Next Article in Journal
Voiceprint Identification for Limited Dataset Using the Deep Migration Hybrid Model Based on Transfer Learning
Previous Article in Journal
Integrity and Collaboration in Dynamic Sensor Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characteristics of Surface Acoustic Wave Sensors with Nanoparticles Embedded in Polymer Sensitive Layers for VOC Detection

Laser Department, Plasma and Radiation Physics, National Institute for Laser, Atomistilor # 409, 077125 Bucharest-Magurele, Romania
*
Author to whom correspondence should be addressed.
Sensors 2018, 18(7), 2401; https://doi.org/10.3390/s18072401
Submission received: 17 May 2018 / Revised: 25 June 2018 / Accepted: 18 July 2018 / Published: 23 July 2018
(This article belongs to the Section Chemical Sensors)

Abstract

:
Surface Acoustic Wave (SAW) sensors with several types of polymer sensing films, containing embedded Fe3O4 nanoparticles (NPs) with various dimensions and concentrations, were studied. A sensor with a sensing film consisting of the polymer alone was used for comparison. NPs with a mean diameter of 7 nm were produced by laser ablation with 5 ns pulse durations, and NPs with 13 nm diameters were obtained with a laser having 10 ps pulse durations. The properties of the Surface Acoustic Wave sensors with such sensing films were analyzed. Their response (frequency shift, sensitivity, noise and response time) to three different volatile organic components (VOCs) at various concentrations were compared with one another. The frequency shift and sensitivity increased with increasing NP concentration in the polymer for a given NP dimension and with decreasing NP diameter for a given concentration. The best results were obtained for the smallest NPs used. The SAW sensor containing 7 nm NPs had a limit of detection (LOD) of 65 ppm (almost five times better than the sensor with polymer alone), and a response time of about 9 s for ethanol.

1. Introduction

Gas sensors have gained great importance in numerous domains, such as in the detection of Chemical Warfare Agents (CWAs), control of emissions, and monitoring of various gases [1]. Several types of gas sensors are currently used, such as resistive sensors and optical devices [2]. Compared to other types of sensors, Surface Acoustic Wave (SAW) sensors have important advantages, such as high sensitivity, fast response time, low cost, ease of fabrication, room temperature operation and/or the possibility of wireless operation [1,3]. Improvements of SAW sensor properties have been achieved by using nanoporous sensitive films [4,5] or various nanostructures, such as bilayers [6,7,8], incorporated nanowires, or nanotubes [9,10].
Among the materials used for the sensitive layers of SAW sensors are polymers. Polymers have high sensitivities and fast responses for various vapors and gases, including toxic gases [11,12,13]. In particular, polyethylenimine (PEI) has proven to have a good response to CWA [14,15]. In literature, numerous methods for polymer deposition onto the surface of SAW sensors have been described, such as Laser-Induced Forward Transfer (LIFT) [16,17], spin coating [18], spray coating [19], etc.
Polymers, however, have a series of disadvantages, such as high losses, low propagation velocity or instability due to affinity to moisture [20]. On the other hand, inorganic metal oxides, which are the other class of materials often used in SAW sensors, have high sensitivities only at relatively high temperatures, and do not operate efficiently at room temperature [21]. In recent years, considerable research effort has gone into hybrid nanocomposites based on incorporation of such inorganic nanomaterials into polymer matrices [21]. Hybrid nanocomposites have led to gas sensing properties which are improved in comparison to both polymers alone and to inorganic metal oxides [22]. Such sensor layers based on polymers have a variety of properties that are better than the base polymers, mainly due to their high specific surface area, large number of sites at which gases can react [10,23], and enhancement of the film conductivity [24]. Incorporation of nanoparticles (NPs) into the polymer sensitive layer, for example, significantly improves the sensitivity of the sensor [19]. Studies have also shown that NPs improve the uniformity of nanocomposite polymer films which contain them [25], and eliminate dewetting of the layer, which is extremely detrimental in SAW sensors [26].
The present paper describes the properties of SAW sensors with nanocomposite sensitive layers consisting in a polymer (PEI) containing embedded Fe3O4 nanoparticles with different diameters and in different concentrations. The sensors were tested on Volatile Organic Compounds (VOCs). The role of the NP dimensions and concentrations in determining essential sensor properties such as frequency shift, sensitivity, noise and response time were investigated. In particular, the influence of NP dimensions on SAW sensor properties has not, to our best knowledge, been investigated.

2. Materials and Methods

The SAW sensor, designed as a two-port resonator, was fabricated on a ST-X cut quartz substrate and has an operating frequency of ~69 MHz. The SAW transducers were fabricated using photolithography and have Interdigital Transducers (IDTs) with Cr and Au metallization film thicknesses of 10 and 150 nm, respectively. Each input and output interdigital transducers consisted of 50 electrode pairs, aperture width 2500 µm, and a wavelength of ~45 µm; the distance between IDTs was 10 mm.
SAW sensors with several types of sensing films were studied, as presented in Table 1: Polymer (polyethylenimine-PEI) nanocomposite films containing Fe3O4 nanoparticles having various dimensions (7 nm, 13 nm or 50 nm) with the same mass concentration (0.4 mg NPs/mL polymer solution), and polymer nanocomposite films having the same NP dimensions (50 nm) with various concentrations of NPs (0.2, 0.4 or 0.8 mg NPs/mL polymer solution). A sensor with a sensing film consisting of the polymer alone was used for comparison.
The NPs with a mean diameter of about 7 nm were produced by laser ablation using a Nd-YAG laser (EKSPLA NL301HT, EKSPLA uab, Vilnius, Lithuania) which had pulse durations of 5 ns and a repetition rate of 10 Hz, while those with about 13 nm diameters were obtained with a Nd-YVO4 laser (Lumera Rapid, Lumera Laser GmbH, Kaiserslautern, Germany) with 10 ps pulse durations and a repetition rate of 500 kHz. A laser wavelength of 355 nm was used in both cases to obtain the nanoparticles in an irradiation chamber with a controlled He atmosphere. A Fe target was irradiated in 400 Torr He, and the resulting nanoparticles were collected on a membrane (Merck Millipore Isopore; 0.1 µm pore size; Merck HGaA, Darmstadt, Germany) placed 45 mm from the target. After deposition, oxygen was introduced into the chamber at a rate of 1 slm. More details on the deposition system are given in [27]. Figure 1 presents Transmission Electronic Microscopy (TEM) images of the nanopowders obtained in the two cases. Such TEM images were used to obtain the size distribution of the nanoparticles shown in Figure 2, which was well-fitted by a lognormal distribution. SAED (Selected Area Electron Diffraction) was used to identify the predominant composition of the NPs as Fe3O4, and their crystalline structure, as shown in Figure 3. The difference in nanoparticle dimensions was due both to the difference in pulse durations and repetition rate, which lead to a difference in interaction between the laser pulse and the ablated target and in expansion of the ablated target species between the target and the membrane on which they are collected [28,29]. The NPs which had a diameter of 50 nm were provided by NaBond Technologies Co, Ltd., Guangdong, China; the size distribution and XRD data are described in Reference [30].
The polymer used was commercially available polyethylenimine (PEI) (Sigma-Aldrich, Inc., Steinheim, Germany). The sensitive layer S1, consisting in polymer only, was obtained by mixing PEI with ethanol in a concentration of 5 mg/mL. The resulting solution was sprayed through a mask (with the role of protecting the IDTs) onto the quartz sensor substrate using an airbrush. Synthetic air was used as a carrier gas. In the case of sensitive layers S2–S6, after the corresponding mass of NPs was added to the 5 mg/mL polymer solution, the resulting suspension was sonicated for 15 min and deposited onto the substrate in the same way as S1. The thickness of the layers, as determined by WLI (White Light Interferometry), was about 250 nm.
A schematic diagram of the set-up used for sensor characterization is given in Figure 4. A CNT-91 Pendulum frequency counter, connected to a computer with Time View 3 software, was used to monitor the frequency change. A DHPVA-100 FEMTO amplifier (10–60 db, 100 MHz), was used to compensate the loss of the signal from the circuit. The response of the sensors to three different volatile organic compounds (VOCs) (ethanol, methanol and toluene) at concentrations between 160 and 16,000 ppm were determined. The acrylic mixer chamber has a volume of 8.5 L; calculated quantities of VOCs mixed with air were injected into it using a microliter pipette. The flow rate of the VOC/air mixture was maintained constant at 150 cm3/s with a diaphragm pump (Pfeiffer MVP 035-2) for all the measurements. The limit of detection (LOD) was defined as three times the noise level per sensitivity.

3. Results

Figure 5 gives the frequency shift of the sensors for the three VOCs at various concentrations. Figure 5a,c,e includes the results for the sensors having different concentrations of the same nanoparticle dimensions, while Figure 5b,d,f is for the same concentrations of different nanoparticle dimensions. All figures include the results for sample S1 with a polymer sensitive layer without NPs. As can be expected, the response increases for increasing concentrations. It is clear from all figures that the samples which contain NPs have a larger frequency shift than the one without NPs in all cases.
A comparison of the responses of the sensors which contain NPs is more visible in Figure 6, which gives the frequency shift of the 6 sensors to a concentration of 1600 ppm of the three VOCs. Sensor S3, which has the smallest nanoparticles, has the largest frequency shift for all three VOCs. Table 2 presents the sensitivities and LODs of the sensors for 1600 ppm of the VOCs. Here, again, it is evident that S3 has the best response, having the highest sensitivity and lowest LOD. On the other hand, the sensor without incorporated NPs has the lowest frequency shift and sensitivity, and highest LOD of all. It can be observed that the differences of sensitivity and LOD between different VOCs for the same sensor is significant, which proves the sensors’ selectivity. Repeating 10 measurements of the frequency deviation for each of the six sensors yielded errors below ±4%.
The response time of the sensors studied was calculated as the necessary time for the oscillation frequency to reach 90% of maximum deviation signal from the frequency base line, and is presented in Figure 7.

4. Discussion and Conclusions

By comparing the sensors with NPs having different diameters and the same concentration, it is clear that the sensor with the smallest NP mean dimensions of 7 nm (and largest number of NPs) has the best properties, the largest frequency shift, and the highest sensitivity for all three VOCs studied. In addition, it has the lowest noise level of 35 Hz (which was determined from data collected for 10 min, and represents the maximum frequency deviation from the trend line) and thus the lowest limit of detection (LOD). The sensor with NPs having 13 nm mean dimensions has the next lowest LOD, and the highest frequency shift and sensitivity in the case of methanol and toluene.
Sensors with nanocomposite sensing films containing NPs with the same mean dimensions of 50 nm have frequency shifts and sensitivities which increase with NP concentration in the domain examined (0.2–0.8 mg/mL). However, the noise level also increases with NP concentration, reaching levels larger than in the case of the sensitive layer without NPs. The noise level for an NP concentration of 0.2 mg/mL (S4) is 55 Hz, while that for 0.8 mg/mL (S6) is 90 Hz; for the sensing layer with polymer only (S1) it is 60 Hz. Thus, the NP concentration cannot be increased indefinitely to improve sensor response.
As can be seen in Figure 7, the sensor with the 7 nm mean nanoparticle diameter has the shortest response time of about 9 s, while the one with a polymer only sensitive layer has the longest response time of about 15 s; both response times are for ethanol with 1600 ppm concentration. The differences between the response times of other sensors with incorporated NPs are not significant, so that no clear conclusions regarding them can be obtained. A possible explanation for a faster response of polymer films which contain NPs is that the structure of the nanocomposite films leads to a higher diffusivity of the VOCs [5]. In Reference [31], the solvent transport within a sensing layer consisting of NPs embedded in PEI was described by classical (Fickian) diffusion. The analysis allows us to conclude that the presence of NPs leads to an increase of the diffusion coefficient in the case of PEI, which in turn leads to a greater absorption speed of the sensing layer.
When a gas is adsorbed in the sensitive layer of the SAW sensor, a series of effects are produced and lead to a frequency shift. The main effects are mass loading, acoustoelectric loading and elastic loading [5,14]. There are several possible explanations for the improved sensor properties when NPs are incorporated into the polymer sensitive film. Introduction of NPs into the polymer increases the surface area, which in turn increases the sensor sensitivity [27]. Since preferred surface sites for gas adsorption are defects, it has been proven that smaller NPs lead to a larger number of sorption sites and sites at which gases react [23,32]. In the present case, for the same NP concentration, the total surface area of the NPs in sample S3 with 7 nm mean diameter is about twice that of sample S2 with 13 nm NPs and over seven times that of sample S5 with 50 nm NPs. This could explain the differences in these samples’ sensitivities; the only exception is the slightly smaller sensitivity of S2 compared to S5 for ethanol.
NPs also modify the electrical conductivity and elastic constant of the polymer matrix they are embedded in. The electrical conductivity of polymers with embedded NPs have been proven to increase for certain values of wt % NPs [33]. It has been established that an increase in the electron conductivity of polymers can enhance the sensitivity of SAW sensors through the increase in free charge carriers, which improves the coupling between the nanocomposite sensitive layer and the electric field associated with the SAW [10]. Research has also shown that incorporation of NPs into a polymer matrix leads to a linear increase of acoustic wave velocity and elastic constant with increasing NP concentration [20]. In particular, Fe3O4 NPs have been proven to increase the rigidity of a polymer, increase its viscosity and make an insulating polymer conductive even at low concentrations [34,35]. All these changes in the properties of the sensitive film lead to complex changes in its interaction with the gas and can potentially lead to improvement of the sensor performance.
NPs have also been proven to improve the uniformity of polymer sensitive films in certain conditions [5]. This in turn leads to a lower noise, since thickness variations of the film produce a loss of coherence of the propagating wavefront and a lower signal to noise ratio [15].
In conclusion, the properties of SAW sensors with polymer sensitive layers without embedded NPs and with NPs of different dimensions and concentrations have been analyzed. Their response (frequency shift, sensitivity, noise and response time) to three different VOCs at various concentrations were compared. The frequency shift and sensitivity increase with increasing NP concentration in the polymer for a given NP dimension and with decreasing NP diameter for a given concentration. The best results were obtained for the smallest NPs used, which led to the largest frequency shift and sensitivity, lowest noise and LOD and shortest response time for all VOCs. This SAW sensor containing 7 nm NPs had a limit of detection (LOD) of 65 ppm (almost five times better than the sensor with polymer alone), and a response time of about 9 s for ethanol. The clear dependence of SAW sensor performance on the dimensions of the NPs embedded in nanocomposite polymer sensitive layers has not, to our knowledge, been studied before.

Author Contributions

Conceptualization, C.V. and D.M.; Formal analysis, C.V.; Investigation, C.V. and D.M.; Writing—original draft, D.M.; Writing—review and editing, D.M.

Funding

This research received no external funding.

Acknowledgments

This work was supported by a grant of the Romanian Ministry of Research and Innovation, CCCDI-UEFISCDI, projects number PN-III-P1-1.2-PCCDI-2017-0172/15PCCDI and project NUCLEU PN 18 13 01 01.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Devkota, J.; Ohodnicki, P.R.; Greve, D.W. SAW Sensors for Chemical Vapors and Gases. Sensors 2017, 17, 801. [Google Scholar] [CrossRef] [PubMed]
  2. Hubert, T.; Boon-Brett, L.; Black, G.; Banach, U. Hydrogen sensors—A review. Sens. Actuators B Chem. 2011, 157, 392–352. [Google Scholar] [CrossRef]
  3. Buryakov, I.A.; Buryakov, T.I.; Matsaev, V.T. Mass-Sensitive Micro- and Nanosensors for Detecting the Vapors of Explosives and Associated Substances. J. Anal. Chem. 2014, 69, 299–310. [Google Scholar] [CrossRef]
  4. Viespe, C. Surface Acoustic Wave Sensors based on Nanoporous Films for Hydrogen Detection. Materials and applications for sensors and transducers III. Key Eng. Mater. 2014, 605, 331–334. [Google Scholar] [CrossRef]
  5. Yang, L.; Yin, C.; Zhang, Z.; Zhou, J.; Xu, H. The investigation of hydrogen gas sensing properties of SAW gas sensor based on Pd surface modified SnO2 thin film. Mater. Sci. Semicond. Process. 2017, 60, 16–28. [Google Scholar] [CrossRef]
  6. Ippolito, S.J.; Ponzoni, A.; Kalantar-Zadeh, K.; Wlodarski, W.; Comini, E.; Faglia, G. Layered WO3/ZnO/36° LiTaO3 SAW gas sensor sensitive towards ethanol vapour and humidity. Sens. Actuators B Chem. 2006, 117, 442–450. [Google Scholar] [CrossRef]
  7. Jakubik, W.P. Surface acoustic wave-based gas sensors. Thin Solid Films 2011, 520, 986–993. [Google Scholar] [CrossRef]
  8. Viespe, C.; Miu, D. Surface Acoustic Wave Sensor with Pd/ZnO bilayer structure for room temperature hydrogen detection. Sensors 2017, 17, 1529. [Google Scholar] [CrossRef] [PubMed]
  9. Marcu, A.; Nicolae, I.; Viespe, C. Active surface geometrical control of noise in nanowire-SAW sensors. Sens. Actuators B Chem. 2016, 231, 469–473. [Google Scholar] [CrossRef]
  10. Penza, M.; Tagliente, M.A.; Aversa, P.; Cassano, G. Organic-vapor detection using carbon-nanotubes nanocomposite microacoustic sensors. Chem. Phys. Lett. 2005, 409, 349–354. [Google Scholar] [CrossRef]
  11. Al-Mashat, L.; Tran, H.D.; Wlodarski, W.; Kaner, R.B.; Kalanter-zadeh, K. Polypyrrole nanofiber surface acoustic wave gas sensors. Sens. Actuators B Chem. 2008, 134, 826–831. [Google Scholar] [CrossRef]
  12. Joo, B.-S.; Huh, J.-S.; Lee, D.-D. Fabrication of polymer SAW sensor array to classify chemical warfare agents. Sens. Actuators B Chem. 2007, 121, 47–53. [Google Scholar] [CrossRef]
  13. Horrillo, M.C.; Fernandez, M.J.; Fontecha, J.L.; Sagayo, I.; Garcia, M.; Aleixandre, M.; Santos, J.P.; Ares, L.; Gutierrez, J.; Garcia, I.; et al. Detection of volatile organic compounds using surface acoustic wave sensors with different polymer coatings. Thin Solid Films 2004, 467, 234–238. [Google Scholar] [CrossRef]
  14. Matatagui, D.; Marti, J.; Fernandez, M.J.; Fontecha, J.L.; Gutierrez, J.; Gracia, I.; Cane, C.; Horillo, M.C. Optimized design of a SAW sensor array for chemical warfare agents simulants detection. Procedia Chem. 2009, 1, 232–235. [Google Scholar] [CrossRef]
  15. Ballantine, D.S.; White, R.M.; Martin, S.I.; Ricco, A.J.; Zellers, E.T.; Frye, G.C.; Wohltjen, H. Acoustic Wave Sensors, Theory, Design and Physico-Chemical Applications; Academic Press: San Diego, CA, USA, 1997; pp. 300–306. [Google Scholar]
  16. Dinca, V.; Fardel, R.; Shaw-Stewart, F.; Di Pietrantonio, D.; Cannata, M.; Benetti, E.; Verona, A.; Palla-Papavlu, A.; Dinescu, M.; Lippert, T. Laser-induced forward transfer: An approach to single-step polymer microsensor fabrication. Sens. Lett. 2010, 8, 436–440. [Google Scholar] [CrossRef]
  17. Dinca, V.; Palla-Papavlu, A.; Dinescu, M.; Shaw-Stewart, F.; Lippert, T.; Di Pietrantonio, D.; Cannata, M.; Benetti, E.; Verona, A. Polymer pixel enhancement by laser-induced forward transfer for sensor applications. Appl. Phys. A 2010, 101, 559–565. [Google Scholar] [CrossRef]
  18. Wei, D.W.; Wang, L.S.; Ma, J.Y.; Jiang, H.M. Synthesis and evaluation of hexafluoroisopropanol-functionalized polysiloxane as a new coating material for sensors. J. Appl. Polym. Sci. 2012, 5, 4136–4140. [Google Scholar] [CrossRef]
  19. Viespe, C.; Grigoriu, C. Surface acoustic wave sensors with carbon nanotubes and SiO2/Si nanoparticles based nanocomposites for VOC detection. Sens. Actuators B Chem. 2010, 147, 43–47. [Google Scholar] [CrossRef]
  20. Kim, H.; Choi, Y.-J.; Kang, K.-M.; Park, H.-H. Directly patternable SnO2 thin films incorporating Pt nanoparticles. Mater. Res. Bull. 2014, 52, 6–10. [Google Scholar] [CrossRef]
  21. Kaushik, A.; Kumar, R.; Arya, S.K.; Nair, M.; Malhotra, B.D.; Bhansali, S. Organic-Inorganic Hybrid Nanocomposite-Based Gas Sensors for Environmental Monitoring. Chem. Rev. 2015, 115, 4571–4606. [Google Scholar] [CrossRef] [PubMed]
  22. Iqbal, S.; Ahmad, S. Recent developments in hybrid conducting polymers: Synthesis, applications and future prospects. J. Ind. Eng. Chem. 2018, 60, 53–84. [Google Scholar] [CrossRef]
  23. Su, P.G.; Peng, Y.-T. Fabrication of a room-temperature H2S gas sensor based on PPy/WO3 nanocomposite films by in-situ photopolimerization. Sens. Actuators B Chem. 2014, 193, 637–643. [Google Scholar] [CrossRef]
  24. Fu, C.; Lee, K.J.; Yang, S.S. Low intensity ultraviolet detection using a surface-acoustic-wave sensor with a Ag-doped ZnO nanoparticle films. Smart Mater. Struct. 2014, 24, 015010. [Google Scholar] [CrossRef]
  25. Nicolae, I.; Viespe, C.; Grigoriu, C. Nanocomposite sensitive polymeric films for SAW sensors deposited by the MAPLE direct write technique. Sens. Actuators B Chem. 2011, 158, 418–422. [Google Scholar] [CrossRef]
  26. Holmes, M.A.; Mackay, M.E.; Giunta, R.K. Nanoparticles for dewetting suppression of thin polymer films used in chemical sensors. J. Nanopart. Res. 2007, 9, 753–763. [Google Scholar] [CrossRef]
  27. Viespe, C.; Grigoriu, C. SAW sensor based on highly sensitive nanoporous palladium thin film for hydrogen detection. Microelectron. Eng. 2013, 108, 218–222. [Google Scholar] [CrossRef]
  28. Huotari, J.; Kekkonen, V.; Haapalainen, T.; Leidinger, M.; Sauerwald, T.; Puustinen, J.; Liimatainen, J.; Lappalainen, J. Pulsed laser deposition of metal nanostructures for highly sensitive gas sensor applications. Sens. Actuators B Chem. 2016, 236, 978–987. [Google Scholar] [CrossRef]
  29. Makarov, G.N. Laser Applications in nanotechnology: Nanofabrication using laser ablation and laser nanolithography. Physics-Uspekhi 2013, 56, 643–682. [Google Scholar] [CrossRef]
  30. Radu, M.; Dinu, D.; Sima, C.; Burlacu, R.; Hermenean, A.; Ardelean, A.; Dinischiotu, A. Magnetite nanoparticles induced adaptive mechanisms countered cell death in human pulmonary fibroblasts. Toxicol. In Vitro 2015, 29, 1492–1502. [Google Scholar] [CrossRef] [PubMed]
  31. Nicolae, I.; Viespe, C.; Serban, N.; Negrila, C.C.; Teodorescu, V.S.; Trupina, L. Increased Diffusion Coefficient of Polymeric Nanocomposite Layer for Gas Sensing Applications. Sens. Lett. 2013, 11, 2327–2332. [Google Scholar] [CrossRef]
  32. Dharmelingam, G.; Joy, N.A.; Grisafe, B.; Carpenter, M.A. Plasmonics-based detection of H2 and CO: Discrimination between reducing gases facilitated by material control. Beilstein J. Nanotechnol. 2012, 3, 712–721. [Google Scholar] [CrossRef] [PubMed]
  33. Van, S.N.; Hadji, R.; Vincent, B.; Rouxel, D.; Sarry, F.; Bauer, F. P(VDF-TrFE)/Al2O3 piezoelectric thin films. In Proceedings of the 2010 IEEE International Symposium on the Applications of Ferroelectrics (ISAF), Edinburgh, UK, 9–12 August 2010; pp. 1–4. [Google Scholar]
  34. Essabir, H.; Raji, M.; Essassi, E.M.; Rodrigue, D.; Bouhfid, R.; Qaiss, A. Morphological, thermal, mechanical, electrical and magnetic properties of ABS/PA6/SBR blends with Fe3O4 nano-particles. J. Mater. Sci. Mater. Electron. 2017, 28, 17120–17130. [Google Scholar] [CrossRef]
  35. Weidenfelder, B.; Hoefer, M.; Schilling, F. Thermal and electrical properties of magnetite filled polymers. Compos. Part A Appl. Sci. Manuf. 2002, 33, 1041–1053. [Google Scholar] [CrossRef]
Figure 1. Transmission Electronic Microscopy (TEM) images of the nanoparticles obtained using (a) a picosecond laser and (b) a nanosecond laser, which are used in the nanocomposite Surface Acoustic Wave (SAW) sensitive layer (S2 and S3 respectively).
Figure 1. Transmission Electronic Microscopy (TEM) images of the nanoparticles obtained using (a) a picosecond laser and (b) a nanosecond laser, which are used in the nanocomposite Surface Acoustic Wave (SAW) sensitive layer (S2 and S3 respectively).
Sensors 18 02401 g001
Figure 2. Distribution of nanoparticles (NP) diameters as obtained using TEM images in the case of (a) ps laser ablation and (b) ns laser ablation. Line is fit with a lognormal function.
Figure 2. Distribution of nanoparticles (NP) diameters as obtained using TEM images in the case of (a) ps laser ablation and (b) ns laser ablation. Line is fit with a lognormal function.
Sensors 18 02401 g002
Figure 3. TEM image and SAED of a nanoparticle.
Figure 3. TEM image and SAED of a nanoparticle.
Sensors 18 02401 g003
Figure 4. Experimental set-up used for measurements of sensor response to various Volatile Organic Compounds (VOCs).
Figure 4. Experimental set-up used for measurements of sensor response to various Volatile Organic Compounds (VOCs).
Sensors 18 02401 g004
Figure 5. Frequency shift of the sensors for different concentrations of ethanol, methanol and toluene VOCs. (a,c,e) compare sensors all having 50 nm NPs and different NP concentrations, and sensor S1 with no NPs. (b,d,f) compare sensors with the same NP concentration of 0.4 mg/mL and different NP diameters, and sensor S1 with no NPs. See Table 1 for sensor identification.
Figure 5. Frequency shift of the sensors for different concentrations of ethanol, methanol and toluene VOCs. (a,c,e) compare sensors all having 50 nm NPs and different NP concentrations, and sensor S1 with no NPs. (b,d,f) compare sensors with the same NP concentration of 0.4 mg/mL and different NP diameters, and sensor S1 with no NPs. See Table 1 for sensor identification.
Sensors 18 02401 g005
Figure 6. Frequency shift of sensors at 1600 ppm concentration of ethanol, methanol and toluene.
Figure 6. Frequency shift of sensors at 1600 ppm concentration of ethanol, methanol and toluene.
Sensors 18 02401 g006
Figure 7. Response time of the devices at a concentration of 1600 ppm of ethanol.
Figure 7. Response time of the devices at a concentration of 1600 ppm of ethanol.
Sensors 18 02401 g007
Table 1. Characteristics of nanocomposite sensitive layers of the sensors studied.
Table 1. Characteristics of nanocomposite sensitive layers of the sensors studied.
SensorMean NP Diameter (nm)NP Concentration (mg/mL)
S1Polymer (PEI) only0
S2130.4
S370.4
S4500.2
S5500.4
S6500.8
Table 2. Sensor characteristics for the three VOCs studied.
Table 2. Sensor characteristics for the three VOCs studied.
SensorSensitivity (Hz/ppm)LOD (ppm)
EthanolMethanolTolueneEthanolMethanolToluene
S10.560.500.69320360262
S20.970.811.4313916695
S31.631.001.946510554
S40.810.531.00203311165
S51.060.661.13212343200
S61.130.881.31240309206

Share and Cite

MDPI and ACS Style

Viespe, C.; Miu, D. Characteristics of Surface Acoustic Wave Sensors with Nanoparticles Embedded in Polymer Sensitive Layers for VOC Detection. Sensors 2018, 18, 2401. https://doi.org/10.3390/s18072401

AMA Style

Viespe C, Miu D. Characteristics of Surface Acoustic Wave Sensors with Nanoparticles Embedded in Polymer Sensitive Layers for VOC Detection. Sensors. 2018; 18(7):2401. https://doi.org/10.3390/s18072401

Chicago/Turabian Style

Viespe, Cristian, and Dana Miu. 2018. "Characteristics of Surface Acoustic Wave Sensors with Nanoparticles Embedded in Polymer Sensitive Layers for VOC Detection" Sensors 18, no. 7: 2401. https://doi.org/10.3390/s18072401

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

Article Metrics

Back to TopTop