Next Article in Journal
A Method Based on Curvature and Hierarchical Strategy for Dynamic Point Cloud Compression in Augmented and Virtual Reality System
Next Article in Special Issue
Identification of the Principle of Taste Sensors to Detect Non-Charged Bitter Substances by 1H-NMR Measurement
Previous Article in Journal
Gold Nanoparticles and Plant Pathogens: An Overview and Prospective for Biosensing in Forestry
Previous Article in Special Issue
Electrical Properties of Two Types of Membrane Component Used in Taste Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon SH-SAW-Based Electronic Nose to Discriminate and Classify Sub-ppm NO2

by
Carlos Cruz
1,2,*,
Daniel Matatagui
1,3,*,
Cristina Ramírez
4,
Isidro Badillo-Ramirez
5,6,
Emmanuel de la O-Cuevas
5,7,
José M. Saniger
5 and
Mari Carmen Horrillo
1
1
Grupo de Tecnología de Sensores Avanzados (SENSAVAN), Instituto de Tecnologías Físicas y de la Información (ITEFI), CSIC, 28006 Madrid, Spain
2
Department of Electronics, University of Alcala, 28871 Alcala de Henares, Madrid, Spain
3
Dpto. de Ingeniería Informática, Escuela Politécnica Superior, Universidad Autónoma de Madrid, 28049 Madrid, Spain
4
Institute of Ceramics and Glass, ICV-CSIC, Kelsen 5, Cantoblanco, 28049 Madrid, Spain
5
Instituto de Ciencias Aplicadas y Tecnología, Universidad Nacional Autónoma de México, Circuito Exterior S/N, Ciudad Universitaria, Ciudad de Mexico 04510, Mexico
6
Center for Intelligent Drug Delivery and Sensing Using Microcontainers and Nanomechanics (IDUN), Department of Health Technology, Technical University of Denmark, 2800 Kongens Lyngby, Denmark
7
Unidad Académica de Física, Universidad Autónoma de Zacatecas, Zacatecas 98068, Mexico
*
Authors to whom correspondence should be addressed.
Sensors 2022, 22(3), 1261; https://doi.org/10.3390/s22031261
Submission received: 2 December 2021 / Revised: 18 January 2022 / Accepted: 1 February 2022 / Published: 7 February 2022

Abstract

:
In this research, a compact electronic nose (e-nose) based on a shear horizontal surface acoustic wave (SH-SAW) sensor array is proposed for the NO2 detection, classification and discrimination among some of the most relevant surrounding toxic chemicals, such as carbon monoxide (CO), ammonia (NH3), benzene (C6H6) and acetone (C3H6O). Carbon-based nanostructured materials (CBNm), such as mesoporous carbon (MC), reduced graphene oxide (rGO), graphene oxide (GO) and polydopamine/reduced graphene oxide (PDA/rGO) are deposited as a sensitive layer with controlled spray and Langmuir–Blodgett techniques. We show the potential of the mass loading and elastic effects of the CBNm to enhance the detection, the classification and the discrimination of NO2 among different gases by using Machine Learning (ML) techniques (e.g., PCA, LDA and KNN). The small dimensions and low cost make this analytical system a promising candidate for the on-site discrimination of sub-ppm NO2.

Graphical Abstract

1. Introduction

Chemical sensors play a relevant role in our modern society for mobile applications, traffic safety and health care. One of the principal functions of these sensors is the monitoring of chemical compounds. It has been becoming increasingly challenging in several applications related to air quality assessment [1,2,3,4,5,6,7] and medical diagnostics [8,9]. These needs have led to the emergence of new generations of low cost, portable and reliable gas sensor devices with high potential discrimination among low concentrations of analytes of interest. The combustion of fossil fuels is the major source of nitrogen oxide (NOx) emissions into the atmosphere. In addition, the long-term exposure to NO2 levels produces harmful effects for humans and living beings, which can be detected by deploying chemical sensors based on carbon material [10,11,12].
Surface acoustic wave (SAW) devices are a type of chemical sensor with high potentialities to detect low gas concentrations at room temperature (RT) due to the high sensitivity and reduced size. In addition, novel sensing applications open the possibility for custom SAW devices [13,14,15,16,17]. SAW sensors are mainly used as gravimetric (mass sensitive) transducers. Adsorption of molecules on the SAW surface produces an increase in density per unit area that modulates the acoustic wave velocity [18,19], which is most commonly measured through a shift of the oscillation frequency. In recent years, elastic sensitivity has been presented as a novel way to measure gases by a small shift in elastic properties of nanostructured layers [11,20,21,22,23]. Several studies to achieve sensors based on nanostructured sensitive layers had been performed [24,25,26], resulting in highly sensitive and selective sensors to improve electronic nose technology for detecting, discriminating and classifying target gases. For example, Bhasker et al. [27] conclude that both mass loading and change in elasticity, are the sensing mechanisms in ammonia detection. On the other hand, Yongliang et al. [28] emphasize the benefits of the elastic loading effect in the SAW sensor to a more significant response to H2S gas with excellent stability, selectivity and humidity resistance. For instance, SAW carbon-based sensors configuration could make it possible to operate under mass and elastic regimens [22].
Pattern recognition methods are commonly applied to the sustainable environment domain [29] and can be performed for classification and discrimination of the sensor field as an important component of the e-nose. Principal component analysis (PCA) is applied for dimensionality reduction features that contain most of the patterns for discriminative information [30,31,32,33]. Artificial neural network algorithms are widely used for pattern recognition of chemical compounds [34]. The fast training and robustness make the suggested patterns procedure an excellent classifier for sensor array data [35]. High accuracy of up to 90% confirms the effectiveness of the classification analysis, especially for the e-nose platforms [36,37]. Linear discriminant analysis (LDA) or K nearest-neighbor (KNN) are also well-known machine learning (ML) approaches for pattern recognition, able to assign class descriptors to the analytes with a high degree of confidence [38,39,40,41]. However, the system needs to be trained with known samples and the results are influenced by the selected samples tested.
In this work, we analyze the e-nose capabilities of different carbon-based sensors for selective NO2 classification. In addition, with the ML techniques, we perform both the extraction of distinctive information to discriminate the observed analytes and the carbon-based sensor potential for low concentration detection of NO2. The sensing performance of the developed e-nose operates at RT and under different exposures to analytes, achieving NO2 selectivity against interfering gases such as CO, NH3, C6H6 and C3H6O. Here, we demonstrate that the elastic and the mass loading effects of carbon-based sensors are suitable for NO2 discrimination by supervised and unsupervised ML techniques in a versatile and compact system.

2. Materials and Methods

2.1. Materials

Graphene oxide (GO) powder was purchased from Graphenea (San Sebastian, Spain). Dopamine (DA) hydrochloride and Tris buffer were purchased from Sigma Aldrich (St. Louis, MO, USA). Graphitized mesoporous carbon (MC) and Gum Arabic (GA) powders were purchased from Sigma Aldrich (Madrid, Spain).

2.2. SH-SAW Sensors Fabrication

SH-SAW was propagated on the ST-cut quartz substrate (perpendicular to the x crystallographic axis). A wave with λ = 28 μm was generated and detected by interdigital transducers (IDTs) with a double electrode configuration. The IDTs, with aluminum of 200 nm thickness, were made by RF sputtering combined with standard lithographic techniques. The distance center to center between IDTs and the acoustic aperture were 150 and 75 λ, respectively. The guided SH-SAW, well-known as the Love wave, was achieved using a 3.5 µm film of SiO2 grown on the piezoelectric substrate by plasma-enhanced chemical vapor deposition, obtaining a resonance frequency of around 160 MHz.

2.3. Sensitive Layer Fabrication

In this research, we developed an array of four sensors using the following carbon-based materials as sensitive layers:
  • Mesoporous Carbon (MC). Gum Arabic (GA) was dissolved in deionized (DI) water to prepare a solution of 2 mg/mL. Then, 1 mg of MC was added to prepare a dispersion using an extended low power ultrasonication bath (Branson 5510) for 7 h. Afterwards, the dispersion was deposited on the SH-SAW sensor by spray coating in order to obtain a reproducible deposition method.
  • Graphene Oxide (GO). An aqueous dispersion of GO (0.5 mg/mL) was prepared by mixing GO powder with DI water:methanol (1:5 volumetric ratio), then the mixture was sonicated for 15 min. The dispersion was used immediately for deposition using the Lagmuir–Blodgett (LB) technique.
  • Reduced GO (rGO). The reduction of graphene oxide (reduced graphene oxide, rGO) was performed in a U reactor with a fritted plate 1.5 cm in diameter, where 100 mg of GO powder was heated at 400 °C for 2 h under constant flow of a 100 mL·min−1 H2/Ar (10% H2 balance Ar) mixture. Then, the reaction was allowed to cool until it reached RT. The synthesized product was a fine black powder. Further, a homogeneous rGO dispersion was obtained by mixing approximately 2.0 mg of rGO powder with 8 mL of methanol and sonicated for 4 h at RT.
  • GO and rGO Langmuir–Blodgett film formation. Langmuir–Blodgett (LB) deposition on the sensor chip was performed in a trough KSV 5000 alternating multilayer LB system (KSV Finland). The trough was first cleaned with chloroform and then filled with ultrapure water (Milli-Q system, 18.2 MΩ cm and simplicity 185, Millipore, Burlington, MA, USA). The rGO/GO dispersion was spread onto the water surface dropwise using a glass syringe at a speed of ~100 μL/min. The surface pressure was monitored using a tensiometer attached to a Wilhelmy plate. Barriers compressed the film at a speed of 15 mm/min. The rGO monolayer was transferred to an SH-SAW sensor using the vertical lifting method at 22 °C. Z-type multilayer structures were prepared by the vertical deposition method at a dipping speed of 10 mm/min. The same procedure was followed to prepare GO monolayer substrates.
  • Polydopamine (PDA) film formation on an rGO chip. PDA was obtained by dissolving DA hydrochloride powder in Tris buffer (pH 8.6) to obtain a DA concentration of 0.01 M. Then, the previously fabricated rGO (LB) sensor chip was immediately immersed in the DA/Tris solution at RT for 30 min. The PDA/rGO chip was rinsed three times with ultrapure water and dried in a dissector.

2.4. Structural and Morphological Characterization

Microstructural characterization of the sensitive layers was performed by field emission scanning electron microscopy (FESEM, S-4700, Hitachi, Barcelona, Spain) and Raman microscopy (Alpha 300RA, WITec GmbH, Ulm, Germany). Raman measurements were acquired with a 532 nm laser excitation wavelength with an integration time of 0.6 s and 10 accumulations; three measurements were acquired at different positions in the central region of the layer and averaged.

2.5. Experimental Setup for Gas Measurement

The SH-SAW sensor array was characterized by an automatic and controlled gas line for sensor characterization in groups of different gaseous environments. The gas sample generator (Figure 1) consisted of three mass flow controllers that obtain desirable concentrations regulating the flow rates of both the synthetic air and the chosen target. The process was performed by switching between a gas sample for 2 min (exposition time) and the synthetic air for 20 min (purge time) at a constant flow of 100 mL/min. Target gases were supplied by gas cylinders with a selectable concentration balanced with the carrier gas, all of them from Nippon Gases: NO2 (1 ppm), CO (10 ppm), NH3 (50 ppm), benzene (C6H6) (50 ppm) and acetone (C3H6O) (50 ppm). It was also possible to distinguish among concentrations with higher sensor responses in terms of increasing concentration steps. Then, the initial sample concentration was diluted with synthetic air to obtain lower concentrations such as 0.1–0.6 ppm for NO2, 1–6 ppm for CO and 10–40 ppm for NH3, benzene (C6H6) or acetone (C3H6O).

2.6. Electronic Nose Configuration

The system (e-nose) was deployed in a modular configuration:
  • The signal conditioning module. The SH-SAW signals were generated using a set of microwave circuits that consisted of two amplification states and a directional coupler.
  • The multiplexor module. The signal of each sensor oscillator was selected and forwarded to a single output, which was mixed with an oscillator-based reference signal, obtaining a new signal around 1 MHz. The reference device also compensates for external disturbances such as changes of temperature.
  • The acquisition and transmission module. A microcontroller was used to measure the resulting signal of the multiplexor module. The e-nose was kept at controlled RT while variations of the frequency over time were recorded by wireless communication (XBEE protocol). The experiment control, the real time data acquisition and the classification analyses were implemented with a PC using the LabVIEW and the Matlab software, respectively.

3. Results and Discussion

3.1. Electrical Characterization

The standard deviation (SD) obtained for the measured SH-SAW sensor signal was 5 Hz. Considering the minimum signal as three times higher than the noise signal, the minimum detectable measurement is a frequency shift of 15 Hz.

3.2. Morphological and Spectroscopic Characterization of Carbon-Based Sensitive Layers

Figure 2 shows the morphology of the carbon-based materials (CBMs) deposited on the quartz substrates. A MC nanostructured layer was created by the combination of stable aqueous dispersion of the particles and air-brush spray deposition technique (Figure 2a). Particle distribution is identified with particle sizes below 100 nm and pores in the order of 100–200 nm. Figure 2b shows the GO layer, where there is deposition of completely extended sheets favored by the LB technique, showing a flat surface with wrinkles, edges and some particular bumped features of 300 nm in diameter [42]. On the other hand, Figure 2b,d shows the sensitive layers based on rGO and PDA/rGO, respectively. The thermal reduction of GO originated the crumpling of rGO sheets, though the use of a PDA treatment had a smoothening effect. It is important to notice that although the quartz surface was effectively coated due to the material’s morphology and deposition technique, the MC layer thickness is expected to be far larger than GO films.
The Raman spectrum of MC (Figure 3 black) presents the three characteristic bands of graphitized materials: D, G and 2D vibrational modes, with moderately ordered structure (ID/IG 0.57). Similarly, GO, rGO and PDA/rGO films show broad D and G Raman bands of similar intensities, indicating the high disorder of graphitic domains obtained by the synthesis route. ID/IG ratios of 2.9 and 3.0 were calculated for rGO and PDA/rGO, respectively, whereas GO presented approximately half of the value. This increment in disorder ratio is associated to the removal of functional groups produced by thermal reduction. In accordance with SEM observations, the tendency to form crumples by rGO could lead to an uneven substrate coating.

3.3. Gas Sensor Characterization

The measured gas could be detected in a complex environment surrounded by gases for different concentrations, requiring sensors with properties of sensitivity and selectivity to discriminate and correctly classify the gases. Among the most important interfering gases, for the above-cited applications where NO2 detection is required, are CO, NH3, benzene (C6H6) and acetone (C3H6O).
Experimental measurements for the gas characterization of the sensitive materials showed that sensor response is dependent on the composition of the sensitive layer. As shown in Figure 4, sensors based on MC and PDA/rGO behaved as gravimetric sensors, decreasing their frequency due to an increase in the density per unit area of the sensing layer as a consequence of the adsorbed molecules. However, the sense of the frequency shift in rGO and GO sensors was the opposite, increasing their frequency due to the change in elastic properties of the nanostructured sensitive layers caused by the interaction with gas target molecules.
The sensitive layers in the developed sensors are different in both surface chemistry and microstructure, which play a key role in the diverse sensing mechanisms that take place between the material surface and the tested gases. The MC layer contains both a large porous surface area per volume ratio and a highly ordered graphitized carbon structure, which provides a high interaction for the of gas adsorption, mainly through physisorption processes, even if chemical adsorption processes might also take part; especially for molecules with easy electronic delocalization, such as NO2, NH3 and C6H6. The GO layer, besides containing a large surface area, also has a high number of functional groups with oxygen due to the oxidation process of the graphene, and therefore it provides many chemical reactive sites for selective gaseous molecule adsorption which causes a large transfer of electrons, such as NO2 and NH3 [43]. Moreover, the rGO layer contains a large site of ordered aromatic groups and in addition shows a reduced number of oxygenated functional groups, in comparison with GO due to the reduction process, which reduces the chemical adsorption of many of the studied gases, but improves the π–π electrostatic interactions with aromatic molecules, such as benzene. The PDA/rGO layer contains different types of functional groups, such as aromatic, amino and hydroxy, which are linked with the reactive gas molecules, in a similar way to the GO with the oxygenated groups; however, the obtained gas adsorption performances are lower. This result might be due to the high number of unordered stacked aromatic groups and hindered reactive oxygenated groups in PDA after a thick layer formation.
The MC sensor presented a high mass response to all gases. However, the rGO and GO sensors showed an elastic response for all of them. Furthermore, the GO sensor exhibited a significant elastic response with NO2 gas. This finding allowed us to compare the rGO and GO sensors and determine their elastic response depending on functional groups. The relation between responses of rGO and GO sensors can be a useful tool to discriminate NO2 among other interfering gases.
The response time was determined by defining τ90 as the time taken to reach 90% of the maximum frequency shift of the response. The MC and GO sensors presented a notable response for 0.4 ppm of NO2 after exposition of 120 s with a complete recovery. On the contrary, rGO and PDA/rGO sensors barely offer a noticeable response approach after NO2 exposition, reaching its maximum value at 80 s. (Figure 4a). All sensors showed a response for NO2, between 300 (sensors with sensitive layer composed of GO or MC) and 50 Hz (sensors with rGO and PDA/rGO). Furthermore, the limit of detection (LOD) achieved for NO2 was ~10 ppb on the MC sensor, ~7.5 ppb on the rGO, ~75 ppb on the GO and ~30 ppb on the PDA/rGO.
Figure 5a shows the performance of the developed sensors for different concentration levels of NO2. The MC and rGO sensors achieved a high linearity response (0.98 and 0.89, respectively, adjusted R squared values) in comparison with the GO and PDA/rGO sensors (0.67 and 0.76, respectively), which shows a saturation regime behavior from 0.4 ppm of NO2 in advance. To verify that sensor materials are suitable as NO2 gas sensors we further investigate the sensors’ behavior by measuring the response in short periods (2 min) and continuous cycles of exposition and purging. Figure 5b shows the reproducibility obtained at 0.1 ppm of NO2. A high and stable response was obtained for MC or GO materials.
The humidity influence on the sensor performance was also evaluated for discrimination capabilities. Figure 6 illustrates the sensor effect with a 20% relative humidity (RH) concentration measured with a handheld thermohygrometer (RS1364) at 21 °C. In the tested humid environment, the frequency decreased in the GO sensor, which implies a mass sensitivity mechanism. This result differs from the cases of the measured toxic gases where the frequency increased due to the elastic sensitivity mechanism (Figure 5). Accordingly, the response set from GO, rGO and PDA/GO sensors provides information valuable for distinguishing among dry and humid environments. The high detection level of sub-ppm NO2 makes MC and GO excellent materials with high performance for toxic gas discrimination, taking rGO and PDA/rGO as references for dry and humid environments, respectively. The negligible number of hydroxyl groups in the rGO sensitive layer implies a non-existent response to high humidity environments. However, after the rGO is covered with PDA the number of hydroxyl groups increased, presenting a high sensitivity to humidity [44,45]. For the tested humid environment, the frequency decreased in the GO sensor, which implies a mass sensitivity mechanism. This result differs from the cases of the measured toxic gases where the frequency increased due to the elastic sensitivity mechanism. Accordingly, the response set from GO, rGO and PDA/GO sensors provide information for distinguishing between dry and humid environments.

3.4. Statistical Treatment and Classification Analysis

Figure 7 shows the responses of the four SH-SAW sensors for each gas concentration, NO2 (0.2 ppm), CO (2 ppm), NH3 (20 ppm), benzene (C6H6) (20 ppm) and acetone (C3H6O) (20 ppm). In this context, the responses were normalized regarding the highest gas response for each sensor. The MC, GO, rGO sensors achieved the shortest response time for CO. Furthermore, the strong point of the elastic loading effect of the GO sensor allowed to obtain a different behavior for NO2 detection, facilitating NO2 discrimination with respect to the rest of the interfering gases.
We configured three ML techniques for pattern discrimination to identify sensors’ behaviors as part of different gas environments. The analysis with the unsupervised methods was focused on PCA visualization. It allowed us to investigate data dimensionality reduction and discriminate between different target gases. PCA analysis provided summarized data and visualized the information from the results containing the individual responses. It is a promising method in case of no labeled data. The scores of the two classes for the most important components—PC1 and PC2—are represented in Figure 8a, which offers a way to study the statistical discrimination and provides a cumulative contribution rate of 87%. The highest possible variation is explained by the first and second components (65% and 23%, respectively). Although low concentrations of both interfering gases, NH3 and acetone (C3H6O), display scattered data, PCA confirmed well-defined patterns in the sensor array response and showed a clear discrimination for NO2.
We also deployed two supervised ML techniques to validate the clustering of the PCA offered. This fact helped us to assure the most adequate classification method to assign the profiles to a group of responses. We conducted an (i) LDA that is a signal classification technique that directly maximizes class separability of different gases, and (ii) KNN, which is applied for classification of mixed gases providing high accuracy results. Figure 8b plots the results of the LDA projections in which the point feature set is defined by different responses. The classifier looks for the linear separability among NO2, interferents and humidity classes, with the examples of each class forming compact clusters and being far from each other. The result for the accuracy value (LDA classifier) was 0.9 (Figure 8b). LDA provided similar results to the PCA and, consequently, both seem to be appropriate methods for pattern discrimination in this gas sensing application.
KNN is a powerful rule to generate highly nonlinear classifications with a set of data. Figure 9 shows the KNN classification performance over 38 samples from the sensors’ responses. Table 1 shows true-quality classification indices for NO2, interferents and humidity. The final result presents an accuracy of 0.95 for the three classes and a value up to 0.9 on both F1 and precision quality scores. The KNN classifier correctly identified NO2 (100%). On the contrary, two false positives were obtained for the interferents class mainly due to the heterogeneous behavior of the responses at low concentrations.
The presented study has demonstrated that a suitable sensitive layer allows measuring changes in the velocity of the SAW that would be then selectively assigned to the adsorption of the different gases. Table 2 illustrates the comparison of our developed sensitive layers with other sensing films SAW sensors reported in the literature for NOx target gas. It also shows that our materials can perform both mass and elastic detection for NO2 and can reach similar and improved sub-ppm detection in comparison with the previous reported SAW sensors [13,14,15,16], within less than 1 min response time. Furthermore, our SAW sensors show excellent sensitivity by mass loading effect for carbon monoxide, ammonia, acetone and benzene, which can extend the environmental gas sensing applications.

4. Conclusions

For the present study, we used a portable array of SH-SAW sensors for a real-time monitoring system in order to provide discrimination of toxic chemicals. The system was designed using a modular architecture, which makes it a very self-contained and versatile platform incorporating ML capabilities for gas detection such as PCA, LDA or KNN.
The e-nose successfully detected nitrogen dioxide (NO2), monoxide (CO), ammonia (NH3), benzene (C6H6) and acetone (C3H6O), in a wide variety of concentrations, with high selectivity and sensitivity, showing a pattern for each toxic agent and high efficiency to discriminate between interfering gases and NO2. The sensor array was effective at detecting NO2 even at a low concentration (0.1 ppm). On the other hand, the elastic-loading effect on the GO-based sensor shows a characteristic behavior for NO2 which, together with the other sensors, allows NO2 to be discriminated from other interferents. The PCA plot data clustering did not present a significant discrimination between the interfering gases, but it showed that NO2 was discriminated from these other gases. Thus, PCA did not provide a clear discrimination among interfering gases with high selectivity and sensitivity, mainly due to the similar responses of acetone or ammonia at low concentrations. The SH-SAW sensors showed different sensitivities at room temperature, excellent repeatability and fast responses, and therefore all of them are considered good candidates for compact array sensors (e-nose) for environmental applications.

Author Contributions

C.C.: Data curation, methodology, investigation, writing—original draft. D.M.: Conceptualization, project administration, data curation, methodology, investigation, writing—review and editing. C.R., I.B.-R., E.d.l.O.-C., J.M.S.: Methodology, investigation, writing—review and editing. M.C.H.: Project administration, methodology, investigation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

Spanish Ministry of Science and Innovation for financing the project RTI2018-095856-B-C22 (AEI/FEDER).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

D.M. acknowledges the financial support from the Fundación General CSIC via Programa ComFuturo. C.R acknowledge the funding from IJCI-2017-34724 grant.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PPy: polypyrrole; PZT: lead zirconate titanate.

References

  1. Tan, X.; Han, L.; Zhang, X.; Zhou, W.; Li, W.; Qian, Y. A review of current air quality indexes and improvements under the multi-contaminant air pollution exposure. J. Environ. Manag. 2021, 279, 111681. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, T.M.; Kuschner, W.G.; Gokhale, J.; Shofer, S. Outdoor air pollution: Nitrogen dioxide, sulfur dioxide, and carbon mon-oxide health effects. Am. J. Med. Sci. 2007, 333, 249–256. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal Oxide Gas Sensors: Sensitivity and Influencing Factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Han, X.; Naeher, L.P. A review of traffic-related air pollution exposure assessment studies in the developing world. Environ. Int. 2006, 32, 106–120. [Google Scholar] [CrossRef] [PubMed]
  5. SM, S.N.; Yasa, P.R.; Narayana, M.V.; Khadirnaikar, S.; Rani, P. Mobile monitoring of air pollution using low-cost sen-sors to visualize spatio-temporal variation of pollutants at urban hotspots. Sustain. Cities Soc. 2019, 44, 520–535. [Google Scholar] [CrossRef]
  6. Covington, J.A.; Marco, S.; Persaud, K.C.; Schiffman, S.S.; Nagle, H.T. Artificial Olfaction in the 21st Century. IEEE Sens. J. 2021, 21, 12969–12990. [Google Scholar] [CrossRef]
  7. Shaw, C.; Boulic, M.; Longley, I.; Mitchell, T.; Pierse, N.; Howden-Chapman, P. The association between indoor and outdoor NO2 levels: A case study in 50 residences in an urban neighbourhood in New Zealand. Sustain. Cities Soc. 2020, 56, 102093. [Google Scholar] [CrossRef]
  8. Shea, K.M.; Truckner, R.T.; Weber, R.W.; Peden, D. Climate change and allergic disease. J. Allergy Clin. Immunol. 2008, 122, 443–453. [Google Scholar] [CrossRef]
  9. Bahos, F.A.; Sainz-Vidal, A.; Sánchez-Pérez, C.; Saniger, J.M.; Gràcia, I.; Saniger-Alba, M.M.; Matatagui, D. ZIF Nanocrystal-Based Surface Acoustic Wave (SAW) Electronic Nose to Detect Diabetes in Human Breath. Biosensors 2018, 9, 4. [Google Scholar] [CrossRef] [Green Version]
  10. Vanotti, M.; Poisson, S.; Soumann, V.; Quesneau, V.; Brandès, S.; Desbois, N.; Yang, J.; André, L.; Gros, C.P.; Blondeau-Patissier, V. Influence of interfering gases on a carbon monoxide differential sensor based on SAW devices functionalized with cobalt and copper corroles. Sens. Actuators B Chem. 2021, 332, 129507. [Google Scholar] [CrossRef]
  11. Zhu, H.; Xie, D.; Lin, S.; Zhang, W.; Yang, Y.; Zhang, R.; Shi, X.; Wang, H.; Zhang, Z.; Zu, X.; et al. Elastic loading enhanced NH3 sensing for surface acoustic wave sensor with highly porous nitrogen doped diamond like carbon film. Sens. Actuators B Chem. 2021, 344, 130175. [Google Scholar] [CrossRef]
  12. Matatagui, D.; López-Sánchez, J.; Peña, A.; Serrano, A.; del Campo, A.; de la Fuente, O.R.; Carmona, N.; Navarro, E.; Marín, P.; Horrillo, M.D.C. Ultrasensitive NO2 gas sensor with insignificant NH3-interference based on a few-layered mesoporous graphene. Sens. Actuators B Chem. 2021, 335, 129657. [Google Scholar] [CrossRef]
  13. Raj, V.B.; Nimal, A.T.; Tomar, M.; Sharma, M.U.; Gupta, V. Novel scheme to improve SnO2/SAW sensor performance for NO2 gas by detuning the sensor oscillator frequency. Sens. Actuators B Chem. 2015, 220, 154–161. [Google Scholar] [CrossRef]
  14. Rana, L.; Gupta, R.; Tomar, M.; Gupta, V. ZnO/ST-Quartz SAW resonator: An efficient NO2 gas sensor. Sens. Actuators B Chem. 2017, 252, 840–845. [Google Scholar] [CrossRef]
  15. Hung, T.-T.; Chung, M.-H.; Chiu, J.-J.; Yang, M.-W.; Tien, T.-N.; Shen, C.-Y. Poly(4-styrenesulfonic acid) doped polypyrrole/tungsten oxide/reduced graphene oxide nanocomposite films based surface acoustic wave sensors for NO sensing behavior. Org. Electron. 2021, 88, 106006. [Google Scholar] [CrossRef]
  16. Rana, L.; Gupta, R.; Kshetrimayum, R.; Tomar, M.; Gupta, V. Fabrication of surface acoustic wave based wireless NO2 gas sensor. Surf. Coat. Technol. 2018, 343, 89–92. [Google Scholar] [CrossRef]
  17. Ghosh, A.; Zhang, C.; Shi, S.; Zhang, H. High temperature CO2 sensing and its cross-sensitivity towards H2 and CO gas using calcium doped ZnO thin film coated langasite SAW sensor. Sens. Actuators B Chem. 2019, 301, 126958. [Google Scholar] [CrossRef]
  18. Sayago, I.; Matatagui, D.; Fernández, M.J.; Fontecha, J.L.; Jurewicz, I.; Garriga, R.; Muñoz, E. Graphene oxide as sensitive layer in Love-wave surface acoustic wave sensors for the detection of chemical warfare agent simulants. Talanta 2016, 148, 393–400. [Google Scholar] [CrossRef]
  19. Matatagui, D.; Fernández, M.; Fontecha, J.; Sayago, I.; Gràcia, I.; Cane, C.; Horrillo, C.; Santos, J. Characterization of an array of Love-wave gas sensors developed using electrospinning technique to deposit nanofibers as sensitive layers. Talanta 2014, 120, 408–412. [Google Scholar] [CrossRef]
  20. Fragoso-Mora, J.; Matatagui, D.; Bahos, F.; Fontecha, J.; Fernandez, M.; Santos, J.; Sayago, I.; Gràcia, I.; Horrillo, M. Gas sensors based on elasticity changes of nanoparticle layers. Sens. Actuators B Chem. 2018, 268, 93–99. [Google Scholar] [CrossRef]
  21. Matatagui, D.; Fernández, M.; Fontecha, J.; Santos, J.; Gràcia, I.; Cane, C.; Horrillo, C. Propagation of acoustic waves in metal oxide nanoparticle layers with catalytic metals for selective gas detection. Sens. Actuators B Chem. 2015, 217, 65–71. [Google Scholar] [CrossRef]
  22. Raj, V.B.; Nimal, A.T.; Parmar, Y.; Sharma, M.U.; Gupta, V. Investigations on the origin of mass and elastic loading in the time varying distinct response of ZnO SAW ammonia sensor. Sens. Actuators B Chem. 2012, 166, 576–585. [Google Scholar]
  23. Raj, V.B.; Singh, H.; Nimal, A.; Tomar, M.; Sharma, M.; Gupta, V. Effect of metal oxide sensing layers on the distinct detection of ammonia using surface acoustic wave (SAW) sensors. Sens. Actuators B Chem. 2013, 187, 563–573. [Google Scholar] [CrossRef]
  24. Matatagui, D.; Bahos, F.A.; Gràcia, I.; Horrillo, M.D.C. Portable Low-Cost Electronic Nose Based on Surface Acoustic Wave Sensors for the Detection of BTX Vapors in Air. Sensors 2019, 19, 5406. [Google Scholar] [CrossRef] [Green Version]
  25. Park, S.Y.; Kim, Y.; Kim, T.; Eom, T.H.; Kim, S.Y.; Jang, H.W. Chemoresistive materials for electronic nose: Progress, perspectives, and challenges. InfoMat 2019, 1, 289–316. [Google Scholar] [CrossRef] [Green Version]
  26. El Kazzy, M.; Weerakkody, J.S.; Hurot, C.; Mathey, R.; Buhot, A.; Scaramozzino, N.; Hou, Y. An overview of artificial olfaction systems with a focus on surface plasmon resonance for the analysis of volatile organic compounds. Biosensors 2021, 11, 244. [Google Scholar] [CrossRef]
  27. Raj, V.B.; Singh, H.; Nimal, A.T.; Sharma, M.U.; Tomar, M.; Gupta, V. Distinct detection of liquor ammonia by ZnO/SAW sensor: Study of complete sensing mechanism. Sens. Actuators B Chem. 2017, 238, 83–90. [Google Scholar] [CrossRef]
  28. Tang, Y.; Xu, X.; Han, S.; Cai, C.; Du, H.; Zhu, H.; Zu, X.; Fu, Y. ZnO-Al2O3 nanocomposite as a sensitive layer for high performance surface acoustic wave H2S gas sensor with enhanced elastic loading effect. Sens. Actuators B Chem. 2020, 304, 127395. [Google Scholar] [CrossRef]
  29. Cruz, C.; Palomar, E.; Bravo, I.; Aleixandre, M. Behavioural patterns in aggregated demand response developments for commu-nities targeting renewables. Sustain. Cities Soci. 2021, 72, 103001. [Google Scholar] [CrossRef]
  30. Marczyński, P.; Szpakowski, A.; Tyszkiewicz, C.; Pustelny, T. Pattern Recognition Applied to Analysis of Gas Sensors’ Array Data. Acta Phys. Pol. A 2012, 122, 847–849. [Google Scholar] [CrossRef]
  31. Jha, S.K.; Yadava, R.D.S. Preprocessing of SAW Sensor Array Data and Pattern Recognition. IEEE Sens. J. 2009, 9, 1202–1208. [Google Scholar] [CrossRef]
  32. Singh, H.; Raj, V.B.; Kumar, J.; Durani, F.; Mishra, M.; Nimal, A.T.; Sharma, M.U. SAW mono sen-sor for identification of harmful vapors using PCA and ANN. Process. Saf. Environ. Prot. 2016, 102, 577–588. [Google Scholar] [CrossRef]
  33. García, M.; Fernández, M.; Fontecha, J.; Lozano, J.; Santos, J.; Aleixandre, M.; Sayago, I.; Gutiérrez, J.; Horrillo, M. Differentiation of red wines using an electronic nose based on surface acoustic wave devices. Talanta 2006, 68, 1162–1165. [Google Scholar] [CrossRef] [PubMed]
  34. Herrera-Chacon, A.; Campos, I.; González-Calabuig, A.; Torres, M.; del Valle, M. Coupling of Sensors and Machine Learning Algorithms in the Qualitative Analysis of Wine. Eng. Proc. 2021, 6, 50. [Google Scholar] [CrossRef]
  35. Ciosek, P.; Wróblewski, W. The analysis of sensor array data with various pattern recognition techniques. Sens. Actuators B Chem. 2006, 114, 85–93. [Google Scholar] [CrossRef]
  36. Zhang, W.; Tian, F.; Song, A.; Hu, Y. Research on electronic nose system based on continuous wide spectral gas sensing. Microchem. J. 2018, 140, 1–7. [Google Scholar] [CrossRef]
  37. Dang, L.; Tian, F.; Zhang, L.; Kadri, C.; Yin, X.; Peng, X.; Liu, S. A novel classifier ensemble for recognition of multiple indoor air contaminants by an electronic nose. Sens. Actuators A Phys. 2014, 207, 67–74. [Google Scholar] [CrossRef]
  38. Sisk, B.C.; Lewis, N.S. Estimation of chemical and physical characteristics of analyte vapors through analysis of the response data of arrays of polymer-carbon black composite vapor detectors. Sens. Actuators B Chem. 2003, 96, 268–282. [Google Scholar] [CrossRef]
  39. Feng, S.; Farha, F.; Li, Q.; Wan, Y.; Xu, Y.; Zhang, T.; Ning, H. Review on Smart Gas Sensing Technology. Sensors 2019, 19, 3760. [Google Scholar] [CrossRef] [Green Version]
  40. Al-Dayyeni, W.S.; Al-Yousif, S.; Taher, M.M.; Al-Faouri, A.W.; Tahir, N.M.; Jaber, M.M.; Ghabban, F.; Najm, I.A.; Alfadli, I.M.; Ameerbakhsh, O.Z.; et al. A Review on Electronic Nose: Coherent Taxonomy, Classification, Motivations, Challenges, Recommendations and Datasets. IEEE Access 2021, 9, 88535–88551. [Google Scholar] [CrossRef]
  41. Gutiérrez, J.; Horrillo, M.C. Advances in artificial olfaction. Sens. Appl. Talanta 2014, 124, 95–105. [Google Scholar] [CrossRef] [PubMed]
  42. de la O-Cuevas, E.; Alvarez-Venicio, V.; Badillo-Ramírez, I.; Islas, S.R.; Carreón-Castro, M.D.P.; Saniger, J.M. Graphenic substrates as modifiers of the emission and vibrational responses of interacting molecules: The case of BODIPY dyes. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 246, 119020. [Google Scholar] [CrossRef] [PubMed]
  43. Lu, G.; Ocola, L.E.; Chen, J. Reduced graphene oxide for room-temperature gas sen-sors. Nanotechnology 2009, 20, 445502. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, D.; Song, X.; Wang, Z.; Chen, H. Ultra-highly sensitive humidity sensing by polydopa-mine/graphene oxide nanostructure on quartz crystal microbalance. Appl. Surface Sci. 2021, 538, 147816. [Google Scholar] [CrossRef]
  45. Hallil, H.; Zhang, Q.; Flahaut, E.; Lachaud, J.-L.; Coquet, P.; Dejous, C.; Rebière, D. Guided SH-SAW sensor based on DWNTs sensitive material for VOCs and humidity detection. J. Integr. Circuits Syst. 2018, 13, 1–4. [Google Scholar] [CrossRef]
Figure 1. Experimental setup deployed for custom concentration measurements.
Figure 1. Experimental setup deployed for custom concentration measurements.
Sensors 22 01261 g001
Figure 2. SEM images of sensitive layers based on (a) mesoporous carbon, (b) reduced graphene oxide, (c) graphene oxide and (d) polydopamine/reduced graphene oxide.
Figure 2. SEM images of sensitive layers based on (a) mesoporous carbon, (b) reduced graphene oxide, (c) graphene oxide and (d) polydopamine/reduced graphene oxide.
Sensors 22 01261 g002
Figure 3. Raman spectra of different carbon-based sensitive layers.
Figure 3. Raman spectra of different carbon-based sensitive layers.
Sensors 22 01261 g003
Figure 4. Real time response and recovery of the SH-SAW sensor array based on MC, rGO, GO and PDA/rGO for a concentration of 0.4 ppm of NO2 (a), 4 ppm of CO (b), 40 ppm of benzene (C6H6) (c), 40 ppm of acetone (C3H6O) (d) and 40 ppm of NH3 (e).
Figure 4. Real time response and recovery of the SH-SAW sensor array based on MC, rGO, GO and PDA/rGO for a concentration of 0.4 ppm of NO2 (a), 4 ppm of CO (b), 40 ppm of benzene (C6H6) (c), 40 ppm of acetone (C3H6O) (d) and 40 ppm of NH3 (e).
Sensors 22 01261 g004
Figure 5. (a) Linearity of MC, GO, rGO and PDA/rGO obtained for different concentrations of NO2. (b) Real time response and recovery of an SH-SAW sensor with a sensitive layer for a concentration of 0.1 ppm of NO2.
Figure 5. (a) Linearity of MC, GO, rGO and PDA/rGO obtained for different concentrations of NO2. (b) Real time response and recovery of an SH-SAW sensor with a sensitive layer for a concentration of 0.1 ppm of NO2.
Sensors 22 01261 g005
Figure 6. Real time response of the SH-SAW sensor array based on MC, rGO, GO and PDA/rGO materials for different humidity values at 21 °C.
Figure 6. Real time response of the SH-SAW sensor array based on MC, rGO, GO and PDA/rGO materials for different humidity values at 21 °C.
Sensors 22 01261 g006
Figure 7. Radial representation of the sensor array’s responses to 0.2 ppm NO2, 2 ppm CO and 20 ppm of benzene (C6H6), acetone (C3H6O) and NH3.
Figure 7. Radial representation of the sensor array’s responses to 0.2 ppm NO2, 2 ppm CO and 20 ppm of benzene (C6H6), acetone (C3H6O) and NH3.
Sensors 22 01261 g007
Figure 8. (a) Principal component analysis (PCA) applied to data for discrimination of interferents (black), humidity (purple) and NO2 (orange) for different concentrations and sensors; (b) LDA classification results for the three classes (NO2, interferents and humidity) indicates a discrimination accuracy of 90%.
Figure 8. (a) Principal component analysis (PCA) applied to data for discrimination of interferents (black), humidity (purple) and NO2 (orange) for different concentrations and sensors; (b) LDA classification results for the three classes (NO2, interferents and humidity) indicates a discrimination accuracy of 90%.
Sensors 22 01261 g008
Figure 9. Confusion matrix obtained by KNN.
Figure 9. Confusion matrix obtained by KNN.
Sensors 22 01261 g009
Table 1. Sensitivity, F1 and precision scores for KNN classification (60% tested out of the total sample).
Table 1. Sensitivity, F1 and precision scores for KNN classification (60% tested out of the total sample).
PrecisionSensitivityF1-Score
Humidity10.940.97
Interferents0.6710.8
NO2111
Accuracy 0.95%
Table 2. Comparison of different SAW sensor parameters reported in the literature for NOx target gas.
Table 2. Comparison of different SAW sensor parameters reported in the literature for NOx target gas.
Sensitive
Layer
Target GasOperating FrequencySensitivityDetected ppmResponse TimeDetection MechanicsReference
SnO2NO2433.9 Mhz-20 ppm2 sElastic loading[13]
ZnONO299.5 Mhz2.9 Hz/ppb400 ppb–16 ppm-Mass loading[14]
PPy/WO/rGONO98 Hhz12 Hz/ppb5–110 ppb<2minMass loading[15]
PZTNO299.4 Mhz9.6 Hz/ppm80–250 ppm-Mass loading[16]
MC
rGO
GO
PDA/rGO
NO2160 Mhz10 Hz/ppb
7.5 Hz/ppb
75 Hz/ppb
30 Hz/ppb
0.1–1 ppm<1minMass loading
Elastic loading
Elastic loading
Mass loading
This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cruz, C.; Matatagui, D.; Ramírez, C.; Badillo-Ramirez, I.; de la O-Cuevas, E.; Saniger, J.M.; Horrillo, M.C. Carbon SH-SAW-Based Electronic Nose to Discriminate and Classify Sub-ppm NO2. Sensors 2022, 22, 1261. https://doi.org/10.3390/s22031261

AMA Style

Cruz C, Matatagui D, Ramírez C, Badillo-Ramirez I, de la O-Cuevas E, Saniger JM, Horrillo MC. Carbon SH-SAW-Based Electronic Nose to Discriminate and Classify Sub-ppm NO2. Sensors. 2022; 22(3):1261. https://doi.org/10.3390/s22031261

Chicago/Turabian Style

Cruz, Carlos, Daniel Matatagui, Cristina Ramírez, Isidro Badillo-Ramirez, Emmanuel de la O-Cuevas, José M. Saniger, and Mari Carmen Horrillo. 2022. "Carbon SH-SAW-Based Electronic Nose to Discriminate and Classify Sub-ppm NO2" Sensors 22, no. 3: 1261. https://doi.org/10.3390/s22031261

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