.A.; Grate, J.W.; Hotchkiss, P.J. The vapor pressures of explosives. Trends Anal. Chem. 2013, 42, 35–48. [Google Scholar] [CrossRef]
  41. Shaw, P.E.; Bun, P.L. Real-time fluorescence quenching-based detection of nitro-containing explosive vapours: What are the key processes? Phys. Chem. Chem. Phys. 2017, 19, 29714–29730. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, J.; Liu, X.; Liu, G.; Zhang, Z.; Wu, H.; Cui, B.; Bai, J.; Zhang, W. Size effect of polystyrene microplastics on sorption of phenanthrene and nitrobenzene. Ecotoxicol. Environ. Saf. 2019, 173, 331–338. [Google Scholar] [CrossRef]
  43. Permissible Exposure Limit for Nitrobenzene in the US. Available online: https://www.cdc.gov/niosh/npg/npgd0450.html (accessed on 28 September 2022).
  44. Rurack, K. Standardization and Quality Assurance in Fluorescence Measurements I; Springer: Berlin/Heidelberg, Germany, 2008; pp. 101–145. [Google Scholar]
  45. Östmark, H.; Wallin, S.; Ang, H.G. Vapor Pressure of Explosives: A Critical Review. Propellants Explos. Pyrotech. 2012, 37, 12–23. [Google Scholar] [CrossRef]
  46. Chuvashov, R.; Baranova, A.; Khokhlov, K.; Verbitskiy, E. A detection system with low sampling distortion for application in optical array sensing in gas phase. In Proceedings of the 2020 7th International Congress on Energy Fluxes and Radiation Effects, EFRE 2020, Tomsk, Russia, 14–26 September 2020; pp. 984–988. [Google Scholar] [CrossRef]
  47. Baranova, A.A.; Khokhlov, K.O.; Chuvashov, R.D.; Verbitskiy, E.V.; Cheprakova, E.M.; Rusinov, G.L.; Charushin, V.N. The portable detector of nitro-explosives in vapor phase with new sensing elements on the base of pyrimidine scaffolds. J. Phys. Conf. Ser. 2017, 830, 012159. [Google Scholar]
Figure 1. Structure of investigated polymers P1P5.
Figure 1. Structure of investigated polymers P1P5.
Molecules 27 06957 g001
Scheme 1. Syntheses of 4-arylstyrenes (I–V) through the microwave-assisted Suzuki coupling reaction.
Scheme 1. Syntheses of 4-arylstyrenes (I–V) through the microwave-assisted Suzuki coupling reaction.
Molecules 27 06957 sch001
Scheme 2. Polymerization of styrene with 4-(9-phenanthrenyl)styrene (III).
Scheme 2. Polymerization of styrene with 4-(9-phenanthrenyl)styrene (III).
Molecules 27 06957 sch002
Scheme 3. Synthesis of poly[styrene-co-(4-arylstyrene)] (P1P5).
Scheme 3. Synthesis of poly[styrene-co-(4-arylstyrene)] (P1P5).
Molecules 27 06957 sch003
Figure 2. Thermogravimetric curves for polymers P1P5.
Figure 2. Thermogravimetric curves for polymers P1P5.
Molecules 27 06957 g002
Figure 3. Absorption spectra of P1P5 in DCM solution.
Figure 3. Absorption spectra of P1P5 in DCM solution.
Molecules 27 06957 g003
Figure 4. Absorption (--), excitation (–), and emission (–) spectra of P4 in DCM.
Figure 4. Absorption (--), excitation (–), and emission (–) spectra of P4 in DCM.
Molecules 27 06957 g004
Figure 5. Structures of utilized nitroaromatic quenchers.
Figure 5. Structures of utilized nitroaromatic quenchers.
Molecules 27 06957 g005
Figure 6. Quenching efficiencies of NB, DNT, TNT, and PA relative to fluorophores P1 (a), P2 (b), P3 (c), P3* (d), P4 (e), and P5 (f) at mM level.
Figure 6. Quenching efficiencies of NB, DNT, TNT, and PA relative to fluorophores P1 (a), P2 (b), P3 (c), P3* (d), P4 (e), and P5 (f) at mM level.
Molecules 27 06957 g006
Figure 7. Fluorescence decay time curves of P5 with DNT in DCM at an excitation wavelength of 300 nm and an emission wavelength of 392 nm.
Figure 7. Fluorescence decay time curves of P5 with DNT in DCM at an excitation wavelength of 300 nm and an emission wavelength of 392 nm.
Molecules 27 06957 g007
Figure 8. The proposed complexation mechanism of copolymer P5 with DNT.
Figure 8. The proposed complexation mechanism of copolymer P5 with DNT.
Molecules 27 06957 g008
Figure 9. Microphoto of M1 material electrospun on foam substrate.
Figure 9. Microphoto of M1 material electrospun on foam substrate.
Molecules 27 06957 g009
Figure 10. Sensor element cartridge under UV (left) and visible (right) illumination.
Figure 10. Sensor element cartridge under UV (left) and visible (right) illumination.
Molecules 27 06957 g010
Figure 11. Fluorescence recorder and cartridge used in the experiment: assembled (a) and disassembled (b).
Figure 11. Fluorescence recorder and cartridge used in the experiment: assembled (a) and disassembled (b).
Molecules 27 06957 g011
Figure 12. Fluorescence responses of materials M1M3 towards NB vapors in saturated vapor scenario. Vertical dashed lines mark the vapor exposure interval.
Figure 12. Fluorescence responses of materials M1M3 towards NB vapors in saturated vapor scenario. Vertical dashed lines mark the vapor exposure interval.
Molecules 27 06957 g012
Figure 13. Fluorescence responses of sensor materials M1M3 towards saturated vapors of interferents and nitroaromatic compounds in a 50 s exposure interval. The 3σ noise levels of registered fluorescence intensity differences are marked with horizontal dashed lines.
Figure 13. Fluorescence responses of sensor materials M1M3 towards saturated vapors of interferents and nitroaromatic compounds in a 50 s exposure interval. The 3σ noise levels of registered fluorescence intensity differences are marked with horizontal dashed lines.
Molecules 27 06957 g013
Figure 14. Fluorescence response of materials M1 and M3 to NB vapors at 1 ppm concentration.
Figure 14. Fluorescence response of materials M1 and M3 to NB vapors at 1 ppm concentration.
Molecules 27 06957 g014
Table 1. Optimization of the copolymerization reaction between styrene and 9-(4-vinylphenyl)phenanthrene (III).
Table 1. Optimization of the copolymerization reaction between styrene and 9-(4-vinylphenyl)phenanthrene (III).
EntryAIBN (mol.%)Yield (%)MwMnPDI
10.552780042001.8
2(d)0.525320023001.4
31.06012,40076001.6
41.57020,00011,0001.8
52.058900050001.8
6(a)1.58016,00098001.6
7(b)1.58424,00014,4001.6
8(c)1.57031,50019,1001.7
9(e)1.53111,50041002.8
Number average molecular weight (Mn), weight average molecular weight (Mw), and polydispersity (PDI) were determined by GPC by eluting with THF. General conditions: monomer III (100 mg, 0.356 mmol) and styrene (372 mg, 3.566 mmol) in 5 mL of THF at 80 °C/12 h; (a) in 2 m L of THF at 80 °C/12 h; (b) in 0.5 mL of THF at 80 °C/3 h; (c) in 0.3 mL of THF at 80 °C/2 h. (d) In toluene. (e) Reactions under microwave irradiation in 5 mL THF at 100 °C/3 h.
Table 2. Polymerization results and thermal properties of P1P5.
Table 2. Polymerization results and thermal properties of P1P5.
CopolymerYield, %MwMnPDITd, °C
P17623,00013,5001.7387
P28225,00015,5001.6387
P38324,00014,4001.6392
P3*58900050001.8385
P47324,00014,0001.7395
P58026,00016,0001.6391
General conditions: 4-arylstyrenes IV (100 mg,1 equiv.), styrene (10 equiv.), and AIBN (1.5 mol.%) in 0.5 mL THF at 80 °C/3 h.
Table 3. Optical properties of polymers P1P5.
Table 3. Optical properties of polymers P1P5.
PolymerUV AbsorptionFluorescence
Solution in DCMSolid
λabsmax (nm)εmax (M−1 cm−1)Excitation λex (nm)Emission λem (nm)τ avg (ns)ΦF aExcitation λex (nm)Emission λem (nm0τ avg (ns)ΦF c
P1292310,00029036221.190.1730536241.250.44
2591,031,00026334521.6035540.31
34640.02
P2294309,0002943582.060.283113592.150.48
2301,271,700230
P3301336,80030137614.570.0835137727.880.24
2581,318,30026535914.7431336427.95
P3*301132,60030137614.410.0735137727.530.19
258517,30025835914.3631336427.21
P4345693,00034547825.230.29 b37545723.500.93
281858,80028140111.46
271585,10027138311.35350
2451,025,300245
P5322552,3003253921.120.303504201.320.77
246332,700246
a ΦF values were determined relative to 2-aminopyridine in 0.1N H2SO4 as standard (ΦF = 0.60); excitation at 300 nm; b excitation at 345 nm. c ΦF values were determined using integrating sphere SC-30 of FS5 Edinburgh Instruments spectrofluorometer.
Table 4. Stern–Volmer constants and detection limits of nitroaromatics for fluorophores P1-P5 in CH2Cl2.
Table 4. Stern–Volmer constants and detection limits of nitroaromatics for fluorophores P1-P5 in CH2Cl2.
PolymerKsv × 104, M−1/DL, mol × L−1
PATNTDNTNB
P1405.79/1.74 × 10−724.32/4.02 × 10−734.33/3.70 × 10−752.26/1.92 × 10−7
P289.16/5.11 × 10−719.91/3.63 × 10−741.78/3.44 × 10−729.56/3.71 × 10−7
P340.25/3.55 × 10−714.70/4.51 × 10−724.81/2.58 × 10−717.14/2.91 × 10−7
P3*42.36/1.18 × 10−618.15/2.19 × 10−725.88/2.11 × 10−717.51/2.39 × 10−7
P454.21/4.78 × 10−77.56/6.63 × 10−75.20/7.18 × 10−73.08/1.04 × 10−6
P576.24/1.74 × 10−75.74/1.54 × 10−74.66/3.71 × 10−72.70/5.55 × 10−7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zen Eddin, M.; Zhilina, E.F.; Chuvashov, R.D.; Dubovik, A.I.; Mekhaev, A.V.; Chistyakov, K.A.; Baranova, A.A.; Khokhlov, K.O.; Rusinov, G.L.; Verbitskiy, E.V.; et al. Random Copolymers of Styrene with Pendant Fluorophore Moieties: Synthesis and Applications as Fluorescence Sensors for Nitroaromatics. Molecules 2022, 27, 6957. https://doi.org/10.3390/molecules27206957

AMA Style

Zen Eddin M, Zhilina EF, Chuvashov RD, Dubovik AI, Mekhaev AV, Chistyakov KA, Baranova AA, Khokhlov KO, Rusinov GL, Verbitskiy EV, et al. Random Copolymers of Styrene with Pendant Fluorophore Moieties: Synthesis and Applications as Fluorescence Sensors for Nitroaromatics. Molecules. 2022; 27(20):6957. https://doi.org/10.3390/molecules27206957

Chicago/Turabian Style

Zen Eddin, Mohamad, Ekaterina F. Zhilina, Roman D. Chuvashov, Alyona I. Dubovik, Alexandr V. Mekhaev, Konstantin A. Chistyakov, Anna A. Baranova, Konstantin O. Khokhlov, Gennady L. Rusinov, Egor V. Verbitskiy, and et al. 2022. "Random Copolymers of Styrene with Pendant Fluorophore Moieties: Synthesis and Applications as Fluorescence Sensors for Nitroaromatics" Molecules 27, no. 20: 6957. https://doi.org/10.3390/molecules27206957

Article Metrics

Back to TopTop