Next Article in Journal
Reflections of High-Frequency Pulsed Ultrasound by Underwater Acoustic Metasurfaces Composed of Subwavelength Phase-Gradient Slits
Previous Article in Journal
Hetero-Deformation Induced Hardening in a CoCrFeNiMn High-Entropy Alloy
Previous Article in Special Issue
Preparation of Novel Electrospun Organic–Inorganic Hybrid Nanofibers Based on Polycaprolactone, Glycine Betaine, and Calcium Carbonate (P/G/CaCO3-EPnF)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ternary Hybrid Materials for Highly Sensitive Acetone Sensing at Room Temperature

by
Nurul Athirah Abu Hussein
1,2,*,
Yew Hoong Wong
3,
Zainal Arif Burhanudin
1,2 and
Huzein Fahmi Hawari
1,2,*
1
Department of Electrical and Electronic Engineering, Universiti Teknologi Petronas, Seri Iskandar 32610, Malaysia
2
Centre of Innovative Nanostructure and Nanodevices (COINN), Universiti Teknologi Petronas, Seri Iskandar 32610, Malaysia
3
Centre of Advanced Materials, Department of Mechanical Engineering, Faculty of Engineering, Universiti Malaya, Kuala Lumpur 50603, Malaysia
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(5), 845; https://doi.org/10.3390/cryst13050845
Submission received: 18 April 2023 / Revised: 11 May 2023 / Accepted: 16 May 2023 / Published: 20 May 2023

Abstract

:
The performance of a conventional metal oxide sensor (MOX) is highly dependent on its high operating temperature. Many researchers have tried to solve the problem by exploring hybrid materials. On the other hand, ternary hybrid materials have emerged as a promising class of materials with unique properties and potential applications in various fields, be it environmental or medical, such as in breath analyzers for prediabetes analysis. This article focuses on the synthesis method, characterization, and application of ternary hybrid materials for room-temperature sensors, as well as recent advances and future developments in the field. The materials consist of three different components, metal oxide (Fe3O4), polymer (polyaniline) and carbon-based materials (reduced graphene oxide), which were synthesized using in-situ methods. Five samples were prepared in different ratios. The properties of these materials were characterized using techniques such as X-ray diffraction (XRD), Raman, scanning electron microscope (SEM) and transmission electron microscopy (TEM). The XRD and Raman analyses showed the existence of all the individual constituents in the hybrid sample. SEM and TEM also showed a strong interaction between the constituent materials as a hybrid nanocomposite. The response and recovery time were studied against 1, 10 and 100 ppm acetone. The results show that the sample with 10 wt%Fe3O4-PANI-RGO (S2_10) has a reaction and recovery time < 32 s against the above ppm and has the highest sensing response at room temperature.

1. Introduction

1.1. Background

Breath is the largest potential source of volatile organic compounds (VOCs) and is also an important indicator of invasive or non-invasive medical diagnosis and health surveillance [1,2]. More than 870 types of volatile organic compounds (VOCs) can be detected in human breath. These VOCs provide important information about metabolic abnormalities or malfunctions in the human body [2], such as cancer [3], cystic fibrosis [4] and diabetes [5,6]. Diabetes is a common health problem in which patients have high blood glucose levels because the body is unable to produce sufficient amounts of insulin [5,7]. Despite enormous efforts to prolong the lives of diabetics, diabetes remains the fifth leading cause of death worldwide with 1.6 million deaths [5]. For early detection of diabetics, acetone gas concentration can be used as a marker to distinguish diabetics from healthy individuals. The acetone concentration in exhaled air is reported to be usually between 0.2 and 0.8 ppm in healthy people, while the threshold for diabetics is 1.8 ppm [8].
In general, gas sensors based on metal oxides are the most promising for monitoring harmful volatile organic compounds (VOCs). They are in great demand in many fields, such as climate, safety, security, and process monitoring as well as for medical devices such as respiratory gas analyzers for the pre-diagnosis of diabetes. However, due to their high operating temperature of up to over 150 °C, the performance of the sensors degrades over time [7,9]. The discovery of hybrid materials generated a lot of interest in various fields. These are basically materials that consist of two or more different components and can combine the advantageous properties of each component to create new or improved functionalities [10]. Over the years, the applications of hybrid materials have expanded to many fields, such as medicine [11], batteries [12], supercapacitors [13], energy storage [14,15] and sensing [16,17,18,19].
In the field of sensor technology, a good sensor should be inexpensive, reliable, stable and easy to manufacture. Hybrid materials can improve the sensitivity, selectivity, stability, and reproducibility of a sensor [20,21]. Hybridization between materials is considered a promising strategy to overcome the drawbacks of organic and inorganic materials for gas sensing. It offers new or enhanced functionalities through a synergic interaction between the two phases [22,23,24]. So far, different materials, such as inorganic semiconductors, metal oxides and conducting polymers, have been used to fabricate sensors with a small size, low power consumption, high sensitivity, and high reliability [16,22,25,26,27].
By taking advantage of the hybridizing materials, the performance of the sensor can be improved. Some of the materials that can be used for hybridization with metal oxides are derivatives of graphene that is reduced to graphene oxide (RGO) and polymers such as polyaniline. The functional groups in RGO have been found to interact strongly with various gases, making RGO a promising platform for room-temperature chemical gas sensing applications [28]. On the other hand, the combination of polymers with iron oxide nanoparticles has attracted much attention in conductive polymer applications due to their structures as well as their chemical and physical properties. It has been proven that the hybridization of iron oxide on the network of the polymer material results in better conductivity and magnetization by generating an electromagnetic field (EM) [17,29,30]. Many researches have been carried out and have shown the advantages of inorganic compounds in the PANI matrix, especially magnetite (Fe3O4). Fe3O4 has good ferromagnetic properties and is resistant to corrosions by acidic solutions. Since Fe3O4 is a hydrophobic material, chemical reactions with PANI increase its dispersions in solution. It also shows good electrical behavior when an external electric field is applied to the material [29].

1.2. Sensing Mechanism

The metal oxide gas sensor works on the principle of chemoresistance, where the gas molecules interacting with the metal oxides act either as donors or acceptors of charge carriers and change the resistance of the metal oxide. The increase or decrease in the resistance of the metal oxide thin film depends on the type of majority carriers in the semiconductor layer and the type of gas molecules (oxidizing or reducing) in the surrounding atmosphere. For n-type materials, oxidizing gases (acceptor) increase the resistance of the thin film, while reducing gases (donor) decrease the resistance, and vice versa for p-type materials [7]. When the sensor is exposed to the atmosphere, oxygen is adsorbed on the surface of the metal oxide and oxidation of the oxygen molecules takes place, with the oxygen atom accepting excited electrons from the conduction band of the metal oxide and becoming oxygen ions ( O 2 O , O2−), and the reaction equations are (1)–(4) [31]. The resistance of the sensor remains the same if the ambient factor remains unchanged.
O2(gas) → O2(adsorbed)
O 2 ( adsorbed ) + e O 2
O2(adsorbed) + e→ 2O
O2(adsorbed) + eO2−
In the present work, we propose the preparation of a novel active material from the ternary hybrid material Fe3O4-PANI-RGO for the detection of acetone at room temperature. To the best of our knowledge, there are papers mentioning the hybridization of the proposed materials but none addressing the application for pre-diabetes diagnosis.

2. Materials and Methods

2.1. Materials

Iron (II) Chloride Tetrahydrate (FeCl2.4H2O), Iron (III) Chloride Hexahydrate (FeCl3.6H2O), Iron (III) Nitrate (FeNO3)3.9H2O, Ammonium hydroxide (25%), Nitric acid (65%), Hydrochloric acid (37%), Aniline (99%), Ammonium peroxodisulfate (98%) Silver (Ag) paste (resistivity 30–35 µΩ-cm), Ascorbic Acid and Graphene oxide (GO) paste (95 wt% purity) were all purchased from Merck. De-ionized (DI) water was used in all preparations. Interdigitated electrode (IDE) with a size of 5 mm × 5 mm was purchased from Nova Scientific (Sturbridge, MA, USA).

2.2. Synthesis of Ternary Hybrid Nanocomposite Fe3O4-PANI-RGO

The ternary hybrid nanocomposite was prepared by an in-situ method, as shown in Figure 1, where the as-synthesized Fe3O4 and PANI were first sonicated in a bath for 30 min. Then both solvents were added and mixed with the GO paste for another 1 h to ensure homogeneity. Then, 0.1 g ascorbic acid was added to the mixture and heated to 90 °C for 2 h. Finally, the product was centrifuged and washed with ethanol and water. The dry product was obtained by drying the mixture in an oven. The synthesis process is illustrated in Figure 1. Five different samples were synthesized with different weight ratios of the materials. The name and description of the samples are listed in Table 1.

2.3. Thin-Film Development and Gas-Sensing Setup

Figure 2 shows the diagram of the development of the thin film on IDE and the construction of the gas sensor. A volume 1.0 μL was developed on the IDE using the drop casting method. The thin film was then dried at room temperature. For the measurement setup, the IDE chamber was connected to the outlet tube of the air and VOC and the multimeter.

2.4. Characterization

Microstructural analysis of the synthesized active materials was studied using an x-ray diffractometer (XRD) (Panalytical, model X’Pert3 Powder, Malvern, UK), structural and morphological analysis was performed using Raman spectroscopy (Horiba, model HR800, 514 nm laser excitation, Kyoto, Japan) for the spectral ranges of 200 to 4000 cm−1, scanning electron microscope (SEM) (Tescan Vega, Brno, Czech Republic) and transmission electron microscope (TEM)(Hitachi, model HT7830, Tokyo, Japan).
A specially designed vapor sensor setup was used to investigate the performance of the active materials at room temperature. The IDE was placed in a chamber and connected to the multimeter. First, air was purged onto the Fe3O4-PANI-RGO-coated IDE for 60 s, then, VOC (acetone) was purged onto the IDE. The analyte from the vapor was adsorbed on the surface of the active materials, resulting in changes in the impedance and capacitance of the IDE due to the polarization and realignment of molecules in the dielectric materials. The process was continued with the purging of air to desorb the analyte from the surface of the active materials. The response and recovery time were monitored in real time on the connected computer.

3. Results and Discussion

3.1. Microstructural Analysis

The microstructural analysis of the materials was investigated by XRD. The pattern consists of the as-synthesized Fe3O4, PANI, GO and S2_10 and is shown in Figure 3. For PANI, the peaks are found to be at 2θ: 15.1, 20.7, 25.5 and 27.3° with corresponding d-spacings of 5.9, 4.3, 3.5, and 3.3 Ǻ, which are very similar to those in the study conducted by (Singu et al., 2011) [32]. For graphene oxide (GO), the main peak is at about ~12.5°, indicating the presence of amorphous GO. The disappearance of this peak in sample S2_10 indicates the complete reduction GO to RGO [33]. The integration of the plane with Fe3O4 and PANI reveales the low crystallinity of RGO and shows a large interfacial contact between all the constituent materials [34,35]. For Fe3O4, all peaks showed consistent agreement with ICDD Card No. 98-015-8746, where the strong and sharp diffraction peaks were seen at 2θ: 30.3, 35.7, 43.2, 53.8, 57.3 and 62.9°, corresponding to the d200, d311, d400, d422, d511 and d440 planes respectively. The calculated cubic lattice parameter was 8.33 Ǻ, which is in agreement with the findings reported in the literature review [35]. The crystallite size of the samples was calculated using the Debye–Scherrer equation:
D = k λ β c o s θ
where D is the particle size (nm), k is a constant ( k = 0.94), λ is the x-ray wavelength, β is the full width at half maximum (FWHM) of the diffraction peak, and θ is the angle of diffraction. The crystallite size calculated from (Equation (5)) for the samples Fe3O4, PANI and S2_10 were 9.68 ± 1.9, 4.17 ± 3.2 and 9.34 ± 1.4 nm, respectively.

3.2. Raman Analysis

Raman spectroscopic analysis was performed to verify the presence of PANI and RGO. Figure 4 shows two observation peaks at 1345 cm−1 and 1594 cm−1 for the ternary hybrid material sample, indicating the D and G bands of GO. While the G band correlates to the sp2-hybridized graphitic carbon atoms in RGO, the D band is related to structural defects and disorder of RGO [36,37]. When comparing the single constituents of RGO with sample S1_10, it was found that the ID/IG decreased from 0.975 to 0.969, which is related to the reduction of the defect following the PANI coating [38]. Two other small peaks from sample S1_10 can be seen at 1157 and 1484 cm−1, which represent the C-H vibration in the quinoid and semiquinone structures of PANI and show the establishment of a PANI chain on GO for the ternary hybrid nanocomposite material [39].

3.3. Morphological Analysis

The structure and morphology of the as-synthesized individual components of Fe3O4, PANI and RGO as well as the ternary hybrid material were investigated by SEM and TEM analysis. Figure 5a shows the SEM image of RGO, where a clear, wrinkled surface at a magnification of 2 µm can be seen. Figure 5b,c show the image of PANI with a high porosity and a rough crystalline image of Fe3O4, respectively. The agglomeration of Fe3O4 is due to its highly magnetic properties [38]. The porosity of PANI could be useful for the absorption of analyte during the sensing process, as it will provide a more active site [40,41]. Figure 5d–h show the ternary hybrid nanocomposite samples of S1_5, S2_10, S3_20, S4_30 and S5_40, showing a similar morphology to the single constituent materials, with hole-like porosity evident for all concentrations of the hybrid materials. A dense PANI and Fe3O4 nanomaterial is seen coated on the sheet-like RGO. This proves that PANI and Fe3O4 are evenly distributed on the surface of RGO.
TEM analyses of the single material Fe3O4, RGO, PANI, the ternary hybrid sample S2_10 and the SAED diffraction are shown in Figure 6a–e, respectively. The single-material Fe3O4 shows a spherical shape with different sizes and a low aggregation with an average particle size of 6.1 ± 2.1 nm, which complements the crystallite size from XRD, as shown in Figure 7. RGO shows a common sheet-like structure with a smooth surface, which is consistent with the image from SEM. This result is similar to the study reported by Gupta et al. (2022) [42]. PANI shows a rod-like structure due to the electrostatic interaction [41]. The size of the rod-like structure is about 15–23 nm in diameter. Figure 6d shows sample S2_10, which has a similar structure to Fe3O4, with the spherical shape of Fe3O4 incorporated with PANI uniformly located on the surface of the RGO sheet. The calculated particle size was 6.3 ± 1.8 nm. These findings support the synthesis of the ternary hybrid nanocomposite with an excellent dispersion. Figure 6e shows the selected area diffraction pattern (SAED) of sample S2_10, in which a well-defined diffraction ring can be seen in agreement with planes (200), (220), (311), (400), (422), (511) and (440), complementing the XRD result.

3.4. Sensing Analysis

Figure 8 shows the schematic diagram of the sensor mechanism. When the metal oxide sensor is placed in an environment with the presence of target gas, the target gas will react with the oxygen species on the surface of the active material (Equations (1)–(4)). The interaction between the target gas and the oxygen species will create a resistance change in the sensor. For acetone, the reaction is simplified as follows [43]:
CH 3 COCH 3   ( gas ) +   O 2   adsorbed   CH 3 C + O + CH 3 O + e
CH 3 C + O   CH 3 + + CO   ( gas )
CO   ( gas ) + O 2   adsorbed   CO 2 + e
Figure 9a–c show the response and recovery time for all samples when expose to 1, 10, and 100 ppm acetone, respectively, while Figure 10a–c show the real-time response of samples S2_10 when expose to 1, 10, and 100 ppm acetone, respectively. The response time is defined as the time from the first contact of the sensor with acetone vapor, in which the sensor resistance reaches 90% of the saturated value of the resistance to acetone vapor. The recovery time is the time from which the acetone resistance returns to 10% of the baseline [44]. To imitate the most common environment for the gas sensor, air was used as the carrier gas.
From the response in Figure 10, we can see that the resistance of the samples decreases/increases upon exposure to acetone and gradually stabilizes. The response shows a concentration-dependent behavior depending on the charge carrier [45]. When p-type semiconductors meet n-type semiconductors, the holes in the p-type semiconductors diffuse into the n-type semiconductors and neutralize the free electrons, resulting in the generation of positive charges in the n-type semiconductors. This explains the emergence of a synergistic effect between the organic p-type material (PANI) and the n-type material (RGO-Fe3O4) [45,46]. From our previous work, the response and recovery time of Fe3O4 to 300 ppm acetone at 300 °C was 2.19/3.6 min, respectively [47]. Based on the current work, the response and recovery time are of which is shown in Figure 9, sample S1_5 had the fastest response and recovery time to 1 ppm acetone, which was 13.6 s/13.5 s, respectively. When exposed to 10 and 100 ppm acetone, sample S2_10 had a better response and recovery time compared to the other samples, with measurements of 20.8 s/17.7 s and 31.3 s/22.4 s, respectively. The performance of the sensor improves at a lower ppm of acetone and at the lower operating temperature, i.e., room temperature. This could be due to an increase in the size of the analyte site leading to an increase in the detected signal as well as the sensitivity of the sensor. One of the reasons for the increasing sensitivity is the pore that can be seen in the SEM image [48]. It was reported by Abokifa et al. (2019) that the sensing at room temperature was attributed to some factors, such as the morphology, small grain size and high surface-to-volume ratio of the nanostructured sensors, as the above-mentioned mechanism (Equation (3)) only takes place under high-operating-temperature conditions (<200 °C) [31]. In this case, 10 wt% was the optimal ratio of the ternary hybrid, as increasing the ratio above this value resulted in the response and recovery time of the sensor also increasing. However, when the concentration was further increased to 40 wt%, the response/recovery time decreased. This finding proves that adjusting the active material of the MOX sensor can affect the performance of the sensor, such as the sensitivity and the response and recovery times.
The sensing response of the sensor (S) was calculated using the following equation:
S   % = R g R a R a × 100
where Rg is the electrical resistance in the presence of gas, and Ra is the electrical resistance in the presence of air. The sensor response shown in Figure 11 for sample S2_10 was observed for 42 days. It shows that the response does not vary significantly. This is due to some factors, such as the ability of the analyte molecules to be adsorbed on the surface of the active materials due to the pore volume [48]. This can be explained by the SEM image in Figure 5e, where the pore volume of sample S2_10 can be obviously seen on the surface of the sample. Therefore, the improvement of the pore volume has a great influence on the sensitivity of the sample [49].
Figure 12 shows the sensory response of all samples when exposed to 1000 ppm formaldehyde, ethanol and acetone. It can be seen that sample S1_5 reacts more strongly to formaldehyde, while samples S2_10, S3_20 and S4_30 react three times more strongly to acetone compared to formaldehyde and ethanol. From this, we can conclude that sample S2_10 has a high response to acetone and is suitable for breath analyzers for pre-diabetic diagnosis.

4. Conclusions

In general, the ternary hybrid nanocomposite materials were synthesized and investigated for their sensing capability. The reaction and recovery time of the ternary hybrid material was studied for 1, 10 and 100 ppm acetone. The fastest response and recovery time for the lowest ppm of acetone was confirmed for sample S1_5. However, as the concentration increased, the sensor response also increased. In this case, sample S2_10 reacted the most to acetone vapor compared to the other samples. This could be due to several reasons, such as the depletion layer, the available active sites and the thickness of the samples cast on IDE. This shows that concentration is dependent on the performance of the sensor. However, further studies need to be carried out to investigate the chemical and physical properties of the active materials, so that modifications can be made to further reduce the reaction and recovery times and improve the sensitivity of the sensor.

Author Contributions

Conceptualization, N.A.A.H. and H.F.H.; Data curation, N.A.A.H.; Formal analysis, N.A.A.H.; Funding acquisition, H.F.H.; Investigation, N.A.A.H.; Methodology, N.A.A.H.; Project administration, H.F.H.; Resources, H.F.H.; Software, N.A.A.H.; Supervision, Y.H.W., Z.A.B. and H.F.H.; Validation, N.A.A.H.; Visualization, H.F.H.; Writing—original draft, N.A.A.H.; Writing—review & editing, N.A.A.H. and H.F.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Higher Education (MOHE) through the Fundamental Research Grant Scheme (Ref No. FRGS/1/2019/TK04/UTP/02/7).

Data Availability Statement

The data is available upon request from the corresponding author (N.A.A.H).

Acknowledgments

The authors would like to thank Centralized Analytical Laboratory (CAL), UTP for their contributions in this publication.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, D.; Hu, Y.; Chen, Y.; Zeng, B. BreathTrack: Tracking Indoor Human Breath Status via Commodity WiFi. IEEE Internet Things J. 2019, 6, 3899–3911. [Google Scholar] [CrossRef]
  2. Zhou, X.; Xue, Z.; Chen, X.; Huang, C.; Bai, W.; Lu, Z.; Wang, T. Nanomaterial-based gas sensors used for breath diagnosis. J. Mater. Chem. B 2020, 8, 3231–3248. [Google Scholar] [CrossRef]
  3. van Keulen, K.E.; Jansen, M.E.; Schrauwen, R.W.M.; Kolkman, J.J.; Siersema, P.D. Volatile organic compounds in breath can serve as a non-invasive diagnostic biomarker for the detection of advanced adenomas and colorectal cancer. Aliment Pharm. 2020, 51, 334–346. [Google Scholar] [CrossRef]
  4. Woollam, M.; Siegel, A.P.; Grocki, P.; Saunders, J.L.; Sanders, D.B.; Agarwal, M.; Davis, M.D. Preliminary method for profiling volatile organic compounds in breath that correlate with pulmonary function and other clinical traits of subjects diagnosed with cystic fibrosis: A pilot study. J. Breath Res. 2022, 16, 027103. [Google Scholar] [CrossRef]
  5. Khan, R.M.M.; Chua, Z.J.Y.; Tan, J.C.; Yang, Y.; Liao, Z.; Zhao, Y. From Pre-Diabetes to Diabetes: Diagnosis, Treatments and Translational Research. Medicina 2019, 55, 546. [Google Scholar] [CrossRef]
  6. Dixit, K.; Fardindoost, S.; Ravishankara, A.; Tasnim, N.; Hoorfar, M. Exhaled Breath Analysis for Diabetes Diagnosis and Monitoring: Relevance, Challenges and Possibilities. Biosensors 2021, 11, 476. [Google Scholar] [CrossRef]
  7. Zhang, H.; Xiao, J.; Wang, Y.; Zhang, L.; Zhao, G.; Yang, H.; Wang, H. A portable acetone detector based on SmFeO3 can pre-diagnose diabetes through breath analysis. J. Alloys Compd. 2022, 922, 166160. [Google Scholar] [CrossRef]
  8. Baharuddin, A.A.; Ang, B.C.; Haseeb, A.S.M.A.; Wong, Y.C.; Wong, Y.H. Advances in chemiresistive sensors for acetone gas detection. Mater. Sci. Semicon Proc. 2019, 103, 104616. [Google Scholar] [CrossRef]
  9. Sharma, B.; Sharma, A.; Kim, J.-S. Recent advances on H2 sensor technologies based on MOX and FET devices: A review. Sens. Actuators B Chem. 2018, 262, 758–770. [Google Scholar] [CrossRef]
  10. Adnan, M.M.; Dalod, A.R.M.; Balci, M.H.; Glaum, J.; Einarsrud, M.A. In Situ Synthesis of Hybrid Inorganic(-)Polymer Nanocomposites. Polymers 2018, 10, 1129. [Google Scholar] [CrossRef] [PubMed]
  11. Broza, Y.Y.; Har-Shai, L.; Jeries, R.; Cancilla, J.C.; Glass-Marmor, L.; Lejbkowicz, I.; Torrecilla, J.S.; Yao, X.; Feng, X.; Narita, A.; et al. Exhaled Breath Markers for Nonimaging and Noninvasive Measures for Detection of Multiple Sclerosis. ACS Chem. Neurosci 2017, 8, 2402–2413. [Google Scholar] [CrossRef]
  12. Wu, S.; Fu, G.; Lv, W.; Wei, J.; Chen, W.; Yi, H.; Gu, M.; Bai, X.; Zhu, L.; Tan, C.; et al. A Single-Step Hydrothermal Route to 3D Hierarchical Cu2O/CuO/rGO Nanosheets as High-Performance Anode of Lithium-Ion Batteries. Small 2018, 14, 1702667. [Google Scholar] [CrossRef]
  13. Lu, X.F.; Chen, X.Y.; Zhou, W.; Tong, Y.X.; Li, G.R. alpha-Fe2O3@PANI Core-Shell Nanowire Arrays as Negative Electrodes for Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 14843–14850. [Google Scholar] [CrossRef] [PubMed]
  14. Notarianni, M.; Liu, J.; Vernon, K.; Motta, N. Synthesis and applications of carbon nanomaterials for energy generation and storage. Beilstein J Nanotechnol 2016, 7, 149–196. [Google Scholar] [CrossRef]
  15. Tabrizi, A.G.; Arsalani, N.; Naghshbandi, Z.; Ghadimi, L.S.; Mohammadi, A. Growth of polyaniline on rGO-Co3S4 nanocomposite for high-performance supercapacitor energy storage. Int. J. Hydrog. Energy 2018, 43, 12200–12210. [Google Scholar] [CrossRef]
  16. Ahmadi Tabr, F.; Salehiravesh, F.; Adelnia, H.; Gavgani, J.N.; Mahyari, M. High sensitivity ammonia detection using metal nanoparticles decorated on graphene macroporous frameworks/polyaniline hybrid. Talanta 2019, 197, 457–464. [Google Scholar] [CrossRef]
  17. Bandgar, D.K.; Navale, S.T.; Mane, A.T.; Gupta, S.K.; Aswal, D.K.; Patil, V.B. Ammonia sensing properties of polyaniline/α-Fe2O3 hybrid nanocomposites. Synth. Met. 2015, 204, 1–9. [Google Scholar] [CrossRef]
  18. Zhao, T.; Ren, Y.; Jia, G.Y.; Zhao, Y.; Fan, Y.; Yang, J.; Zhang, X.; Jiang, W.; Wang, L.; Wei, L. Facile synthesis of mesoporous WO3@graphene aerogel nanocomposites for low-temperature acetone sensing. Chin. Chem. Lett. 2019, 30, 2032–2038. [Google Scholar] [CrossRef]
  19. Abu Hussein, N.A.; Hawari, H.F.; Wong, Y.H. Synthesis of Iron Oxide/Polyaniline/Reduced Graphene Oxide Nanocomposite Materials as Active Sensing Material. In Proceedings of the 2020 8th International Conference on Intelligent and Advanced Systems (ICIAS), Kuching, Malaysia, 13–15 July 2021. [Google Scholar] [CrossRef]
  20. Dong, Z.M.; Xia, Q.; Ren, H.; Shang, X.; Lu, X.; Joo, S.W.; Huang, J. Preparation of hollow SnO2/ZnO cubes for the high-performance detection of VOCs. Ceram. Int. 2023, 49, 4650–4658. [Google Scholar] [CrossRef]
  21. Foronda, J.R.F.; Aryaswara, L.G.; Santos, G.N.C.; Raghu, S.N.V.; Muflikhun, M.A. Broad-class volatile organic compounds (VOCs) detection via polyaniline/zinc oxide (PANI/ZnO) composite materials as gas sensor application. Heliyon 2023, 9, e13544. [Google Scholar] [CrossRef] [PubMed]
  22. Zhao, H.; Zhang, T.; Qi, R.; Dai, J.; Liu, S.; Fei, T.; Lu, G. Development of solution processible organic-inorganic hybrid materials with core-shell framework for humidity monitoring. Sens. Actuators B Chem. 2018, 255, 2878–2885. [Google Scholar] [CrossRef]
  23. Liu, Z.; Yang, X.; Huo, L.; Tian, X.; Qi, T.; Yang, F.; Wang, X.; Yu, K.; Ma, F.; Sun, J. P-CuPcTS/n-SnO2 organic-inorganic hybrid film for ppb-level NO2 gas sensing at low operating temperature. Sens. Actuators B Chem. 2017, 248, 324–331. [Google Scholar] [CrossRef]
  24. Kim, M.Y.; Hwang, J.Y.; Mirzaei, A.; Choi, S.-W.; Kim, S.-I.; Kim, H.-S.; Kim, S.-J.; Roh, J.W.; Choi, M.S.; Lee, K.H.; et al. NO2 Gas Sensing Properties of Ag-Functionalized Porous ZnO Sheets. Adsorpt. Sci. Technol. 2023, 2023, 9021169. [Google Scholar] [CrossRef]
  25. Andre, R.S.; Mercante, L.A.; Facure, M.H.M.; Mattoso, L.H.C.; Correa, D.S. Enhanced and selective ammonia detection using In2O3/reduced graphene oxide hybrid nanofibers. Appl. Surf. Sci. 2019, 473, 133–140. [Google Scholar] [CrossRef]
  26. Zhang, B.; Cheng, M.; Liu, G.; Gao, Y.; Zhao, L.; Li, S.; Wang, Y.; Liu, F.; Liang, X.; Zhang, T.; et al. Room temperature NO2 gas sensor based on porous Co3O4 slices/reduced graphene oxide hybrid. Sens. Actuators B Chem. 2018, 263, 387–399. [Google Scholar] [CrossRef]
  27. Lee, Z.Y.; Hawari, H.F.B.; Djaswadi, G.W.; Kamarudin, K. A Highly Sensitive Room Temperature CO2 Gas Sensor Based on SnO2-rGO Hybrid Composite. Materials 2020, 14, 522. [Google Scholar] [CrossRef]
  28. Drmosh, Q.A.; Yamani, Z.H.; Hendi, A.H.; Gondal, M.A.; Moqbel, R.A.; Saleh, T.A.; Khan, M.Y. A novel approach to fabricating a ternary rGO/ZnO/Pt system for high-performance hydrogen sensor at low operating temperatures. Appl. Surf. Sci. 2019, 464, 616–626. [Google Scholar] [CrossRef]
  29. Hussain, N.H.I.; Mustafa, M.K.; Asman, S. Synthesis of PANI/Iron (II, III) Oxide Hybrid Nanocomposites Using SolGel Method. J. Sci. Technol. 2018, 10, 1–4. [Google Scholar] [CrossRef]
  30. Li, Y.; Zhao, H.; Ban, H.; Yang, M. Composites of Fe2O3 nanosheets with polyaniline: Preparation, gas sensing properties and sensing mechanism. Sens. Actuators B Chem. 2017, 245, 34–43. [Google Scholar] [CrossRef]
  31. Abokifa, A.A.; Haddad, K.; Fortner, J.; Lo, C.S.; Biswas, P. Sensing mechanism of ethanol and acetone at room temperature by SnO2 nano-columns synthesized by aerosol routes: Theoretical calculations compared to experimental results. J. Mater. Chem. A 2018, 6, 2053–2066. [Google Scholar] [CrossRef]
  32. Sydulu Singu, B.; Srinivasan, P.; Pabba, S. Benzoyl Peroxide Oxidation Route to Nano Form Polyaniline Salt Containing Dual Dopants for Pseudocapacitor. J. Electrochem. Soc. 2011, 159, A6–A13. [Google Scholar] [CrossRef]
  33. Amir Faiz, M.S.; Che Azurahanim, C.A.; Raba'ah, S.A.; Ruzniza, M.Z. Low cost and green approach in the reduction of graphene oxide (GO) using palm oil leaves extract for potential in industrial applications. Results Phys. 2020, 16, 102954. [Google Scholar] [CrossRef]
  34. Manna, R.; Srivastava, S.K. Reduced Graphene Oxide/Fe3O4/Polyaniline Ternary Composites as a Superior Microwave Absorber in the Shielding of Electromagnetic Pollution. ACS Omega 2021, 6, 9164–9175. [Google Scholar] [CrossRef] [PubMed]
  35. Mathew, J.; Sathishkumar, M.; Kothurkar, N.K.; Senthilkumar, R.; Sabarish Narayanan, B. Polyaniline/Fe3O4-RGO Nanocomposites for Microwave Absorption. IOP Conf. Ser. Mater. Sci. Eng. 2018, 310, 012138. [Google Scholar] [CrossRef]
  36. Kumar, R.; Youssry, S.M.; Ya, K.Z.; Tan, W.K.; Kawamura, G.; Matsuda, A. Microwave-assisted synthesis of Mn3O4-Fe2O3/Fe3O4@rGO ternary hybrids and electrochemical performance for supercapacitor electrode. Diam. Relat. Mater. 2020, 101, 107622. [Google Scholar] [CrossRef]
  37. Patil, V.S.; Thoravat, S.S.; Kundale, S.S.; Dongale, T.D.; Patil, P.S.; Jadhav, S.A. Synthesis and testing of polyaniline grafted functional magnetite (Fe3O4) nanoparticles and rGO based nanocomposites for supercapacitor application. Chem. Phys. Lett. 2023, 814, 140334. [Google Scholar] [CrossRef]
  38. Cai, H.; Feng, C.; Xiao, H.; Cheng, B. Synthesis of Fe3O4/rGO@PANI with three-dimensional flower-like nanostructure and microwave absorption properties. J. Alloys Compd. 2022, 893, 162227. [Google Scholar] [CrossRef]
  39. Mondal, S.; Rana, U.; Malik, S. Reduced Graphene Oxide/Fe3O4/Polyaniline Nanostructures as Electrode Materials for an All-Solid-State Hybrid Supercapacitor. J. Phys. Chem. C 2017, 121, 7573–7583. [Google Scholar] [CrossRef]
  40. Ahirrao, D.J.; Pal, A.J.; Singh, V.; Jha, N. Nanostructured porous polyaniline (PANI) coated carbon cloth (CC) as electrodes for flexible supercapacitor device. J. Mater. Sci. Technol. 2021, 88, 168–182. [Google Scholar] [CrossRef]
  41. Ma, Y.; Zhou, Y.; Sun, Y.; Chen, H.; Xiong, Z.; Li, X.; Shen, L.; Liu, Y. Tunable magnetic properties of Fe3O4/rGO/PANI nanocomposites for enhancing microwave absorption performance. J. Alloys Compd. 2019, 796, 120–130. [Google Scholar] [CrossRef]
  42. Gupta, M.; Hawari, H.F.; Kumar, P.; Burhanudin, Z.A. Copper Oxide/Functionalized Graphene Hybrid Nanostructures for Room Temperature Gas Sensing Applications. Crystals 2022, 12, 264. [Google Scholar] [CrossRef]
  43. Zhang, D.Z.; Wu, Z.L.; Zong, X.Q. Metal-organic frameworks-derived zinc oxide nanopolyhedra_S, N_ graphene quantum dots_polyaniline ternary nanohybrid for high-performance acetone sensing. Sens. Actuators B Chem. 2019, 288, 232–242. [Google Scholar] [CrossRef]
  44. Gupta, M.; Abu Hussein, N.A.; Fahmi Hawari, H. Graphene derivative coated QCM-based gas sensor for volatile organic compound (VOC) detection at room temperature. Indones. J. Electr. Eng. Comput. Sci. 2020, 18, 1279–1286. [Google Scholar] [CrossRef]
  45. Jiang, T.; Wang, Z.; Li, Z.; Wang, W.; Xu, X.; Liu, X.; Wang, J.; Wang, C. Synergic effect within n-type inorganic–p-type organic nano-hybrids in gas sensors. J. Mater. Chem. C 2013, 1, 3017–3025. [Google Scholar] [CrossRef]
  46. Yan, Y.; Yang, G.; Xu, J.L.; Zhang, M.; Kuo, C.C.; Wang, S.D. Conducting polymer-inorganic nanocomposite-based gas sensors: A review. Sci. Technol. Adv. Mater. 2021, 21, 768–786. [Google Scholar] [CrossRef]
  47. Abu Hussein, N.A.; Hawari, H.F.; Wong, Y.H.; Haseeb, A.S.M.A. Preparation and sensing characterization of hybrid iron oxide/polyaniline/reduced graphene oxide at room temperature. Int. J. Chem. Biochem. Sci. 2021, 20, 90–95. [Google Scholar]
  48. Shafa, M.; Priante, D.; ElAfandy, R.T.; Hedhili, M.N.; Mahmoud, S.T.; Ng, T.K.; Ooi, B.S.; Najar, A. Twofold Porosity and Surface Functionalization Effect on Pt-Porous GaN for High-Performance H2-Gas Sensors at Room Temperature. ACS Omega 2019, 4, 1678–1684. [Google Scholar] [CrossRef]
  49. Nabeel, M.; Varga, M.; Kuzsella, L.; Fiser, B.; Vanyorek, L.; Viskolcz, B. The Effect of Pore Volume on the Behavior of Polyurethane-Foam-Based Pressure Sensors. Polymers 2022, 14, 3652. [Google Scholar] [CrossRef]
Figure 1. Synthesis process of ternary hybrid materials.
Figure 1. Synthesis process of ternary hybrid materials.
Crystals 13 00845 g001
Figure 2. Gas-sensing setup.
Figure 2. Gas-sensing setup.
Crystals 13 00845 g002
Figure 3. XRD analysis of Fe3O4, PANI and S2_10.
Figure 3. XRD analysis of Fe3O4, PANI and S2_10.
Crystals 13 00845 g003
Figure 4. Raman analysis of the single constituents and the ternary hybrid nanocomposite material of sample S2_10.
Figure 4. Raman analysis of the single constituents and the ternary hybrid nanocomposite material of sample S2_10.
Crystals 13 00845 g004
Figure 5. SEM Images of (a) RGO, (b) PANI, (c) Fe3O4 and (dh) S1_5, S2_10, S3_20, S4_30 and S5_40, respectively.
Figure 5. SEM Images of (a) RGO, (b) PANI, (c) Fe3O4 and (dh) S1_5, S2_10, S3_20, S4_30 and S5_40, respectively.
Crystals 13 00845 g005aCrystals 13 00845 g005b
Figure 6. TEM analysis for samples (a) Fe3O4, (b) RGO, (c) PANI and (d) S2_10, and (e) SAED pattern.
Figure 6. TEM analysis for samples (a) Fe3O4, (b) RGO, (c) PANI and (d) S2_10, and (e) SAED pattern.
Crystals 13 00845 g006
Figure 7. Crystallite size and particle comparison from XRD and TEM respectively.
Figure 7. Crystallite size and particle comparison from XRD and TEM respectively.
Crystals 13 00845 g007
Figure 8. The mechanism of the sensor towards acetone vapor.
Figure 8. The mechanism of the sensor towards acetone vapor.
Crystals 13 00845 g008
Figure 9. The response and recovery time in seconds for all samples to (a) 1, (b) 10 and (c) 100 ppm acetone vapor.
Figure 9. The response and recovery time in seconds for all samples to (a) 1, (b) 10 and (c) 100 ppm acetone vapor.
Crystals 13 00845 g009
Figure 10. The response of IDE sensor of all samples to (a) 1, (b) 10 and (c) 100 ppm acetone, respectively.
Figure 10. The response of IDE sensor of all samples to (a) 1, (b) 10 and (c) 100 ppm acetone, respectively.
Crystals 13 00845 g010aCrystals 13 00845 g010b
Figure 11. The stability of all samples to 1 ppm acetone at room temperature for seven weeks.
Figure 11. The stability of all samples to 1 ppm acetone at room temperature for seven weeks.
Crystals 13 00845 g011
Figure 12. The selectivity response of the IDE sensor to 1000 ppm formaldehyde, ethanol and acetone at room temperature.
Figure 12. The selectivity response of the IDE sensor to 1000 ppm formaldehyde, ethanol and acetone at room temperature.
Crystals 13 00845 g012
Table 1. Sample name and description of the ratios used.
Table 1. Sample name and description of the ratios used.
Sample NameDescription
S1_55 wt%Fe3O4-PANI-90 wt%RGO
S2_1010 wt%Fe3O4-PANI-80 wt%RGO
S3_2020 wt%Fe3O4-PANI-60 wt%RGO
S4_3030 wt%Fe3O4-PANI-40 wt%RGO
S5_4040 wt%Fe3O4-PANI-20 wt%RGO
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abu Hussein, N.A.; Wong, Y.H.; Burhanudin, Z.A.; Hawari, H.F. Ternary Hybrid Materials for Highly Sensitive Acetone Sensing at Room Temperature. Crystals 2023, 13, 845. https://doi.org/10.3390/cryst13050845

AMA Style

Abu Hussein NA, Wong YH, Burhanudin ZA, Hawari HF. Ternary Hybrid Materials for Highly Sensitive Acetone Sensing at Room Temperature. Crystals. 2023; 13(5):845. https://doi.org/10.3390/cryst13050845

Chicago/Turabian Style

Abu Hussein, Nurul Athirah, Yew Hoong Wong, Zainal Arif Burhanudin, and Huzein Fahmi Hawari. 2023. "Ternary Hybrid Materials for Highly Sensitive Acetone Sensing at Room Temperature" Crystals 13, no. 5: 845. https://doi.org/10.3390/cryst13050845

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