Next Article in Journal
A Non-Contact Measurement System for the Range of Motion of the Hand
Next Article in Special Issue
Maximum Measurement Range and Accuracy of SAW Reflective Delay Line Sensors
Previous Article in Journal
Distributed Synchronization Technique for OFDMA-Based Wireless Mesh Networks Using a Bio-Inspired Algorithm
Previous Article in Special Issue
Design of a Thermoacoustic Sensor for Low Intensity Ultrasound Measurements Based on an Artificial Neural Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Different Sensitive Behaviors of a Hydrogen-Bond Acidic Polymer-Coated SAW Sensor for Chemical Warfare Agents and Their Simulants

State Key Laboratory of Electronic Thin Films and Integrated Devices, School of Optoelectronic Information, University of Electronic Science and Technology of China (UESTC), Chengdu 610054, China
*
Author to whom correspondence should be addressed.
Sensors 2015, 15(8), 18302-18314; https://doi.org/10.3390/s150818302
Submission received: 30 April 2015 / Revised: 8 July 2015 / Accepted: 22 July 2015 / Published: 28 July 2015
(This article belongs to the Special Issue Acoustic Waveguide Sensors)

Abstract

:
A linear hydrogen-bond acidic (HBA) linear functionalized polymer (PLF), was deposited onto a bare surface acoustic wave (SAW) device to fabricate a chemical sensor. Real-time responses of the sensor to a series of compounds including sarin (GB), dimethyl methylphosphonate (DMMP), mustard gas (HD), chloroethyl ethyl sulphide (2-CEES), 1,5-dichloropentane (DCP) and some organic solvents were studied. The results show that the sensor is highly sensitive to GB and DMMP, and has low sensitivity to HD and DCP, as expected. However, the sensor possesses an unexpected high sensitivity toward 2-CEES. This good sensing performance can’t be solely or mainly attributed to the dipole-dipole interaction since the sensor is not sensitive to some high polarity solvents. We believe the lone pair electrons around the sulphur atom of 2-CEES provide an electron-rich site, which facilitates the formation of hydrogen bonding between PLF and 2-CEES. On the contrary, the electron cloud on the sulphur atom of the HD molecule is offset or depleted by its two neighbouring strong electron-withdrawing groups, hence, hydrogen bonding can hardly be formed.

1. Introduction

Trace level detections of harmful or toxic chemicals, such as nerve agents [1,2,3,4,5,6,7,8,9,10,11], explosives [12,13,14,15,16,17,18], and volatile organic compounds (VOCs) [19,20,21,22,23], have been intensely studied by means of a wide variety of sensing technologies. In general, application of chemical sensors in this area requires coating of a sorptive material onto the surface of specific transducers [1,2,3,4,5,6,7,8,16,17,18,19,20,21,24,25]. Considerable amounts of research on key sensing materials involve fine-tuned metal oxides, carbon nanotubes, graphene, dyes and polymers/polymer composites. Among the above materials, polymer-based efforts for the detection of organic vapors have gained wide popularity [1,2,3,4,5,6,17,18,19,20,24]. The large-area processability, as well as widely tailorable physical and chemical properties of polymeric materials has stimulated their exploitation in practical applications.
Strongly hydrogen-bond acidic (HBA) polymers are commonly a set of siloxane- or carbosilane-based molecules functionalized by fluorinated alcohol or fluorinated phenol groups [1,2,3,4,5,6,26,27,28]. Structures of some typical ones are shown in Figure 1. They are essential for the detection of nerve agents originally based on acoustic wave sensors, and have been introduced in microcantilever, chemiresistor and chemicapacitor sensors, and a number of optical approaches [29,30,31,32,33]. In principle, they can interact with the organophosphorus compounds by hydrogen bonding, which promotes the sorption of the analytes into the polymer coating on the transducer surface, and as a consequence, increases the sensor response. Sensitive properties toward nitroaromatic explosives have also been studied with these polymers [17]. The hydrogen bonding is formed by interaction between the nitro group of the explosive compound and the hydroxyl of the functional polymer.
Figure 1. Some typical previously reported HBA polymers.
Figure 1. Some typical previously reported HBA polymers.
Sensors 15 18302 g001
As a representative HBA polymer, linear functionalized polymer (PLF) was introduced by the Demathieu group in 2000, who emphasized the partition coefficients between PLF and a series of reference solutes [34]. Zimmermann initially reported a PLF-coated Love wave device for the detection of organophosphorus compounds, and the sensor showed high sensitivity and reversibility. Subsequently, dimethyl methylphosphonate (DMMP) and sarin (GB) vapor detections were compared to study the different sensitive properties based on an identical Love wave sensor, which demonstrated a higher sensitivity of the sensor to DMMP than to GB [35].
The work detailed here describes our efforts to study the different sorbent properties of a HBA polymer for chemical sensing applications based on a surface acoustic wave (SAW) device. The acoustic sensor works as a frequency-determining electrical oscillator, therefore, the response of the sensor is recognized as the shift of the resonant frequency, which is believed to be primarily attributed to the mass and modulus changes of the sensitive coating [36,37]. The feasibility and high sensitivity of the PLF coated SAW sensor to nitroaromatic explosives had been demonstrated and previously reported by our group [38]. This work focuses on the different sensitive properties of the PLF sensor to real chemical warfare agents and their stimulants, including sarin and its simulant DMMP, mustard gas (HD) and its simulant 2-chloroethyl ethyl sulphide (2-CEES) and 1,5-dichloropentane (DCP). Thereinto, the HBA polymer shows a high sensitivity to 2-CEES, while low sensitivities to HD itself and another simulant DCP. These results are firstly reported and compared here. The real-time responses of the sensor are presented and the corresponding sensitive mechanisms are discussed.

2. Experimental Section

Exposure to sarin results in high morbidity and psychological impact, but low mortality. The lethal dose of sarin causing death in 50% of persons exposed (LD50) is estimated to be 100 mg·min/m3 [39], that is 0.1 mg·min/L, so special safety protocols need to be implemented for exposure to this substance. The synthesis and characterization details of PLF were described in the previous paper [38]. The fabrication of the PLF-coated SAW sensor was performed as follows: A 434-MHz two-port SAW resonator device from Luguang Electronics (Shenzhen, China), was used. PLF was dissolved in chloroform (J & K Scientific, Beijing, China) and spray-coated onto the surface of a bare device by an airbrush apparatus. After film deposition, a frequency shift of 777 kHz was recorded, corresponding to a film thickness of about 27.9 nm. The coated SAW device had a Q value of 3000 and an insertion loss of −14 dB, monitored by a network analyzer (E5070B, Agilent Technologies, Shanghai, China). The coated device, as well as a bare one, were mounted onto a printed circuit board and independently excited in an oscillator loop. The frequency differences between the two devices were obtained by a mixer, and recorded by a SS7200A general counter (Suin Digital Instruments, Shijiazhuang, China).
The test vapors were generated by a MF-3C dynamic vapor generator (China National Metrology Technology Development Co., Beijing, China). The vapors were injected into a small test chamber, where the sensor and a bare device were located in a queue configuration. The chamber was made of aluminium alloy with gas flowing closely over the sensor printed circuit board. The tests were carried out at room temperature (25 °C), and the flow rates of all test vapors were consistently regulated at 1 L/min. The sensitive properties of the sensor were investigated by exposure to various analytes in increasing concentration steps.

3. Results and Discussion

The signals were effectively recorded by measuring the change of the shifts of frequency. One test cycle was about 6 min, corresponding to 3-min exposure time to the target vapors and 3-min to pure nitrogen.

3.1. Responses to DMMP and Sarin

We first set out to investigate the responses of the sensor to DMMP and sarin. Figure 2 shows typical responses of the sensor to DMMP at concentrations ranging from 1 to 20 mg/m3. As we can see, the response changed immediately when the chamber was purged with DMMP vapor, and positive values were obtained at any test concentration. At the lowest test concentration of 1 mg/m3, the response reached 80% maximum of the corresponding peak height within 50 s, and finally achieved a value of 3.1 kHz. When pure nitrogen was injected into the chamber, desorption took place rapidly as well and the response could return to the original value. The signals recorded in the curve showed that each adsorption–desorption test cycle exhibited an analogous square wave shape.
Figure 2. Real-time responses of the PLF sensor to DMMP at concentrations ranging from 1 to 20 mg/m3 (3 min DMMP and 3 min N2).
Figure 2. Real-time responses of the PLF sensor to DMMP at concentrations ranging from 1 to 20 mg/m3 (3 min DMMP and 3 min N2).
Sensors 15 18302 g002
The inset of Figure 2 shows the response of the sensor as a function of the vapor concentration. PLF is a viscous liquid at room temperature and the responses will be influenced by the viscoelastic effect of the polymer. Thus a parabolic fitting was chosen instead of the linear fit, which is more suitable for pure gravimetric devices like the quartz crystal microbalance (QCM). After polynomial fitting, an expression of Δf = 3024.34x25.99x2 is obtained where Δf is in Hz and vapor concentration x is in mg/m3. According to this expression, the response of the sensor at 0.009 mg/m3 is estimated to be 28 Hz. Thus, the limit of detection (LOD) of the sensor, taking into account the signal-to-noise ratio of 3:1, is deducted to be 0.009 mg/m3 since the noise level of the coated SAW sensor is around 5–9 Hz.
Figure 3 shows the responses of the sensor to sarin at the concentrations of 0.5, 1 and 10 mg/m3. Unlike DMMP, the responses of the sensor to sarin exhibited a triangular-wave curve for each sorption–desorption cycle. The response value increased linearly with the exposure time to the sarin vapor with a time-dependent sensitivity of 260 Hz/min at the concentration of 0.5 mg/m3, which indicated that PLF coating couldn’t reach an equilibrium state within an exposure of 3 min. Actually, at the highest test concentration of 10 mg/m3, the response value of the sensor could reach 26 kHz at 16 min, which wasn’t even a steady state, shown in Figure 4.
Figure 3. Real-time responses of the sensor to sarin at the concentrations of 0.5, 1 and 10 mg/m3 (3 min sarin and 3 min N2).
Figure 3. Real-time responses of the sensor to sarin at the concentrations of 0.5, 1 and 10 mg/m3 (3 min sarin and 3 min N2).
Sensors 15 18302 g003
Figure 4. Response of the sensor to sarin at the concentration of 10 mg/m3 (16 min sarin and 16 min N2).
Figure 4. Response of the sensor to sarin at the concentration of 10 mg/m3 (16 min sarin and 16 min N2).
Sensors 15 18302 g004
The response comparison of a poly(methyl-{3-[2-hydroxy-4,6-2(3methyl)]phenyl}-propyl siloxane (DKAP)-coated SAW sensor to DMMP and sarin has already been investigated by our group [2]. Like PLF, the response of the DKAP sensor to sarin is slower than to DMMP. The reason leading to the result is the size discrepancy of the two analytes and the hydrogen bond strength.

3.2. Responses to 2-CEES, DCP and HD

Figure 5 includes the response variations in terms of the concentrations of 2-CEES. Like the DMMP tests, at all test concentrations, the sensor shows fast adsorption processes during the exposure to 2-CEES vapor. Thereinto, the responses of the sensor towards 2-CEES vapor at the concentrations of 1, 2 and 5 mg/m3, could reach steady states which represent the adsorption-desorption equilibrium of the sensitive coating. Mass effects play a primary role in the frequency shift response under such conditions. However, an interesting phenomenon is onserved in that the responses decline slightly after reaching maximum values at relatively high test concentrations of 10 and 19 mg/m3. These adsorption kinetic behaviors are interpreted in terms of the viscoelastic effect, which is based on the swelling-induced modulus changes of the PLF coating [36,37]. Thus, in a sense, the modulus changes could offset the response values of the sensor induced by the mass effect. These results indicate that both the mass effect and viscoelastic effect contribute to the frequency shift response when exposed to relatively high concentrations of 2-CEES. The inset of Figure 5 also shows the response of the sensor as a function of the vapor concentration. An expression of Δf = 2841.51x − 69.52x2 is obtained. According to this expression, the response of the sensor at 0.01 mg/m3 is estimated to be 28 Hz. Thus, the LOD of the sensor to 2-CEES is deducted to be 0.01 mg/m3.
Figure 5. Response variations of the PLF sensor in terms of the concentration change of 2-CEES (3 min 2-CEES and 3 min N2).
Figure 5. Response variations of the PLF sensor in terms of the concentration change of 2-CEES (3 min 2-CEES and 3 min N2).
Sensors 15 18302 g005
Two of the common simulants for HD, 2-CEES and 1,5-dichloropentane (DCP) have been studied with various sensitive materials [40,41,42,43]. However, to the best of our knowledge, none of these studies concerned the discrepancy between the sensitive properties of the two compounds. Our work indicated that the PLF sensor showed a nearly 20 times higher response to 2-CEES than to DCP at the test concentration of 10 mg/m3, as shown in Figure 6. The inset of Figure 6 shows the response of the sensor as a function of the vapor concentration. An expression of Δf = 134.04x − 0.29x2 is obtained. According to this expression, the response of the sensor at 0.21 mg/m3 is estimated to be 28 Hz. Thus, the LOD of the sensor to DCP is deducted to be 0.21 mg/m3.
Figure 6. Real-time responses of the PLF sensor to DCP (3 min DCP and 3 min N2).
Figure 6. Real-time responses of the PLF sensor to DCP (3 min DCP and 3 min N2).
Sensors 15 18302 g006
A further study on the sensitive property of the sensor to HD has also been conducted. Figure 7 is the plot for the response of the sensor with the variation of HD concentration. Like DCP, the sensor showed a low sensitivity of about 500 Hz to HD at the test concentration of 5 mg/m3, which was only 1/20 as compared to 2-CEES. Furthermore, the response rate of the sensor to HD was much slower than to 2-CEES.
Figure 7. Responses of the PLF sensor to HD at the concentrations of 2, 5 and 20 mg/m3 (3 min HD and 3 min N2).
Figure 7. Responses of the PLF sensor to HD at the concentrations of 2, 5 and 20 mg/m3 (3 min HD and 3 min N2).
Sensors 15 18302 g007
The molecular structures of the three compounds are shown in Figure 8. Compared to HD, a chlorine atom and a sulphur atom are lacking in the molecular structures of 2-CEES and DCP, respectively. The chlorine atom shows strong electronegativity and the sulphur atom is electron-rich. The participation of the two atoms results in different polarities and electron cloud distributions of the three molecules, which may explain the discrepancy in the sensitive behavior of the sensor.
Figure 8. Molecular structures of HD, its simulants and 1-chloropentane.
Figure 8. Molecular structures of HD, its simulants and 1-chloropentane.
Sensors 15 18302 g008
In a word, the results indicate that the PLF sensor shows tremendously different sensitive behaviors to chemical warfare agents and their simulants, which means that the sensitivity studies of the sensors to the simulants of the real agents may not be suitable for practical application.

3.3. Responses to Interfering Vapors

Responses of the sensor to various vapors are shown in Figure 9 and Table 1. Thereinto, DMMP, sarin and 2-CEES are organic vapors that possess flexible electron-rich cloud around oxygen or sulphur atom. The rest interference vapors can be fallen into different categories by the functional groups in their molecular structures. Ethanol and acetone are oxidized polar hydrocarbon, pentane and toluene are low polar hydrocarbons, 1,2-dichloroethane (DCE) is halogenated non-polar hydrocarbon, 1-chloropentane is halogenated polar hydrocarbon and N,N-dimethylformamide (DMF) is nitric polar hydrocarbon.
Figure 9. Responses of the PLF sensor to various analytes at 10 mg/m3.
Figure 9. Responses of the PLF sensor to various analytes at 10 mg/m3.
Sensors 15 18302 g009
Table 1. Sensor responses to various vapors at 10 mg/m3.
Table 1. Sensor responses to various vapors at 10 mg/m3.
AcetoneEthanolDCEDMFToluenePentaneDCP
Response (Hz)test 19981059991165582767951
test 2101110891031184599759975
test 310061081981178596765969
Pressure at 25 ℃ (mmHg)224.858.499.95.531.5512.81.1
HD1-chloro-pentane2-CEESDMMPSarin
Response (Hz)test 1996209207542764226601
test 2991213207912770626655
test 3997217208372776126651
Pressure at 25 ℃ (mmHg)0.130.043.40.8330.4
The response values were obtained at a consistent test concentration of 10 mg/m3. As we can see, responses of the sensor to the three electron-rich compounds were obviously larger than to the interferent vapors. The response to water vapor was also investigated in this paper and former work [44]. Obviously, compared to the non-polar and low polar compounds, the sensor was more sensitive to the polar analytes.

3.4. Sensitive Mechanism

It is well established that the interactions between HBA polymers and organophosphorus are based on hydrogen bonding, as shown in Figure 10 [3,4,5].
Figure 10. Hydrogen-bonding interactions between DMMP, sarin, 2-CEES and HBA polymer PLF.
Figure 10. Hydrogen-bonding interactions between DMMP, sarin, 2-CEES and HBA polymer PLF.
Sensors 15 18302 g010
For the pendant group of PLF, the electron withdrawing effect of the two adjacent trifluoromethyl groups increases the activity of the hydroxyl group, which greatly promotes the hydrogen bond acidity of the HFIP group. As to DMMP and sarin, the oxygen atom shows a higher electonegativity than the phosphorus atom, thus, the strong electron-withdrawing properties lead to a high density of the electron cloud of the oxygen atom. When contacting the HBA polymer, the electron-rich oxygen atom can act as an electron donor. As a result, hydrogen bonding is formed between the pendant group of PLF and the oxygen atoms of DMMP or sarin.
As shown in Figure 9, 2-CEES showed a response of 20.8 kHz at 10 mg/m3, which is much larger than its similar counterparts HD (0.99 kHz) and DCP (0.97 kHz). The saturation vapor pressures of 2-CEES, HD and DCP are 3.4, 0.1059 and 1.125 Torr, respectively. As for pure physical adsorption, the lower the saturation vapor pressure, the easier the vapor swelling into the adsorption material [45]. Obviously the adsorption of 2-CEES is far more than physical adsorption. The reason is speculated to be that the lone pair electrons on the sulphur atom can form weak hydrogen bonding with the hydroxyl in PLF. As is known, the ethyl and the chlorethylidene groups are electron-donating and electron-withdrawing groups, respectively. For HD, the electron-withdrawing effect of the two symmetric chlorethylidene groups decreases the density of electron cloud on the sulphur atom collaboratively, which is believed to lead to the low sensitivity of the sensor.
The mechanism is confirmed by a comparison test to 1-chloropentane. The only difference between 1-chloropentane and 2-CEES is the sulphur atom in 2-CEES while a carbon atom occupies that position in 1-chloropentane. Since the PLF sensor shows a much higher sensitivity to 2-CEES than to 1-chloropentane, as shown in Figure 9, we can safely draw a conclusion that the sulphur atom plays a vital role in the 2-CEES adsorption, by providing the hydrogen-bonding site. Furthermore, Hussain et al. [46] also found that the surface hydroxyl groups of Ag/TiOx–Al2O3 absorb sulfur aromatics primarily via hydrogen-bonding, which is consistent with our findings.
Another reason might be that the active hydroxyl group in PLF can form ion-dipole interactions with 2-CEES, thus the 2-CEES molecule shows remarkable polarity arising from the asymmetry of the molecular structure, while as to HD and DCP, their polarity is very tiny. However, as shown in Figure 9, the sensor exhibited low sensitivity to ethanol, DMF and 1-chloropentane, which are all polar compounds. Thus, the dipole-dipole interaction is a synergic factor instead of a primary factor that explains to the high responses of the sensor when exposed to DMMP, sarin and 2-CEES.

4. Conclusions

A HBA polymer (PLF)-coated SAW sensor was introduced for the detection a variety of analytes, including DMMP, sarin and 2-CEES. Care was taken to investigate the sensitive properties of the sensor to these analytes. High sensitivity of the sensor is demonstrated and the reason is interpreted as the formation of hydrogen bonding between PLF and the analytes that have electron-rich atoms. Each of these compounds possesses an electron-rich site that acts as an electron donor, which can interact with the hydroxyl group of PLF by hydrogen bonding. However, the sensor shows tremendously different sensitive behaviors to real agents and their simulants, which is believed to due to the discrepancy in their molecular structures. The results show that the sensitivity study of the sensor to the simulants of the real agents is not enough for practical application. Furthermore, we will evaluate the feasibility of the sensor for the detection of other organosulphur compounds, such as dimethyl sulphide and methyl disulfide, in continuing efforts.

Acknowledgments

This work was partially supported by the National High-Tech Research & Development Program of China (Grant No. 2014AA063010), Specialized Research Fund for the Doctoral Program of Higher Education (Grant No. 20130185110017) and National Science Foundation of China via Grant No. 61235006 and 61421002.

Author Contributions

Yin Long, Yang Wang and Xiaosong Du created the major part of this paper. Luhua Cheng and Penglin Wu helped the material and film fabrication. Yadong Jiang contributed a lot during the revision process.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grate, J.W. Acoustic wave microsensor arrays for vapor sensing. Chem. Rev. 2000, 100, 2627–2648. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Y.; Du, X. The response comparison of a hydrogen-bond acidic polymer to sarin, soman and dimethyl methyl phosphonate based on a surface acoustic wave sensor. Anal. Methods 2014, 6, 1951–1955. [Google Scholar] [CrossRef]
  3. Grate, J.W. Hydrogen-bond acidic polymers for chemical vapor sensing. Chem. Rev. 2008, 108, 726–745. [Google Scholar] [CrossRef] [PubMed]
  4. Hartmann-Thompson, C.; Hu, J. Hydrogen-bond acidic hyperbranched polymers for surface acoustic wave (SAW) sensors. Chem. Mater. 2004, 25, 5357–5364. [Google Scholar] [CrossRef]
  5. Hartmann-Thompson, C.; Keeley, D.L. Hydrogen-bond acidic polyhedral oligosilsesquioxane filled polymer coatings for surface acoustic wave sensors. J. Appl. Polym. Sci. 2007, 104, 3171–3182. [Google Scholar] [CrossRef]
  6. He, W.; Liu, Z. Analytical application of poly{methyl[3-(2-hydroxy-3,4-difluoro)phenyl] propyl siloxane} as a QCM coating for DMMP detection. Talanta 2008, 76, 698–702. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, F.; Gu, H. Carbon nanotube/polythiophene chemiresistive sensors for chemical warfare agents. J. Am. Chem. Soc. 2008, 130, 5392–5393. [Google Scholar] [CrossRef] [PubMed]
  8. Fifield, L.S.; Grate, J.W. Hydrogen-bond acidic functionalized carbon nanotubes (CNTs) with covalently-bound hexafluoroisopropanol groups. Carbon 2010, 48, 2085–2088. [Google Scholar] [CrossRef]
  9. Mäkinen, M.A.; Anttalainen, O.A. Ion mobility spectrometry and its applications in detection of chemical warfare agents. Anal. Chem. 2010, 82, 9594–9600. [Google Scholar] [CrossRef] [PubMed]
  10. Li, Y.; Du, X. Improvement of column efficiency in MEMS-based gas chromatography column. RSC Adv. 2014, 4, 3742–3747. [Google Scholar] [CrossRef]
  11. Goswami, S.; Manna, A. 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]
  12. Saravanan, N.P.; Venugopalan, S. Voltammetric determination of nitroaromatic and nitramine explosives contamination in soil. Talanta 2006, 69, 656–662. [Google Scholar] [CrossRef] [PubMed]
  13. Mullen, C.; Irwin, A. Detection of explosives and explosives-related compounds by single photon laser ionization time-of-flight mass spectrometry. Anal. Chem. 2006, 78, 3807–3814. [Google Scholar] [CrossRef] [PubMed]
  14. Bonnot, K.; Bernhardt, P. Tunable generation and adsorption of energetic compounds in the vapor phase at trace levels: A tool for testing and developing sensitive and selective substrates for explosive detection. Anal. Chem. 2010, 82, 3389–3393. [Google Scholar] [CrossRef] [PubMed]
  15. Das, S.; Chatterjee, D.P. Thermo and pH responsive water soluble polythiophene graft copolymer showing logic operation and nitroaromatic sensing. RSC Adv. 2013, 3, 17540–17550. [Google Scholar] [CrossRef]
  16. Rabbany, S.Y.; Lane, W.J. Trace detection of explosives using a membrane-based displacement immunoassay. J. Immunol. Methods 2000, 246, 69–77. [Google Scholar] [CrossRef]
  17. Houser, E.J.; Mlsna, T.E. Rational materials design of sorbent coatings for explosives: Applications with chemical sensors. Talanta 2001, 54, 469–485. [Google Scholar] [CrossRef]
  18. McGill, R.A.; Stepnowski, J. The design of functionalized silicone polymers for chemical sensor detection of nitroaromatic compounds. Sens. Actuators B Chem. 2000, 65, 5–9. [Google Scholar] [CrossRef]
  19. Zhang, G.Z.; Zellers, E.T. Portable instrument employing a surface acoustic wave sensor with a regenerable reagent coating for direct measurement of 1,3-butadiene and styrene. Rev. Sci. Instrum. 1995, 66, 239–246. [Google Scholar] [CrossRef]
  20. Rowe, M.P.; Steinecker, W.H. Exploiting charge-transfer complexation for selective measurement of gas-phase olefins with nanoparticle-coated chemiresistors. Anal. Chem. 2007, 79, 1164–1172. [Google Scholar] [CrossRef] [PubMed]
  21. Tsai, C.J.; Chen, M.L. Single SnO2 gas sensor as a practical tool for evaluating the efficiency of odor control engineering at food waste composting plants. Sens. Actuators B Chem. 2012, 169, 248–254. [Google Scholar] [CrossRef]
  22. Mochalski, P.; Wzorek, B. Improved pre-concentration and detection methods for volatile sulphur breath constitutes. J. Chromatogr. B 2009, 877, 1856–1866. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, Q.F.; Milburn, R.K. Real time monitoring of hazardous airborne chemicals: A styrene investigation. J. Hazard. Mater. 2006, 132, 261–268. [Google Scholar] [CrossRef] [PubMed]
  24. Hierlemann, A.; Zellers, E.T. Use of linear solvation energy relationships for modeling responses from polymer-coated acoustic-wave vapor sensors. Anal. Chem. 2001, 73, 3458–3466. [Google Scholar] [CrossRef] [PubMed]
  25. Du, D.; Wang, J. Biomonitoring of organophosphorus agent exposure by reactivation of cholinesterase enzyme based on carbon nanotube-enhanced flow-injection amperometric detection. Anal. Chem. 2009, 81, 9314–9320. [Google Scholar] [CrossRef] [PubMed]
  26. Du, X.; Ying, Z. Synthesis and evaluation of a new polysiloxane as SAW sensor coatings for DMMP detection. Sens. Actuators B Chem. 2008, 134, 409–413. [Google Scholar] [CrossRef]
  27. Du, X.; Wang, Z. A new polysiloxane coating on QCM sensor for DMMP vapor detection. J. Mater. Sci. 2009, 44, 5872–5876. [Google Scholar] [CrossRef]
  28. Huang, J.; Jiang, Y. A new siloxane polymer for chemical vapor sensor. Sens. Actuators B Chem. 2010, 146, 388–394. [Google Scholar] [CrossRef]
  29. Grate, J.W.; Kaganove, S.N. Hybrid organic/inorganic copolymers with strongly hydrogen-bond acidic properties for acoustic wave and optical sensors. Chem. Mater. 1997, 9, 1201–1207. [Google Scholar] [CrossRef]
  30. Krech, J.H.; Rose-Pehrsson, S.L. Detection of volatile organic compounds in the vapor phase using solvatochromic dye-doped polymers. Anal. Chim. Acta 1997, 341, 53–62. [Google Scholar] [CrossRef]
  31. Johnson, S.R.; Sutter, J.M. Identification of multiple analytes using an optical sensor array and pattern recognition neural networks. Anal. Chem. 1997, 69, 4641–4648. [Google Scholar] [CrossRef]
  32. Albert, K.J.; Gill, S.D. Automatic decoding of sensor types within randomly ordered, high-density optical sensor arrays. Anal. Bioanal. Chem. 2002, 373, 792–802. [Google Scholar] [CrossRef] [PubMed]
  33. White, J.; Kauer, J.S. Rapid analyte recognition in a device based on optical sensors and the olfactory system. Anal. Chem. 1996, 68, 2191–2202. [Google Scholar] [CrossRef] [PubMed]
  34. Demathieu, C.; Chehimi, M.M. Inverse gas chromatographic characterization of functionalized polysiloxanes. Relevance to sensors technology. Sens. Actuators B Chem. 2000, 62, 1–7. [Google Scholar] [CrossRef]
  35. Zimmermann, C.; Mazein, P. Detection of GB and DMMP vapors by love wave acoustic sensors using strong acidic fluoride polymers. IEEE Sens. J. 2004, 4, 479–488. [Google Scholar] [CrossRef]
  36. Grate, J.W.; Kaganove, S.N. Comparisons of polymer/gas partition coefficients calculated from responses of thickness shear mode and surface acoustic wave vapor sensors. Anal. Chem. 1998, 70, 199–203. [Google Scholar] [CrossRef] [PubMed]
  37. Lucklum, R.; Behling, C. Role of mass accumulation and viscoelastic film properties for the response of acoustic-wave-based chemical sensors. Anal. Chem. 1999, 71, 2488–2496. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, Y.; Du, X. Detection of 2,4-dinitrotoluene using hydrogen-bond acidic polymer coated SAW sensor. Chin. Sci. Bull. 2014, 59, 2608–2612. [Google Scholar] [CrossRef]
  39. Raber, E.; Jin, A. Decontamination issues for chemical and biological warfare agents: How clean is clean enough? Int. J. Environ. Health Res. 2001, 11, 128–148. [Google Scholar] [CrossRef] [PubMed]
  40. Patil, L.A.; Deo, V.V. Sensing of 2-chloroethyl ethyl sulfide (2-CEES)-a CWA simulant-using pure and platinum doped nanostructured CdSnO3 thin films prepared from ultrasonic spray pyrolysis technique. Sens. Actuators B Chem. 2011, 160, 234–243. [Google Scholar] [CrossRef]
  41. Patil, L.A.; Deo, V.V. Improved 2-CEES sensing performance of spray pyrolized Ru-CdSnO3 nanostructured thin films. Sens. Actuators B Chem. 2014, 191, 130–136. [Google Scholar] [CrossRef]
  42. Tomchenko, A.A.; Harmer, G.P. Detection of chemical warfare agents using nanostructured metal oxide sensors. Sens. Actuators B Chem. 2005, 108, 41–55. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Fan, J.; Yu, J.; Zheng, S.; Chen, W.; Li, H.; Wang, Z.; Zhang, W. New poly (N,N-Dimethylaminoethyl Methacrylate)/polyvinyl alcohol copolymer coated QCM sensor for interaction with CWA simulants. J. ACS Appl. Mater. Interfaces 2012, 4, 944–949. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Du, X. Real-time detection of styrene using SAW sensors based on hexafluoroisopropanol group functionalized hydrogen-bond acidic polymers. Sens. Actuators B Chem. 2015, 206, 272–277. [Google Scholar] [CrossRef]
  45. Munoz, B.C.; Steinthal, G. Conductive polymer-carbon black composites-based sensor arrays for use in an electronic nose. Sens. Rev. 1999, 19, 300–305. [Google Scholar] [CrossRef]
  46. Hussain, A.H.M.S.; Yang, H. Investigation of Organosulfur Adsorption Pathways from Liquid Fuels onto Ag/TiOx–Al2O3 Adsorbents at Ambient Conditions. Energy Fuel 2013, 27, 4353–4362. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Long, Y.; Wang, Y.; Du, X.; Cheng, L.; Wu, P.; Jiang, Y. The Different Sensitive Behaviors of a Hydrogen-Bond Acidic Polymer-Coated SAW Sensor for Chemical Warfare Agents and Their Simulants. Sensors 2015, 15, 18302-18314. https://doi.org/10.3390/s150818302

AMA Style

Long Y, Wang Y, Du X, Cheng L, Wu P, Jiang Y. The Different Sensitive Behaviors of a Hydrogen-Bond Acidic Polymer-Coated SAW Sensor for Chemical Warfare Agents and Their Simulants. Sensors. 2015; 15(8):18302-18314. https://doi.org/10.3390/s150818302

Chicago/Turabian Style

Long, Yin, Yang Wang, Xiaosong Du, Luhua Cheng, Penglin Wu, and Yadong Jiang. 2015. "The Different Sensitive Behaviors of a Hydrogen-Bond Acidic Polymer-Coated SAW Sensor for Chemical Warfare Agents and Their Simulants" Sensors 15, no. 8: 18302-18314. https://doi.org/10.3390/s150818302

Article Metrics

Back to TopTop