N-Doped Graphene and Derivatives as Resistive Gas Sensors: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 1 by Ali Mirzaei.

Graphene and its derivatives with a 2D structure are among the most encouraging materials for gas-sensing purposes, because a 2D lattice with high surface area can maximize the interaction between the surface and gas, and a small variation in the carrier concentration of graphene can cause a notable modulation of electrical conductivity in graphene. Furthermore, graphene has a high mobility of electrons and good conductivity.

  • graphene
  • reduced graphene oxide
  • nitrogen
  • doping
  • DFT
  • gas sensor

1. Introduction

Gases are found everywhere and we cannot imagine life without them. However, some gases and organic volatile compounds (VOCs) such as NO2, CO, C6H6, and C7H8 are highly hazardous and their emission into the air leads to air pollution and subsequent problems related to human health. Today, due to the presence of many industries and plants and a large amount of car traffic, access to clean air is difficult; more than 90% of people live in cities with air pollution and, inevitably, they suffer from polluted air. Indeed, air pollution is one of the causes of premature death and sickness globally, and in total, it causes 16% of deaths in the world [1]. While the presence of some gases contributes to air pollution and health problems, some other gases such as H2S are considered biomarkers for the diagnosis of some diseases such as halitosis and Down syndrome [2].
Humans have a sophisticated olfactory system which is constructed from millions of receptor cells. When odorant molecules come into contact with olfactory receptors through the glomerulus, mitral cells send a signal to the brain which will be detected. Hence, we are able to sense hundreds of smells easily and instantly. We can even instantaneously sense very low concentrations of odorants. For instance, we can detect methanethiol in 1 in 2.5 billion units of air [3].
However, in the case of odorless gases and substances such as CO, or extremely low concentrations of gases, we are unable to detect them. Moreover, people cannot be present everywhere. Standardized techniques for the detection of gases such as high-performance gas chromatography and ion chromatography need complex preparation, expensive and large equipment, and cannot provide fast results [5][4]. Hence, the realization of fast dynamics and stable gas sensors to detect low amounts of gases is of importance [6][5]. Considering these factors, various gas and humidity [7,8][6][7] sensors have been introduced which are fabricated from different materials with different sensing principles.
In general, gas sensors are electronic devices which can sense the presence of a target gas in their surrounding atmosphere and generate an electronic signal based on the nature and concentration of the target gas. Resistance or resistive sensors are, nowadays, highly well-liked thanks to their low cost, simple operation, ease of fabrication and measurement, high response, high stability, ease of integration, etc. [9][8]. In [10][9], Bradeen and Bradeen were the first to discover the effect of the surrounding atmosphere on the electrical properties of germanium [11][10]. Later, gas explosions in Japan inspired scientists to develop high-sensitivity gas sensors, and Seiyama et al. [12][11] introduced the first resistance gas sensor using ZnO for the detection of C7H8, CO2, and propane gases. Additionally, Taguchi significantly worked on the development of resistive gas sensors and, finally, he commercialized the first resistive gas sensor based on SnO2.
Mostly, resistive gas sensors are fabricated from semiconducting metal oxides. Nonetheless, they generally show unacceptable selectivity and should be heated to high temperatures to offer their best performance, especially when they are in pure form [13][12]. Thus, other materials such as conducting polymers (CPs) [14][13], carbon-based materials [15][14], MXenes [16][15], and metal chalcogenides [17][16] have also been employed for the realization of gas-sensing devices. Most of these materials can work at low or room temperature, which is a great advantage as it can remarkably decrease the power consumption of fabricated gas sensors. Generally, there are two common configurations for resistive gas sensors: planar and tubular. In both configurations, the sensing layer is deposited on the sensor substrate which is an insulating material, such as alumina equipped with conductive electrodes such as Au and Pt. Additionally, to provide temperature to the gas sensor, a heater is attached on back of the sensor for planar gas sensors, while for tubular gas sensors, generally, heater wire is put inside of the tube and by applying voltage to it, heat is generated via the Joule heating effect.
Thanks to the outstanding progress in the synthesis and discovery of new materials, nowadays, there are many engineering materials with unique features. Among them, carbon-based materials have high potential for sensing studies thanks to their good conductivity, high mechanical properties, relatively low prices, numerous varieties, and high chemical stability [19][17]. Among the different carbon allotropes, graphene with a 2-dimensional (2D) morphology, which was isolated from graphite in 2004 [20][18], is one of the most fascinating materials. Graphene refers to a single layer of “C” atoms which are arranged into a 2D honeycomb structure, and it is regarded as a basic building block for graphite (3D), carbon nanotubes (1D), and fullerenes (0D) [21][19].
Nowadays, graphene can be easily synthesized; however, it is easily agglomerated due to Van der Waals interactions. Unfortunately, graphene suffers from the lack of band gap, which limits its application as a gas sensor, especially in pristine form. Furthermore, it has poor water solubility. Hence, it can be partially oxidized by powerful oxidants and, in this way, graphene oxide (GO) is synthesized and different functional groups such as epoxy groups on its surface lead to water solubility. Due to the presence of carboxyl, carbonyl, and epoxide groups on the surface of GO, it is an electrical insulator because of the disruption of its sp2 bonding networks [23,24][20][21]. Therefore, it has limited potential for gas-sensing purposes. Reduced graphene oxide (rGO), a reduced from of graphene oxide (GO), can also be considered as promising material for gas-sensing applications, mainly thanks to having many functional groups and defects on it [25][22]. However, even though an exact comparison between N-doped graphene and rGO is difficult, rGO has no band gap, whereas N-doped graphene has a small band gap. Additionally, N-doped graphene has greater electron mobility than rGO. Furthermore, N-doped graphene may provide more active sites for gas molecules than rGO. In both materials, however, there are more defects than pristine graphene [26,27][23][24].
Referring back to graphene, it has been considered for different applications thanks to its nontoxicity, high transparency, high mechanical properties, high electrical and thermal conductance, and high carrier mobility at room temperature [28][25]. Accordingly, it has been used in different areas such as gas sensors [29][26]. Generally, doping is a good technique to modify the features of different materials [31][27]. There are two types of doping in graphene. Metallic-cluster-induced doping, gate-controlled doping, or substrate-induced doping are known as electrical doping. Indeed, this type of doping does not affect the composition of graphene. Another type of doping, namely chemical doping, tunes the lattice structure via chemical routes such as substitution with foreign atoms [32][28].

2. N-Doped Graphene and Its Derivatives as Resistive Gas Sensors: Theoretical Studies

With the progressive advancement of computational power and the availability of better supercomputers, theoretical studies of materials have become more feasible. Since gas sensor reactions include chemical reactions in which bonds are formed or broken and often result in a transfer of charge, the first principle (ab initio) or a quantum approach is preferred over a classical mechanics method. Quantum approach methods are useful for studying gas–surface interactions owing to their ability to model the adsorption mechanisms, the adsorption energy, charge transfer, electronic modification after gas adsorption, the structural changes, and the electronic properties involved. In this regard, density functional theory (DFT) calculations are considered a powerful technique to carry out many simulations, including the gas-sensing capability of different materials, without any side effects. DFT calculations are required for understanding experimental results at the molecular level, or as a predictive tool for the design of new sensing materials [73,74,75][29][30][31].
The adsorption of some gases, namely CO, NO, O2, and NO2, on defective graphene and N-doped graphene was investigated through DFT calculations. It was theoretically revealed that all CO, NO, and O2 gases were chemisorbed on the defective graphene with large charge transfers. This can lead to the poor selectivity of CO in the presence of other gases. In contrast, only CO molecule with a large charge transfer was chemisorbed on the N-doped graphene, while other gases were only weakly adsorbed on it, reflecting good selectivity for CO gas sensing. Hence, through DFT calculations, it was shown that with a simple doping of “N” in a graphene lattice, it is possible to improve the selectivity to CO gas [76][32].
Based on DFT studies, the adsorption of CO molecules on pristine, N-doped, as well as Al-doped graphene was studied. The results showed that the most stable adsorption configuration was on top of a carbon (N or Al) atom of graphene sheets. The adsorption energies of CO molecules on pristine, N-doped, and Al-doped graphene were −0.01 eV, −0.03 eV, and −2.69 eV, respectively. In fact, the adsorption of CO molecules on pristine and N-doped graphene was physical adsorption, while it was chemisorbed on the Al-doped graphene. Moreover, calculations on band structure, density of states (DOS), and differential charge density showed that the adsorption of CO on Al-doped graphene was remarkably different from other materials in their study [77][33]. In another work, the adsorption of HCHO on Si-, Al-, Cr-, B-, N-, Au-, and Mn-doped graphene was studied using DFT calculations. It was reported that the HCHO molecule was weakly adsorbed on B- and N-doped graphene, while HCHO chemisorption was predicted on Si-, Al-, Cr-, Mn-, and Au-doped graphene. Additionally, based on adsorption geometries, charge transfers, adsorption energies, and density of states, it was found that Al and Mn dopants are the best ones for sensing enhancement towards HCHO gas [78][34]. The next work was related to an adsorption study of CO2 on the P-, N-, B-, and Al- doped monovacancy graphene through DFT calculations. The band gap was changed from 0.004 to 0.478 eV and the electrical conductivity was changed from metallic to semiconducting when a monovacancy was created in graphene. Furthermore, the doped graphene had a larger band gap relative to its undoped counterpart. The authors reported that Al-doped monovacancy graphene was better than other gas sensors for CO2 gas sensing due to the suitable adsorption strength and band gap. Before the adsorption of CO2 gas, the band gaps of N-doped monovacancy graphene and P-doped monovacancy graphene had no change, but those of B-doped monovacancy graphene and Al-doped monovacancy graphene changed. The band gap of Al-doped monovacancy graphene changed from almost 0 to 0.229 eV [79][35]. Hence, the electrical features were significantly changed, showing the high potential of Al-doped monovacancy graphene for CO2 gas-sensing application.
BF3 and BCl3 are colorless gases and useful reagents in organic synthesis [80][36]. The energetic and electronic properties of pristine and N-doped graphene upon the adsorption of BC13, B(OCH3)3, and BF3 gases were studied using DFT calculations. The computed dipole moment revealed that the amount of change in dielectric of the sensing material depends on the type of molecule. Furthermore, it was reported that N-doped graphene had much higher adsorption energy and higher net charge transfer relative to pristine graphene thanks to Lewis acid–base interactions between these gases and the sensing material. Additionally, the Lewis acidity increased as follows: BF3 < BC13 < B(OCH3)3 with adsorption energies of −8.7, −18.3, and −26.5 kJ/mol, respectively, revealing better adsorption of the B(OCH3)3 on the N-doped graphene. Furthermore, very low adsorption energies were calculated on pristine graphene, suggesting a poor sensing response of pristine graphene to these gases [81][37].
Based on the DFT calculation results, it can be concluded that, overall, N doping can enhance the adsorption of gas molecules on graphene or rGO, relative to pristine (undoped) graphene or rGO. However, it should be also mentioned that, in comparison to metallic dopants such as Al, Ti, or even B dopant, generally, N-doped graphene shows weaker adsorption of the target gas molecules. However, the adsorption behavior of codoped graphene such as Ti/N, and codoping in graphene in general, need to be further explored using DFT calculations.

3. N-Doped Graphene and Its Derivatives as Resistive Gas Sensors: Experimental Studies

Graphene quantum dots (GQDs) have lateral dimensions smaller than 10 nm and the combination of QDs and graphene characteristics in these materials leads to revealing enhanced features such as low toxicity, chemical stability, high strength, tunable energy band gaps, outstanding optical properties, and excellent microelectronic properties [92][38]. In this regard, Masemola et al. [87][39] prepared a N-doped GQDs/PANI composite for ethanol-sensing studies.
The NGQDs/PANI sensor revealed a better response compared to that of pristine PANI along with a faster response time and recovery time of 85 s and 62 s, respectively, to 100 ppm ethanol vapor at 25 °C. The enhanced response of the composite sensor was related to the incorporation of N in the GQDs. It was related to the enhanced conductive pathways between the PANI and N-doped GQDs, which induced conductive pathways created by strong synergetic effects through interchain π interactions between the NGQDs and PANI. Furthermore, the surface area of the composite sensor (41.0 m2 g−1) was higher than that of bare PANI (31.2 m2 g−1), leading to higher adsorption of ethanol molecules on the composite sensor.
CPs are used in gas sensors due their tunable properties, low cost, and room-temperature operation [94][40]. In another study, Gavgani et al. [95][41] realized VOC gas sensors based on pristine N-doped GQDs, poly (3,4 ethylenedioxythiophene)-poly(styrenesulfonate) (PEDOT-PSS), and N-doped GQDs/PEDOT-PSS nanocomposites. The fabricated sensors at room temperature were exposed to 100 ppm acetone, toluene, propanol, chlorobenzene, methanol, ethanol, and water. It was reported that the N-doped GQDs/PEDOT-PSS nanocomposite gas sensor showed a good response to methanol, whereas it revealed almost no response to other gases. The response of the optimal sensor to 100 ppm methanol was 154.4%, which was 12 times as high as that of the PEDOT-PSS sensor. The sensing mechanism of the optimal sensor was described using the following reasons: (i) The swelling of CPs due to the diffusion of methanol into the CPs. In this case, the interchain distance of CPs increases and the electron hopping process becomes more difficult. (ii) N-doped GQDs integrated into the CP matrix acted as conductive paths, favoring the electron hopping. Therefore, a combination of the above two processes leads to the modulation of the electrical resistance in the presence of methanol gas.
Nitrogen dioxide (NO2), with a highly toxic nature, emits from fossil fuels and can damage the human respiratory system [96][42]. In this regard, a series of N-doped GQD 3-dimensional ordered macroporous (3DOM) In2O3 composites were synthesized for NO2 gas-sensing studies. The results showed that the sensor exhibited a better response to NO2 gas with fast and recovery times and a low working temperature (100 °C), relative to pristine In2O3 and rGO/3DOM In2O3 sensors. Furthermore, it showed good selectivity and high long-term stability. The conductivity enhancement due to the presence of N-doped GQDs and the generation of heterojunctions between In2O3 and N-GQDs, as well as the presence of macrospores, accounted for the enhanced response to NO2 gas [97][43].
P-type rGO and N-doped rGO were explored for NO gas detection at 25 °C. The response of the N-rGO sensor to 1000 ppb NO (1.7) was higher than that of the rGO sensor (0.012) and it was even able to detect 0.4 ppm NO gas. The increase in the NO gas-sensing performance was related to N doping. First, the N doping produced polarization in the sp2 carbon network and further affected the physical and the electrical properties of the sensor. Second, the improved gas response was related to the improved active sites of pyridinic N and/or pyrrolic N. In particular, pyridinic N, where the N atom is bonded to two C atoms and donates one p-electron to the aromatic π system, modified the band structure of carbon and lowered the work function. Additionally, the electron density on the neighboring C atom decreased, which resulted in the transfer of electrons from the C to N atoms, and N back-donate electrons to the neighboring C pz orbitals. These processes facilitated the dissociation of oxygen molecules on the C atoms and helped to form a strong bond between O and C atoms [98][44].
TiO2 is an n-type semiconductor which offers high chemical stability and excellent electronic properties. Yan et al. [99][45] investigated the sensing properties of C- and N-doped rGO/TiO2 composites with special exposed facets synthesized via a hydrothermal method at 180 °C using HF. The gas-sensing performance of the N-doped rGO/TiO2 sensor was better than that of the C-doped rGO/TiO2. The N-doped rGO/TiO2 sensor exhibited the highest response to isopropanol, ethanol, and acetone gases at 210, 240, and 270 °C, respectively. The N-doped rGO/TiO2 sensor exhibited the highest response to 100 ppm acetone with a response of 14.12, while its responses to ethanol and isopropanol were 5.85 and 3.47, respectively. 
Lin et al. [100][46] fabricated mesoporous p-type Co3O4/N-doped rGO nanocomposites via a metal–organic framework (MOF) template. Elemental mapping revealed the uniform distribution of Co, O, C, and N elements in the sample. The effect of initial rGO concentration on the gas-sensing characteristics of the composite sensor was studied. It was found the Co3O4/N-doped rGO-0.5 (rGO mass = 0.5 mg) nanocomposite sensor had a better response to ethanol than other sensors, and a response of 44.5 to 200 ppm ethanol at 200 °C was obtained. The enhanced sensing response was related to the presence of rGO with high electron mobility, which facilitated the migration of carriers in the sensing process, leading to improved sensing performance. Furthermore, the coupling effect between Co3O4 and N-doped rGO enhanced the oxygen reduction ability in the gas-sensing process and increased the transfer rate of charge carriers. In addition, the sensor had a high specific surface area, and so provided more effective active sites during ethanol adsorption than the other samples. The N atoms were mainly bonded with rGO as pyridine-N. They were bonded with the “C” atoms of the rGO sheets, which provided lone pair electrons to increase the electronic partial density of states near the Fermi level. The higher density of states facilitated the electron flow during the sensing reactions.
Modak et al. [101][47] introduced NO2 gas sensors using a SnO2-N-doped rGO composite for NO2 sensing. SnO2-rGO composites with three amounts of SnO2 (low, medium, and high) were synthesized via a hydrothermal reaction. The samples were coded as SR1, SR2, and SR3. For N-doped rGO, urea was used as a source of N atoms. They were coded as SRN1, SRN2, and SRN3 based on 0.09, 0.18, and 0.27 mg of urea, respectively. 
N doping led to the generation of more active sites which were beneficial for NO2 adsorption. Additionally, the higher surface area and higher pore volume of the SRN2 sensor facilitated a higher response to NO2 gas. Furthermore, the large number of defects and oxygen vacancies in SRN2 enhanced the gas response. Indeed, N-doped rGO had a greater number of defects in the rGO sheets, which acted as adsorption sites for the NO2 gas. Additionally, the SRN2 sensor had higher conductivity compared to SR sensors, owing to the presence of a lone pair of electrons on the N atom.
Nie et al. [102][48] synthesized N- and Si-codoped (N/Si-doped) graphene. The N atoms acted as the active sites for the NO2 adsorption, whereas the Si atoms remarkably modified the electronic features of the graphene. The N/Si-doped graphene exhibited p-type conductivity, indicating that the hole doping from Si atoms overcame the electron doping from N atoms. The sensor was able to detect NO2 gas at 25 °C.
NH3, as a toxic gas, is an irritant to the eyes, skin, and respiratory system [103][49]. Among different CPs, PANI is popular thanks to merits such as the ease of synthesis, low cost, stability, and good response to NH3. Nevertheless, in pristine form, it has a response [104][50]. Therefore, it is often used in composite form. In this regard, Gavgani and coworkers [105][51] introduced a flexible NH3 sensor using a S/N-codoped GQDs/PANI composite. The response of the S/N-codoped GQDs/PANI sensor to 100 ppm NH3 was 5-fold higher than that of the PANI sensor, with a faster response time and recovery time. Additionally, the S/N-doped GQDs/PANI hybrid gas sensor showed better selectivity to NH3 gas relative to the PANI sensor.
The increased NH3-sensing features were attributed to the swelling of the PANI in the presence of NH3, as well as the reversible acid–base doping/dedoping of the PANI in the presence of NH3 gas, which changed it from a conductive state to an insulating state, leading to an increase in the resistance and formation of heterojunctions between the PANI and the S/N-codoped GQDs. In another study, an NH3 gas sensor based on a combination of PANI and different amounts of N-doped GQDs was fabricated and Ag or Al was used for the electrodes. The sensor with Ag and Al contact showed Ohmic and Schottky behaviors, respectively. The sensor with a Ag electrode had a much higher response to NH3 than that of the sensor with an Al electrode. The (50 wt%) N-GQDs/PANI sensor with a Ag electrode displayed a high response of 110.92 to 1500 ppm NH3 at room temperature, which was around 30% higher than that of the Al electrode [106][52]. In another research, the sensing response of a PEDOT-PSS sensor to 1500 ppm NH3 at room temperature increased from 30.13% to 212.32% with 50 wt% N-doped GQDs. Furthermore, the sensor with N-GQD showed faster dynamics and higher stability [107][53]. Purbia et al. reported N-doped GQDs/SnO2 nanocomposites for the detection of NO2 gas at 150 °C. The response to 100 ppb NO2 gas was 292 with a response time of ~3 min. Furthermore, it revealed good selectivity to NO2 gas. The formation of SnO2-N-doped GQDs, the presence of defects, and the high surface area of the gas sensor led to a better response in the composite sensor relative to the bare SnO2 gas sensor [108][54].
3. Conclusions

4. Conclusions

There are various methods to fabricate N-doped graphene or rGO. N doping in graphene and rGO can not only induce defects, but can also increase the electrical conductivity and induce n-type conductivity in graphene. Through N doping, it is feasible to tune the band gap of graphene and rGO. Furthermore, N dopants are cheaper than noble metals such as Au, Pt, and Pd.
 

References

  1. Navale, S.; Mirzaei, A.; Majhi, S.M.; Kim, H.W.; Kim, S.S. State-of-the-art research on chemiresistive gas sensors in Korea: Emphasis on the achievements of the research labs of professors Hyoun Woo Kim and Sang Sub Kim. Sensors 2022, 22, 61.
  2. Navale, S.; Shahbaz, M.; Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S. CuxO nanostructure-based gas sensors for H2S detection: An overview. Chemosensors 2021, 9, 127.
  3. Mirzaei, A.; Ansari, H.R.; Shahbaz, M.; Kim, J.-Y.; Kim, H.W.; Kim, S.S. Metal Oxide Semiconductor Nanostructure Gas Sensors with Different Morphologies. Chemosensors 2022, 10, 289.
  4. Mirzaei, A.; Kim, S.S.; Kim, H.W. Resistance-based H2S gas sensors using metal oxide nanostructures: A review of recent advances. J. Hazard. Mater. 2018, 357, 314–331.
  5. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Recent advances in energy-saving chemiresistive gas sensors: A review. Nano Energy 2021, 79, 105369.
  6. Chen, C.; Jiang, M.; Luo, X.; Tai, H.; Jiang, Y.; Yang, M.; Xie, G.; Su, Y. Ni-Co-P hollow nanobricks enabled humidity sensor for respiratory analysis and human-machine interfacing. Sens. Actuators B Chem. 2022, 370, 132441.
  7. Su, Y.; Liu, Y.; Li, W.; Xiao, X.; Chen, C.; Lu, H.; Yuan, Z.; Tai, H.; Jiang, Y.; Zou, J.; et al. Sensing-transducing coupled piezoelectric textiles for self-powered humidity detection and wearable biomonitoring. Mater. Horiz. 2023, 10, 842–851.
  8. Mirzaei, A.; Kordrostami, Z.; Shahbaz, M.; Kim, J.-Y.; Kim, H.W.; Kim, S.S. Resistive-Based Gas Sensors Using Quantum Dots: A Review. Sensors 2022, 22, 4369.
  9. Navale, S.; Shahbaz, M.; Mirzaei, A.; Kim, S.S.; Kim, H.W. Effect of Ag addition on the gas-sensing properties of nanostructured resistive-based gas sensors: An overview. Sensors 2021, 21, 6454.
  10. Brattain, W.H.; Bardeen, J. Surface properties of germanium. Bell Syst. Tech. J. 1953, 32, 1–41.
  11. Seiyama, T.; Kato, A.; Fujiishi, K.; Nagatani, M. A new detector for gaseous components using semiconductive thin films. Anal. Chem. 1962, 34, 1502–1503.
  12. Zhang, J.; Liu, X.; Neri, G.; Pinna, N. Nanostructured materials for room-temperature gas sensors. Adv. Mater. 2016, 28, 795–831.
  13. Wong, Y.C.; Ang, B.C.; Haseeb, A.S.M.A.; Baharuddin, A.A.; Wong, Y.H. Conducting polymers as chemiresistive gas sensing materials: A review. J. Electrochem. Soc. 2020, 167, 037503.
  14. Llobet, E. Gas sensors using carbon nanomaterials: A review. Sens. Actuators B Chem. 2013, 179, 32–45.
  15. Deshmukh, K.; Kovářík, T.; Pasha, S.K.K. State of the art recent progress in two dimensional MXenes based gas sensors and biosensors: A comprehensive review. Coord. Chem. Rev. 2020, 424, 213514.
  16. Kumar, R.; Goel, N.; Hojamberdiev, M.; Kumar, M. Transition metal dichalcogenides-based flexible gas sensors. Sens. Actuators A Phys. 2020, 303, 111875.
  17. Zhang, Y.; Zhang, Y.; Zhang, D.; Li, S.; Jiang, C.; Su, Y. Confinement preparation of Au nanoparticles embedded in ZIF-67-derived N-doped porous carbon for high-performance detection of hydrazine in liquid/gas phase. Sens. Actuators B Chem. 2019, 285, 607–616.
  18. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669.
  19. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191.
  20. Kaur, R.; Arora, A.; Tripathi, S.K. Fabrication and characterization of metal insulator semiconductor Ag/PVA/GO/PVA/n-Si/Ag device. Microelec. Eng. 2020, 233, 111419.
  21. Ren, P.G.; Yan, D.X.; Ji, X.; Chen, T.; Li, Z.M. Temperature dependence of graphene oxide reduced by hydrazine hydrate. Nanotechnology 2010, 22, 055705.
  22. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Reduced graphene oxide (rGO)-loaded metal-oxide nanofiber gas sensors: An overview. Sensors 2021, 21, 1352.
  23. He, D.; Jiang, Y.; Lv, H.; Pan, M.; Mu, S. Nitrogen-doped reduced graphene oxide supports for noble metal catalysts with greatly enhanced activity and stability. Appl. Catal. B Environ. 2013, 132, 379–388.
  24. Luo, Z.; Lim, S.; Tian, Z.; Shang, J.; Lai, L.; MacDonald, B.; Fu, C.; Shen, Z.; Yu, T.; Lin, J. Pyridinic N doped graphene: Synthesis, electronic structure, and electrocatalytic property. J. Mater. Chem. 2011, 21, 8038–8044.
  25. Kazemi, A.; Rodner, M.; Fadavieslam, M.R.; Kaushik, P.D.; Ivanov, I.G.; Eriksson, J.; Syväjärvi, M.; Yakimova, R.; Yazdi, G.R. The effect of Cl- and N-doped MoS2 and WS2 coated on epitaxial graphene in gas-sensing applications. Surf. Interfaces 2021, 25, 101200.
  26. Varghese, S.S.; Lonkar, S.; Singh, K.K.; Swaminathan, S.; Abdala, A. Recent advances in graphene based gas sensors. Sens. Actuators B Chem. 2015, 218, 160–183.
  27. Wang, C.-N.; Li, Y.-L.; Gong, F.-L.; Zhang, Y.-H.; Fang, S.-M.; Zhang, H.-L. Advances in doped ZnO nanostructures for gas sensor. Chem. Rec. 2020, 20, 1553–1567.
  28. Lv, R.; Terrones, M. Towards new graphene materials: Doped graphene sheets and nanoribbons. Mater. Lett. 2012, 78, 209–218.
  29. Yan, Z.; Zhang, Y.; Kang, W.; Deng, N.; Pan, Y.; Sun, W.; Ni, J.; Kang, X. TiO2 Gas sensors combining experimental and DFT calculations: A review. Nanomaterials 2022, 12, 3611.
  30. Spencer, M.J. Gas sensing applications of 1D-nanostructured zinc oxide: Insights from density functional theory calculations. Prog. Mater. Sci. 2012, 57, 437–486.
  31. Cruz-Martínez, H.; Rojas-Chávez, H.; Montejo-Alvaro, F.; Peña-Castañeda, Y.A.; Matadamas-Ortiz, P.T.; Medina, D.I. Recent developments in graphene-based toxic gas sensors: A theoretical overview. Sensors 2021, 21, 1992.
  32. Ma, C.C.; Shao, X.H.; Cao, D.P. Nitrogen-doped graphene as an excellent candidate for selective gas sensing. Sci. China Chem. 2014, 57, 911–917.
  33. Wang, W.; Zhang, Y.; Shen, C.; Chai, Y. Adsorption of CO molecules on doped graphene: A first-principles study. AIP Adv. 2016, 6, 025317.
  34. Liu, X.-Y.; Zhang, J.-M. Formaldehyde molecule adsorbed on doped graphene: A first-principles study. Appl. Surf. Sci. 2014, 293, 216–219.
  35. Zheng, Z.; Wang, H. Different elements doped graphene sensor for CO2 greenhouse gases detection: The DFT study. Chem. Phys. Lett. 2019, 721, 33–37.
  36. Hamaya, N.; Ishizuka, M.; Onoda, S.; Guishan, J.; Ohmura, A.; Shimizu, K. Pressure-induced phase transition, metallization, and superconductivity in boron triiodide. Phys. Rev. B 2010, 82, 094506.
  37. Rad, A.S.; Shadravan, A.; Soleymani, A.A.; Motaghedi, N. Lewis acid-base surface interaction of some boron compounds with N-doped graphene; first principles study. Curr. Appl. Phys. 2015, 15, 1271–1277.
  38. Kaur, M.; Kaur, M.; Sharma, V.K. Nitrogen-doped graphene and graphene quantum dots: A review on synthesis and applications in energy, sensors and environment. Adv. Colloid Interface Sci. 2018, 259, 44–64.
  39. Rad, A.S.; Esfahanian, M.; Maleki, S.; Gharati, G. Application of carbon nanostructures toward SO2 and SO3 adsorption: A comparison between pristine graphene and N-doped graphene by DFT calculations. J. Sulfur Chem. 2016, 37, 176–188.
  40. Kumar, V.; Mirzaei, A.; Bonyani, M.; Kim, K.-H.; Kim, H.W.; Kim, S.S. Advances in electrospun nanofiber fabrication for polyaniline (PANI)-based chemoresistive sensors for gaseous ammonia. TrAC-Trends Anal. Chem. 2020, 129, 115938.
  41. Gavgani, J.N.; Dehsari, H.S.; Hasani, A.; Mahyari, M.; Shalamzari, E.K.; Salehi, A.; Taromi, F.A. A room temperature volatile organic compound sensor with enhanced performance, fast response and recovery based on N-doped graphene quantum dots and poly(3,4-ethylenedioxythiophene)–poly(styrenesulfonate) nanocomposite. RSC Adv. 2015, 5, 57559–57567.
  42. Duan, K.; Li, W.; Zhu, C.; Li, J.; Xu, J.; Wang, X. Promoting sensitivity and selectivity of NO2 gas sensor based on (P, N)-doped single-layer WSe2: A first principles study. Results Phys. 2022, 34, 105296.
  43. Lv, Y.-K.; Li, Y.-Y.; Zhou, R.-H.; Pan, Y.-P.; Yao, H.-C.; Li, Z.-J. N-doped graphene quantum dot-decorated three-dimensional ordered macroporous In2O3 for NO2 sensing at low temperatures. ACS Appl. Mater. Interfaces 2020, 12, 34245–34253.
  44. Chang, Y.-S.; Chen, F.-K.; Tsai, D.-C.; Kuo, B.-H.; Shieu, F.-S. N-doped reduced graphene oxide for room-temperature NO gas sensors. Sci. Rep. 2021, 11, 20719.
  45. Yan, W.-Y.; Zhou, Q.; Chen, X.; Huang, X.-J.; Wu, Y.-C. C-doped and N-doped reduced graphene oxide/TiO2 composites with exposed (0 0 1) and (1 0 1) facets controllably synthesized by a hydrothermal route and their gas sensing characteristics. Sens. Actuators B Chem. 2016, 230, 761–772.
  46. Lin, G.; Wang, H.; Lai, X.; Yang, R.; Zou, Y.; Wan, J.; Liu, D.; Jiang, H.; Hu, Y. Co3O4/N-doped RGO nanocomposites derived from MOFs and their highly enhanced gas sensing performance. Sens. Actuators B Chem. 2020, 303, 127219.
  47. Modak, M.; Jagtap, S. Low temperature operated highly sensitive, selective and stable NO2 gas sensors using N-doped SnO2-rGO nanohybrids. Ceram. Int. 2022, 48, 19978–19989.
  48. Niu, F.; Liu, J.-M.; Wang, W.; Song, W.-G. Nitrogen and silica co-doped graphene nanosheets for NO2 gas sensing. J. Mater. Chem. A 2013, 1, 6130–6133.
  49. Timmer, B.; Olthuis, W.; Van Den Berg, A. Ammonia sensors and their applications—A review. Sens. Actuators B Chem. 2005, 107, 666–677.
  50. Wu, Z.; Chen, X.; Zhu, S.; Zhou, Z.; Yao, Y.; Quan, W.; Liu, B. Enhanced sensitivity of ammonia sensor using graphene/polyaniline nanocomposite. Sens. Actuators B Chem. 2013, 178, 485–493.
  51. Gavgani, J.N.; Hasani, A.; Nouri, M.; Mahyari, M.; Salehi, A. Highly sensitive and flexible ammonia sensor based on S and N co-doped graphene quantum dots/polyaniline hybrid at room temperature. Sens. Actuators B Chem. 2016, 229, 239–248.
  52. Hakimi, M.; Salehi, A.; Boroumand, F.A.; Mosleh, N. Fabrication of a room temperature ammonia gas sensor based on polyaniline with N-doped graphene quantum dots. IEEE Sens. J. 2018, 18, 2245–2252.
  53. Hakimi, M.; Salehi, A.; Boroumand, F.A. Fabrication and characterization of an ammonia gas sensor based on PEDOT-PSS with N-doped graphene quantum dots dopant. IEEE Sens. J. 2016, 16, 6149–6154.
  54. Purbia, R.; Kwon, Y.M.; Kim, H.D.; Lee, Y.S.; Shin, H.; Baik, J.M. Zero-dimensional heterostructures: N-doped graphene dots/SnO2 for ultrasensitive and selective NO2 gas sensing at low temperatures. J. Mater. Chem. A 2020, 8, 11734–11742.
More
Video Production Service