Next Article in Journal
Recent Advances in Benzodiazepine Electroanalysis
Previous Article in Journal
Investigation of Distinct Odor Profiles of Blood over Time Using Chemometrics and Detection Canine Response
Previous Article in Special Issue
Electrochemical Monitoring of Bisphenol A Degradation in Leachate by Trichoderma harzianum Using a Sensitive Sensor of Type SPE in Microbial Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Noble Metal-Decorated In2O3 for NO2 Gas Sensor: An Experimental and DFT Study

by
Parameswari Raju
1,
Jafetra Rambeloson
1,
Dimitris E. Ioannou
1,
Abhishek Motayed
2 and
Qiliang Li
3,*
1
Department of Electrical and Computer Engineering, George Mason University, Fairfax, VA 22030, USA
2
N5 Sensors, Inc., Rockville, MD 20850, USA
3
Advanced Manufacturing and Robotics, Peking University, Beijing 100871, China
*
Author to whom correspondence should be addressed.
Chemosensors 2025, 13(9), 350; https://doi.org/10.3390/chemosensors13090350
Submission received: 20 June 2025 / Revised: 24 July 2025 / Accepted: 5 September 2025 / Published: 11 September 2025
(This article belongs to the Special Issue Nanomaterial-Based Sensors: Design, Development and Applications)

Abstract

Indium oxide-based gas sensors have been proven to be a promising material for detecting nitrogen dioxide (NO2) gas because of its wide bandgap and stability. In this paper, the enhancement mechanism for the sensitivity of indium oxide NO2 gas sensors was systematically investigated using density functional theory (DFT) calculations and experimental validation with noble metals like Au, Ag, Pt, Pd, and Cu. We have fabricated a GaN nanowire-based NO2 gas sensor functionalized with In2O3 and decorated with noble metals using a standard fabrication technique. Experimental tests showed that Au/In2O3 sensors exhibited the highest response of 38.9% followed by bare In2O3 with 10% for 10 ppm NO2 at room temperature. The sensing properties were mainly attributed to a spillover effect or catalytic performance of Au with In2O3. The adsorption energies, charge transfers, and band gap confirm the enhanced sensing capability of Au-decorated Indium oxide for a NO2 gas sensor.

1. Introduction

The early detection and continuous monitoring of toxic gases is of paramount importance, due to their potential harmful effects on both human health and the environment. Many industrial processes, vehicular emissions, and natural events contribute to the release of hazardous gases such as nitrogen dioxide (NO2), carbon monoxide (CO), chlorine (Cl2), and sulfur dioxide (SO2). Prolonged exposure to these pollutants can lead to respiratory diseases, cardiovascular issues, and environmental degradation including acid rain and photochemical smog. Consequently, there is a growing need for reliable, sensitive, and cost-effective gas sensors that can operate in diverse environmental conditions and detect low concentrations of harmful gases with high accuracy [1,2,3,4,5].
Among various types of gas sensors, metal oxide semiconductor (MOS)-based sensors have garnered significant attention in recent years. Their popularity stems from several advantages, such as high surface-to-volume ratio, low cost, fast response and recovery times, ease of fabrication, and tunable sensitivity through structural and compositional modifications. These sensors operate on a simple yet effective principle: the change in the electrical resistance of the sensing material upon exposure to target gases. In ambient air, oxygen molecules are adsorbed onto the surface of the metal oxide and capture free electrons from the conduction band, forming ionized oxygen species. This process leads to the formation of an electron depletion layer at the surface, which increases the sensor’s resistance. Upon exposure to reducing gases like CO or SO2, the gases react with the adsorbed oxygen, releasing the trapped electrons back into the conduction band, thereby reducing the resistance. In contrast, oxidizing gases like NO2 and O3 can further extract electrons, increasing the resistance [5,6,7,8,9,10].
Enhancing the performance of MOS gas sensors is a major area of research. Structural modifications such as transforming the sensing layer into nanowires, nanoribbons, ultrathin nanotubes, or two-dimensional flakes significantly increase the surface area available for gas interaction. Additionally, doping or decorating the sensor surface with noble or transition metals such as Ag, Pd, Cu, and Au improves the sensor’s response, due to phenomena such as the spillover effect, electronic sensitization, and catalytic enhancement. These modifications not only increase the number of active sites for gas adsorption but also alter the electronic structure of the sensing material, thereby improving sensitivity and selectivity. Light activation is another method used to enhance gas-sensing performance by generating additional charge carriers or activating the catalyst at lower operating temperatures [11,12,13,14,15].
Commonly used metal oxides in gas sensing include SnO2, In2O3, WO3, ZnO, and TiO2. Among them, indium oxide (In2O3), an n-type semiconductor with a wide bandgap (Eg = 3.55–3.75 eV), is particularly attractive due to its excellent photocatalytic activity, chemical stability, and high electron mobility [16,17,18,19,20]. Several studies have demonstrated that modifying In2O3 with noble or transition metals significantly enhances its sensing performance. For example, Ag-decorated In2O3 exhibited improved adsorption energy for formaldehyde (HCHO), changing from −1.6 eV to −0.84 eV, owing to an increase in surface area and enhanced catalytic activity [21]. Similarly, In2O3 decorated with CuO nanowires via hydrothermal synthesis showed improved gas response due to the formation of p-n heterojunctions and increased active sites [22,23].
Recent advancements in heterostructure engineering have further boosted the performance of In2O3-based sensors. Rb-loaded ZnO/In2O3 heterojunctions demonstrated a sensor response of 24.2 for 1 ppm NO2 with a rapid response time of just 55 s [8]. Another notable example includes In2O3–graphene–Cu nanocomposites, which achieved a sensitivity of 46 ppm to NO2, attributed to the synergistic effect of highly defective graphene and the catalytic nature of Cu nanoparticles [6]. Similarly, a modified WS2 microsheet with In2O3 and Au yielded an impressive response of 46.91 for 50 ppm NO2, mainly due to the increased ionized oxygen adsorption sites [9]. Furthermore, the performance of a Pd-AlGaN/GaN NO2 sensor was enhanced by optimizing the gate bias voltage, achieving a sensitivity of 91.6% with a reduced response time of just 9 s at −1V bias [24]. ITO thin films deposited on AlN ceramic substrates exhibited maximum responses of 2.21 to 2000 ppm H2 at 400 °C, 2.39 to 1000 ppm NH3 at 350 °C, and 2.14 to 100 ppm NO2 at 350 °C, demonstrating their effective gas-sensing performance at elevated temperatures [25]. NO2 gas sensors based on laser-induced graphene (LIG) and its heterostructure with SnO2 were developed for room-temperature operation. The sensors feature a highly porous architecture synthesized through a one-step laser-scribing process. The LIG/SnO2 hybrid sensor demonstrated effective NO2 detection at room temperature. With increasing temperature, the sensor’s response improved; however, its sensitivity to different NO2 concentrations decreased. Additionally, in humid environments, the response rate increased. Selectivity tests using CO2 showed no notable response, confirming the sensor’s specificity toward NO2 [26,27]. This study investigates the gas-sensing performance of tin oxide (SnO2) nanoparticles for the detection of NO2, NH3, CO, and H2S across a range of operating temperatures. SnO2 powders were synthesized via the sol–gel method, and thin films were deposited using spin coating. At room temperature, the sensor exhibited sensitivity to NO2 at a low concentration of 2 ppm, with response and recovery times of 184 s and 432 s [28]. Light-enhanced ultra-sensitive NO2 gas detection was achieved using flexible In2O3/MXene-based sensors, demonstrating excellent linearity in the 5–100 ppb range and a remarkably low detection limit of 0.79 ppb [29].
To gain a deeper understanding of the sensing mechanisms and to guide the design of novel sensor materials, first-principles calculations such as with the application of density functional theory (DFT) have been extensively employed. DFT provides valuable insights into the electronic and structural properties of materials, including adsorption energy, charge transfer, adsorption distance, energy gap, and density of states (DOS). These computational studies allow researchers to evaluate various combinations of sensing materials and predict their sensing behavior before experimental realization. By simulating gas adsorption on different material surfaces, DFT helps in identifying the optimal doping levels, suitable metal decorations, and appropriate heterostructure configurations for enhanced sensitivity and selectivity.
In this study, we focus on improving the NO2 gas-sensing performance of indium oxide through decoration with noble metals. Our aim is to systematically investigate the interaction of NO2 with modified In2O3 surfaces using both experimental characterization and DFT-based simulations. The outcomes of this research are expected to contribute to the development of next-generation gas sensors that are not only highly sensitive and selective but also energy-efficient and capable of real-time monitoring in practical environments.

2. Device Fabrication and Characterization

The GaN on silicon wafers were purchased from EpiGaN. All the metal-oxide sputtering targets were obtained from Kurt J. Lesker company.
The sensors were fabricated following steps similar to those of our prior paper [30]. Sensor fabrication was started with a standard RCA cleaning process, followed by a stepper lithography-assisted dry-etching process to form GaN nanowires of width 200–400 nm. To protect GaN nanowires during inductively coupled plasma etching, process patterned metals were used. With a standard electron beam evaporator, ohmic contacts were formed. Then, plasma-enhanced chemical vapor deposition was used to deposit the SiO2 layer, and also to protect nanowires, which are metal contacts that can be damaged by high-temperature processing or etching. Active area on GaN nanowire was created by a reactive ion-etching process to functionalize it with metal oxide. A thin layer of In2O3 was deposited by RF magnetron sputtering, and then noble metals (Au, Ag, Pt, Pd and Cu) were deposited to form a sensor, shown in Figure 1a. The In2O3 layer thickness was adjusted by controlling the sputtering deposition by varying the doping concentration from 1 × 1017 cm−3 to 1 × 1020 cm−3, depending on the sputtering and annealing conditions.
To complete the fabrication process, a rapid thermal annealing at 600–700 °C was performed to crystallize the noble metal-decorated In2O3. This also helps to improve the ohmic contacts. The microstructure and surface morphology of the fabricated metal-oxide/GaN sensors were analyzed using field-emission scanning electron microscopy (FESEM). Imaging was conducted with a Zeiss Ultra 60 FESEM to investigate the structural features of the synthesized samples. Elemental composition of the GaN nanowires was confirmed via energy-dispersive X-ray spectroscopy (EDS). The crystallographic structure and phase purity of the metal oxide films were assessed using X-ray diffraction, (XRD) performed on a Rigaku SmartLab system, Virgina, USA equipped with Cu-Kα radiation (Table 1). Additionally, the surface topography and roughness of the metal-oxide/GaN sensors were characterized using an Asylum Cypher high-resolution atomic force microscope (AFM). Device EDS analysis (Figure 1c) and a sensor SEM micrograph are shown in Figure 1b.
The fabricated device was placed in the designated gas stainless steel chamber for gas-sensing testing. Then, NO2 and compressed air were let in to the sensing apparatus while maintaining the net flow at 100 SCCM. The flow rate was controlled by a mass flow controller, and the device was biased with a 5 V constant voltage supply. The sensor was illuminated with a 365 nm, 470 mW/cm2 UV light power source. The photon energy of a 365 nm UV source is 3.4 eV which is equal to the GaN band gap and necessary for electron hole pair generation; UV power helps to operate the sensor at room temperature, as under dark conditions the sensor response was lower, due to insufficient active sites. The UV source helps to save power by operating the sensor at room temperature. All our experimental measurements were carried out at room temperature.
With NO2 present, change in the resistance leads to the change in the sensitivity S
S = R N O 2 R a i r R a i r
where Rair and RNO2 represent baseline resistance without NO2 and with NO2. A transient response of 10 ppm NO2 with various noble metals is shown in Figure 2a,b. Throughout the gas-sensing measurement, the device was allowed to regain the baseline (i.e., recover) after it was exposed to NO2 within 10 min of “buffer” time. NO2, upon exposure to the Au catalyst layer, dissociated as NO and Oxygen ions and then transferred to the In2O3 layer. This charge transfer changes the resistance of the sensor [24,25,26,27,28,29,30,31,32]. The relative change in resistance for Au/In2O3 is from 1.37 M to 1.9 M and bare In2O3 is from 800K to 883K in the presence of NO2 gas. In room temperature, the change in resistance or sensitivity is higher for Au when compared to other noble metals. Au/In2O3 sensors exhibited the highest response, at 38.9%, followed by bare In2O3 at 10%. Nanoparticle decoration increases the adsorption capabilities of NO2 on an In2O3 surface, due to a spill-over effect [24,25,26,27,28,29,30,31,32]. Increased sensitivity of Au/In2O3 is attributed to an increased area of active site, and increased electron transfer is due to gold decoration on In2O3. In air, a few free electrons in the conduction band of the sensor are captured by the oxygen, forming a depletion layer on the In2O3 surface. As Au/In2O3 has more free electrons, the depletion layer widens. Under NO2 exposure, negatively charged NO2 species are formed on the In2O3, and the depletion layer thickness increases. Adsorption of negatively charged NO2 by Au/In2O3 is stronger than by bare In2O3, since there are more free electrons in Au/In2O3 than in bare In2O3. The observed superiority of Au noble metal is further studied with the help of density functional theory.

3. DFT Calculation: Results and Discussion

All density functional theory calculations were performed using quantum espresso with generalized gradient approximation (GGA) and ultra-soft pseudo potential [33,34]. We designed our substrate using the In2O3 (111) surface, comprising 140 O atoms and 96 In atoms. We kept a vacuum layer of 15 Å to avoid adjacent layer interactions. All the atoms were relaxed until the force between them is less than 0.003 Ry/Bohr using the Broyden–Fletcher–Goldfarb– Shanno (BFGS) algorithm [35]. The Monkhorst–Pack k-point grid of 2 × 2 × 1 was used for geometrical optimization and a denser k-point of 6 × 6 × 1 was used to evaluate the electronic properties. Cutoff for wave function and charge was set to 55 Ry and 550 Ry.
To understand the adsorption behavior and site preference of noble metals on an In2O3 substrate, we first calculated the binding energy (Eb). This provides insight into the thermodynamic favorability of metal adsorption and helps identify the most stable configuration. The expression used to evaluate Eb is as follows [36]:
E b = E I n 2 O 3 + E M e t a l E I n 2 O 3 + M e t a l
where E I n 2 O 3 + M e t a l , E I n 2 O 3 , E M e t a l denote the total energies of the noble metal adsorbed In2O3, bare In2O3 and energy of non-interacting noble metals (Au, Ag, Pt, Pd and Cu). When greater, the absolute value of the binding energy is the most optimal site for noble metal. The resulting binding energies range from −3.5 eV to −4.0 eV across the different metals. These negative values indicate a spontaneous and favorable adsorption process and underscore a strong metal–support interaction. A higher absolute binding energy suggests a stronger bond and, thus, a more stable adsorption site.
Following the binding energy analysis, we explored the interaction between NO2 gas and the metal-decorated In2O3 surfaces by calculating the adsorption energy (Eads) [36]:
E a d s = E M e t a l I n 2 O 3 + N O 2 E M e t a l / I n 2 O 3 + E N O 2
where E M e t a l I n 2 O 3 + N O 2 , E M e t a l / I n 2 O 3 and E N O 2 are the total energies of the gas molecules adsorbed on different structures with noble metal, metal-attached structures and isolated gas molecules, respectively.
These energies are essential to assess the thermodynamics of gas adsorption and the interaction strength between the gas molecule and the surface.
The optimized configurations of noble metal-decorated In2O3 surfaces with NO2 adsorption are depicted in Figure 3, while the corresponding geometric parameters, such as bond lengths and bond angles, are listed in Table 2. The bonding patterns are distinctive for each noble metal:
  • Au and Pd each bond with one surface oxygen atom.
  • Pt forms a bond with one oxygen atom and one indium atom.
  • Ag binds to one indium atom.
  • Cu establishes bonds with two oxygen atoms and one indium atom.
The bond lengths and angles in the Au–NO2 and Pt–NO2 systems showed more significant deviation from the pristine In2O3 structure than in other cases. This observation suggests that NO2 undergoes molecular activation in the presence of these metals, thereby demonstrating their catalytic potential.
The computed NO2 adsorption energies for different systems are as follows:
  • Bare In2O3: −0.77 eV
  • Au/In2O3: −1.97 eV
  • Pt/In2O3: −0.99 eV
  • Pd/In2O3: −0.98 eV
  • Ag/In2O3: −0.94 eV
  • Cu/In2O3: −0.85 eV
Clearly, all noble metals enhance the adsorption energy compared to the bare substrate, with Au showing the most pronounced effect. The adsorption distance also decreases with metal decoration, indicating stronger interactions. These adsorption energies reflect that NO2 adsorption on Au is most favorable and likely involves chemisorption, while weaker values for Ag, Pt, Pd, and Cu indicate physisorption. The highly negative values confirm the spontaneous nature of the process.
To further elucidate the adsorption mechanism, we examined charge transfer between NO2 and the In2O3 surface using Löwdin population analysis. This analysis helps quantify the electronic redistribution during adsorption. As an electron acceptor, NO2 draws electrons from the surface, forming an electron depletion region that influences the material’s electrical properties.
The charge transferred to NO2 from the surface for each noble metal system is as follows:
  • Au/In2O3: 0.23 e
  • Pt/In2O3: 0.09 e
  • Pd/In2O3: 0.07 e
  • Ag/In2O3: 0.10 e
  • Cu/In2O3: 0.05 e
This transfer reduces the conduction band electron density, leading to an increase in surface resistance—especially relevant for resistive-type gas sensors. The largest transfer observed for Au implies that Au enhances the electron exchange and facilitates stronger NO2 adsorption. This effect widens the depletion layer and reduces carrier mobility, thus boosting sensor sensitivity. These trends align well with the calculated adsorption energies. The summary of adsorption energy, adsorption distance and charge transfer are all listed in Table 1.
To analyze the electronic structure modifications upon NO2 adsorption, we computed the density of states (DOS) before and after gas exposure. As illustrated in Figure 4, the DOS of Ag, Pt, Pd, and Cu systems showed minimal change, suggesting weak electronic interactions. In contrast, the Au-decorated system exhibited a significant DOS alteration near the Fermi level, affirming robust charge transfer and a stronger interaction with NO2 [26,27,28,29,30,31].
This finding is also consistent with adsorption energy and charge transfer calculation. We have also calculated charge density difference using [26]
ρ = ρ M e t a l I n 2 O 3 + N O 2 ρ M e t a l / I n 2 O 3 ρ N O 2
where ρ M e t a l I n 2 O 3 + N O 2 , ρ M e t a l / I n 2 O 3 and ρ N O 2 are the charge density of the gas adsorbed In2O3 with noble metal, and In2O3 with noble metal and gas molecule. The charge density difference in all the structures is displayed in Figure 5. Figure 5 reveals blue regions (electron depletion) and yellow regions (electron accumulation). The most intense charge redistribution is observed for Au/In2O3, again confirming that Au induces a higher degree of electronic reorganization, consistent with its catalytic role [22,23,24,25,26,27,28,29,30,31,32]. The findings support the hypothesis that Au serves as an excellent mediator for NO2. This result is also consistent with adsorption energy, charge transfer, and DOS plot.
A classic relationship between band gap (Eg) and conductivity is as follows [26]
σ α A e x p E g 2 k T
where σ is the conductivity, k is the Boltzmann constant (1.38 × 10−23 J/K), A is a fixed parameter and T is the temperature. As like in other papers, the band gap of In2O3 is less than the experimental value, since the DFT calculation underestimates the band gap [27].
The bare In2O3 exhibited a band gap of 0.8725 eV, which increased to 0.8899 eV after NO2 adsorption. This shift implies a decrease in conductivity, as expected when electrons are withdrawn by an oxidizing gas like NO2 [22,23,24,25,26,27,28,29,30,31,32].
Noble metal decoration reduced the band gap to 0.77 eV, enabling better electron mobility and enhancing NO2 capture. Post adsorption, the band gap of Au/In2O3 increased to 0.8155 eV, signifying that NO2 adsorption results in reduced conductivity—a typical signal in chemiresistive sensors (Table 3). Band gap variations for other metals were less prominent, correlating with weaker charge transfer.
The changes in Eg before and after NO2 adsorption are visualized in Figure 6. The largest ΔEg was observed for Au/In2O3, reinforcing Au’s pivotal role in promoting high NO2 sensitivity.
Through comprehensive DFT-based analysis, this study demonstrates the superior performance of Au-decorated In2O3 in NO2-sensing applications. The combination of high binding and adsorption energies, significant charge transfer, marked DOS variation, charge redistribution, and band gap modulation all converge to establish Au as the most effective catalyst among the noble metals studied.
This work provides theoretical validation for the enhanced sensing capability of Au/In2O3 composites, supporting their practical application in the development of next-generation, high-sensitivity NO2 gas sensors.

4. Conclusions

In this work, we have successfully fabricated a GaN nanowire-based NO2 gas sensor, functionalized with In2O3 and decorated by noble metals. The sensing properties of bare Indium oxide and that decorated with noble metals were investigated using density functional theory and experimental tests. Experimental tests showed that Au/In2O3 sensors exhibited the highest response of 38.9%, followed by bare In2O3 with 10% for 10 ppm NO2 at room temperature. The adsorption energy (Eads) calculated for bare/In2O3, Au/In2O3, Pt/In2O3, Pd/In2O3, Ag/In2O3, and Cu/In2O3 are −0.77, −1.97, −0.99, −0.98, −0.94, and −0.85, respectively. The charge transfers from substrate to gas were about 0.23, 0.09, 0.07, 0.1, and 0.05 e for Au/In2O3, Pt/In2O3, Pd/In2O3, Ag/In2O3, and Cu/In2O3, respectively. For Au/In2O3, Eg increased from 0.77 to 0.8155 eV, indicating a higher amount of NO2 adsorbed on it when compared to other noble metals. Both experimental and DFT results proved the enhancement of In2O3 with Au for NO2 gas sensors.

Author Contributions

Conceptualization, Q.L.; Methodology, P.R.; Validation, Q.L.; Formal analysis, P.R.; Investigation, P.R. and J.R.; Resources, A.M.; Data curation, P.R. and J.R.; Writing—original draft, P.R.; Writing—review & editing, P.R. and D.E.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NASA research grant on hybrid gas sensors and Virginia Microelectronics Consortium (VMEC) research grant.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original data used in this study can be obtained from the submitting author if requested.

Conflicts of Interest

Author Abhishek Motayed was employed by the company N5 Sensors (United States). The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Raju, P.; Li, Q. Semiconductor Materials and Devices for Gas Sensors. J. Electrochem. Soc. 2022, 169, 057518. [Google Scholar] [CrossRef]
  2. He, Y.; Jiao, M. A Mini-Review on Metal Oxide Semiconductor Gas Sensors for Carbon Monoxide Detection at Room Temperature. Chemosensors 2024, 12, 55. [Google Scholar] [CrossRef]
  3. Zhu, L.Y.; Ou, L.X.; Mao, L.W.; Wu, X.Y.; Liu, Y.P.; Lu, H.L. Advances in Noble Metal-Decorated Metal Oxide Nanomaterials for Chemiresistive Gas Sensors: Overview. Nano-Micro Lett. 2023, 15, 89. [Google Scholar] [CrossRef]
  4. Shah, S.; Hussain, S.; Din, S.T.U.; Shahid, A.; Amu-Darko, J.N.O.; Wang, M.; Tianyan, Y.; Liu, G.; Qiao, G. A review on In2O3 nanostructures for gas sensing applications. J. Environ. Chem. Eng. 2024, 12, 112538. [Google Scholar] [CrossRef]
  5. Xuan, J.; Zhao, G.; Sun, M.; Jia, F.; Wang, X.; Zhou, T.; Yin, G.; Liu, B. Low-temperature operating ZnO-based NO2 sensors: A review. RSC Adv. 2020, 10, 39786–39807. [Google Scholar] [CrossRef] [PubMed]
  6. Khort, A.; Haiduk, Y.; Taratyn, I.; Moskovskikh, D.; Podbolotov, K.; Usenka, A.; Lapchuk, N.; Pankov, V. High-performance selective NO2 gas sensor based on In2O3–graphene–Cu nanocomposites. Sci. Rep. 2023, 13, 7834. [Google Scholar] [CrossRef]
  7. Li, Y.; Wei, X.; Liu, Q.; Zang, D.; You, R. Visible Light-Activated Room Temperature NO2 Gas Sensing Based on the In2O3@ZnO Heterostructure with a Hollow Microtube Structure. ACS Sens. 2024, 9, 3741–3753. [Google Scholar] [CrossRef]
  8. Yang, Y.; Cui, J.; Luo, Z.; Luo, Z.; Sun, Y. Enhanced NO2 Gas Sensing Properties Based on Rb-Doped ZnO/In2O3 Heterojunctions at Room Temperature: A Combined DFT and Experimental Study. Sensors 2024, 24, 5311. [Google Scholar] [CrossRef]
  9. Luo, C.; Dong, Y.; Sun, Y.; Huang, Y.; Xue, L.; Ren, L.; Wang, D.; Qin, M.; Li, X. In2O3 and Au Nanoparticles Co-Modified WS2 Microsheets for Enhanced NO2 Gas Sensing at Room Temperature. IEEE Sens. J. 2024, 24, 25291–25298. [Google Scholar] [CrossRef]
  10. Tan, N.H.; Le, D.T.T.; Hoang, T.T.; Duy, N.M.; Tonezzer, M.; Xuan, C.T.; Van Duy, N.; Hoa, N.D. Metal-decorated indium oxide nanofibers used as nanosensor for triethylamine sensing towards seafood quality monitoring. Colloids Surf. A Physicochem. Eng. Asp. 2024, 703, 135268. [Google Scholar] [CrossRef]
  11. Wu, M.-J.; Wu, Y.-Z.; Tung, Y.-F.; Chang, T.-C.; Lee, C.-T.; Lee, H.-Y. Surface Modification and Decoration of Nano-Honeycombed ZnO Structure with Gold-Black Nanoparticles for High Responsivity Detection of NO2 Gas Sensors. IEEE Sens. J. 2024, 24, 38699–38707. [Google Scholar] [CrossRef]
  12. Ren, Y.; Xie, W.; Li, Y.; Ma, J.; Li, J.; Liu, Y.; Zou, Y.; Deng, Y. Noble Metal Nanoparticles Decorated Metal Oxide Semiconducting Nanowire Arrays Interwoven into 3D Mesoporous Superstructures for Low-Temperature Gas Sensing. ACS Cent. Sci. 2021, 7, 1885–1897. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, Y.; Li, S.; Xiao, S.; Du, K. In2O3 microtubes decorated with Ag nanoparticles for NO2 gas detection at room temperature. Vacuum 2022, 202, 111197. [Google Scholar] [CrossRef]
  14. Wang, H.; Yuan, Z.; Gao, H.; Zhang, R.; Meng, F. Ppb-Level N-Propanol Gas Sensor Based on Ce-Doped Flower-like In2O3 Compounded on rGO Nanosheets. IEEE Trans. Instrum. Meas. 2024, 73, 9517109. [Google Scholar] [CrossRef]
  15. Yuan, Z.; Zhang, H.; Li, J.; Meng, F.; Yang, Z.; Zhang, H. Ag Nanoparticle-Modified ZnO–In2O3 Nanocomposites for Low-Temperature Rapid Detection Hydrogen Gas Sensors. IEEE Trans. Instrum. Meas. 2023, 72, 9503712. [Google Scholar] [CrossRef]
  16. Guo, L.; Zeng, H.; Feng, Y.; Wang, Y.; Zhao, W.; Liu, X. Ar Plasma-Treated In2O3 Microtubes with Rich Oxygen Vacancies, High Surface Area, and Large Pores for Enhanced Formaldehyde Detection. IEEE Sens. J. 2023, 23, 22207–22216. [Google Scholar] [CrossRef]
  17. Yuan, Z.; Jin, J.; Zhu, H.; Zhang, H.; Meng, F. Ag-Modified Co3O4-In2O3 Nanohollow Microspheres for H2 Detection. IEEE Sens. J. 2024, 24, 25299–25307. [Google Scholar] [CrossRef]
  18. Cai, Z.; Park, J.; Jun, D.; Park, S. High-performance UV-activated room temperature NO2 sensors based on TiO2/In2O3 composite. CrystEngComm 2023, 25, 2546–2556. [Google Scholar] [CrossRef]
  19. Ilin, A.; Martyshov, M.; Forsh, E.; Forsh, P.; Rumyantseva, M.; Abakumov, A.; Gaskov, A.; Kashkarov, P. UV effect on NO2 sensing properties of nanocrystalline In2O3. Sens. Actuators B Chem. 2016, 231, 491–496. [Google Scholar] [CrossRef]
  20. Rambeloson, J.; Ioannou, D.E.; Raju, P.; Wang, X.; Motayed, A.; Yun, H.J.; Li, Q. Photoactivated In2O3-GaN Gas Sensors for Monitoring NO2 with High Sensitivity and Ultralow Operating Power at Room Temperature. Chemosensors 2022, 10, 405. [Google Scholar] [CrossRef]
  21. Liu, Y.; Li, J.; Hou, W.; Zhou, Q.; Zeng, W. Pristine and Ag decorated In2O3 (110): A gas-sensitive material to selective detect NO2 based on DFT study. J. Mater. Res. Technol. 2022, 18, 4236–4247. [Google Scholar] [CrossRef]
  22. Khalili, S.; Majidi, M.; Bahrami, M.; Roshanaei, M.; Madrakian, T.; Afkhami, A. A portable gas sensor based on In2O3@CuO P–N heterojunction connected via Wi-Fi to a smartphone for real-time carbon monoxide determination. Sci. Rep. 2024, 14, 13594. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, Y.; Han, S.; Wang, M.; Liu, S.; Liu, G.; Meng, X.; Xu, Z.; Wang, M.; Qiao, G. Electrospun Cu-doped In2O3 hollow nanofibers with enhanced H2S gas sensing performance. J. Adv. Ceram. 2022, 11, 427–442. [Google Scholar] [CrossRef]
  24. Nguyen, V.C.; Kim, K.; Kim, H. Performance Optimization of Nitrogen Dioxide Gas Sensor Based on Pd-AlGaN/GaN HEMTs by Gate Bias Modulation. Micromachines 2021, 12, 400. [Google Scholar] [CrossRef]
  25. Almaev, A.V.; Kopyev, V.V.; Novikov, V.A.; Chikiryaka, A.V.; Yakovlev, N.N.; Usseinov, A.B.; Karipbayev, Z.T.; Akilbekov, A.T.; Koishybayeva, Z.K.; Popov, A.I. ITO Thin Films for Low-Resistance Gas Sensors. Materials 2023, 16, 342. [Google Scholar] [CrossRef]
  26. Hwang, J.; Jung, H.; Shin, H.-S.; Kim, D.-S.; Kim, D.S.; Ju, B.-K.; Chun, M. The Effect of Noble Metals on CO Gas Sensing Properties of In2O3 Nanoparticles. Appl. Sci. 2021, 11, 4903. [Google Scholar] [CrossRef]
  27. Kumar, R.; Mamta; Kumari, R.; Singh, V.N. SnO2-based NO2 gas sensor with outstanding sensing performance at room temperature. Micromachines 2023, 14, 728. [Google Scholar] [CrossRef]
  28. Soydan, G.; Ergenc, A.F.; Alpas, A.T.; Solak, N. Development of an NO2 Gas Sensor Based on Laser-Induced Graphene Operating at Room Temperature. Sensors 2024, 24, 3217. [Google Scholar] [CrossRef]
  29. Jo, S.H.; Rhyu, H.; Park, C.; Kang, M.H.; Yoon, D.H.; Song, W.; Lee, S.S.; Lim, J.; Myung, S. Light-enhanced ultra-sensitive NO2 gas detection using In2O3/MXene-based flexible sensors. Appl. Surf. Sci. 2025, 704, 163448. [Google Scholar] [CrossRef]
  30. Khan, M.A.H.; Thomson, B.; Yu, J.; Debnath, R.; Motayed, A.; Rao, M.V. Scalable Metal Oxide Functionalized GaN Nanowire for Precise SO2 Detection. Sens. Actuators B Chem. 2020, 318, 128223. [Google Scholar] [CrossRef]
  31. Ueda, T.; Boehme, I.; Hyodo, T.; Shimizu, Y.; Weimar, U.; Barsan, N. Enhanced NO2-Sensing Properties of Au-Loaded Porous In2O3 Gas Sensors at Low Operating Temperatures. Chemosensors 2020, 8, 72. [Google Scholar] [CrossRef]
  32. Chen, C.; Zhang, Q.; Xie, G.; Yao, M.; Pan, H.; Du, H.; Tai, H.; Du, X.; Su, Y. Enhancing Visible Light-Activated NO2 Sensing Properties of Au NPs Decorated ZnO Nanorods by Localized Surface Plasmon Resonance and Oxygen Vacancies. Mater. Res. Express 2020, 7, 015924. [Google Scholar] [CrossRef]
  33. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  34. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  35. Møller, M.F. A scaled conjugate gradient algorithm for fast supervised learning. Neural Netw. 1993, 6, 525–533. [Google Scholar] [CrossRef]
  36. Opoku, F.; Govender, P.P. SF6 decomposed gas sensing performance of van der Waals layered cobalt oxyhydroxide: Insights from a computational study. J. Mol. Model. 2021, 27, 158. [Google Scholar] [CrossRef]
Figure 1. (a) Fabricated Au/In2O3/GaN nanowire field-effect transistor (FET) sensor, (b) top view of SEM images of GaN NW, (c) EDS spectrum of Au/In2O3.
Figure 1. (a) Fabricated Au/In2O3/GaN nanowire field-effect transistor (FET) sensor, (b) top view of SEM images of GaN NW, (c) EDS spectrum of Au/In2O3.
Chemosensors 13 00350 g001
Figure 2. (a,b) Transient response of 10 ppm NO2 at room temperature with various noble metals.
Figure 2. (a,b) Transient response of 10 ppm NO2 at room temperature with various noble metals.
Chemosensors 13 00350 g002
Figure 3. The most stable structures for NO2 adsorption on In2O3 (111) surface (a) bare, (b) Au decorated, (c) Pt decorated, (d) Pd decorated, (e) Ag decorated and (f) Cu decorated.
Figure 3. The most stable structures for NO2 adsorption on In2O3 (111) surface (a) bare, (b) Au decorated, (c) Pt decorated, (d) Pd decorated, (e) Ag decorated and (f) Cu decorated.
Chemosensors 13 00350 g003
Figure 4. Density of states diagram of bare and noble metal (a), Au (b), Pt (c), Pd (d), Ag (e) and Cu. (f) Decorated In2O3 before and after NO2 adsorption.
Figure 4. Density of states diagram of bare and noble metal (a), Au (b), Pt (c), Pd (d), Ag (e) and Cu. (f) Decorated In2O3 before and after NO2 adsorption.
Chemosensors 13 00350 g004aChemosensors 13 00350 g004b
Figure 5. Charge density difference in (a) bare, (b) Au, (c) Pt, (d) Pd, (e) Ag and (f) Cu decorated In2O3. The blue and yellow regions denote charge depletion and accumulation, respectively, and the isosurface value is set to 0.035 e Å−3.
Figure 5. Charge density difference in (a) bare, (b) Au, (c) Pt, (d) Pd, (e) Ag and (f) Cu decorated In2O3. The blue and yellow regions denote charge depletion and accumulation, respectively, and the isosurface value is set to 0.035 e Å−3.
Chemosensors 13 00350 g005aChemosensors 13 00350 g005b
Figure 6. Change in band gap Eg trend before and after NO2 adsorption.
Figure 6. Change in band gap Eg trend before and after NO2 adsorption.
Chemosensors 13 00350 g006
Table 1. Elemental mixing ratio.
Table 1. Elemental mixing ratio.
In2O3Au/In2O3
ElementsWeight %Atomic %Weight %Atomic %
N31.8469.9333.7371.73
Ga68.1630.0766.1528.26
Au000.110.02
Table 2. Adsorption parameter of bare and noble metal decorated In2O3, including adsorption energy and charge transfer along with gas molecules bond length, bond angle, and distance.
Table 2. Adsorption parameter of bare and noble metal decorated In2O3, including adsorption energy and charge transfer along with gas molecules bond length, bond angle, and distance.
ModelBond Length
(O-N) Å
Bond Angle
(O-N-O) Deg
Adsorption Energy (ev)Charge Transfer (e)Distance (Å)
In2O3–NO21.21, 1.28121.46−0.770.032.86
Au/In2O3–NO21.22, 1.3119.72−1.970.232.06
Ag/In2O3–NO21.22, 1.26124.34−0.940.12.17
Pt/In2O3–NO21.19, 1.42115.02−0.990.091.95
Pd/In2O3–NO21.22, 1.28124.07−0.980.071.99
Cu/In2O3–NO21.21, 1.29122.17−0.850.051.89
Table 3. Band gap value of bare and noble metal decorated In2O3, before and after gas adsorption.
Table 3. Band gap value of bare and noble metal decorated In2O3, before and after gas adsorption.
StructureBand Gap (eV)
Bare/In2O30.8725
Bare/In2O3–NO20.8899
Au/In2O30.77
Au/In2O3–NO20.8155
Ag/In2O30.8361
Ag/In2O3–NO20.8248
Pt/In2O30.8112
Pt/In2O3–NO20.8196
Pd/In2O30.8294
Pd/In2O3–NO20.8194
Cu/In2O30.8534
Cu/In2O3–NO20.8503
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

Raju, P.; Rambeloson, J.; Ioannou, D.E.; Motayed, A.; Li, Q. Noble Metal-Decorated In2O3 for NO2 Gas Sensor: An Experimental and DFT Study. Chemosensors 2025, 13, 350. https://doi.org/10.3390/chemosensors13090350

AMA Style

Raju P, Rambeloson J, Ioannou DE, Motayed A, Li Q. Noble Metal-Decorated In2O3 for NO2 Gas Sensor: An Experimental and DFT Study. Chemosensors. 2025; 13(9):350. https://doi.org/10.3390/chemosensors13090350

Chicago/Turabian Style

Raju, Parameswari, Jafetra Rambeloson, Dimitris E. Ioannou, Abhishek Motayed, and Qiliang Li. 2025. "Noble Metal-Decorated In2O3 for NO2 Gas Sensor: An Experimental and DFT Study" Chemosensors 13, no. 9: 350. https://doi.org/10.3390/chemosensors13090350

APA Style

Raju, P., Rambeloson, J., Ioannou, D. E., Motayed, A., & Li, Q. (2025). Noble Metal-Decorated In2O3 for NO2 Gas Sensor: An Experimental and DFT Study. Chemosensors, 13(9), 350. https://doi.org/10.3390/chemosensors13090350

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