Next Article in Journal
TFF-Net: A Feature Fusion Graph Neural Network-Based Vehicle Type Recognition Approach for Low-Light Conditions
Previous Article in Journal
Recent Advances in Multitone Microwave Frequency Measurement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mesoporous Bi2S3/Bi2O3 Heterostructure-Based Sensors for Sub-ppm NO2 Detection at Room Temperature

1
School of Electronics and Information Engineering, Nanjing University of Information Science and Technology, Nanjing 210044, China
2
School of Microelectronics and Control Engineering, Changzhou University, Changzhou 213164, China
3
School of Integrated Circuits, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Sensors 2025, 25(12), 3612; https://doi.org/10.3390/s25123612
Submission received: 2 May 2025 / Revised: 31 May 2025 / Accepted: 3 June 2025 / Published: 9 June 2025
(This article belongs to the Section Chemical Sensors)

Abstract

:
Novel Bi2S3/Bi2O3 hybrid materials with unique mesoporous structures were successfully synthesized via a facile in situ elevated-temperature thermal oxidation method using the Bi2S3 as a precursor in air. The as-prepared Bi2S3/Bi2O3 heterostructure-based sensor exhibits an excellent performance for detecting sub-ppm concentrations of NO2 at room temperature (RT). In the presence of 8 ppm NO2, the sensor registers a response of approximately 7.85, reflecting a 3.5-fold increase compared to the pristine Bi2S3-based sensor. The response time is 71 s, while the recovery time is 238 s, which are reduced by 32.4% and 24.2%, respectively, compared to the pristine Bi2S3-based sensor. The Bi2S3/Bi2O3 heterostructure-based sensor achieves an impressively low detection limit of 0.1 ppm for NO2, and the sensor has been demonstrated to possess superior signal repeatability, gas selectivity, and long-term stability. The optimal preparation conditions of the hybrid materials were explored, and the formation of mesoporous structure was analyzed. The obviously improved gas sensitivity of the Bi2S3/Bi2O3 heterostructure-based sensor can be assigned to the combined influence of electronic sensitization and its distinctive morphological structure. The potential gas-sensitive mechanisms were revealed by employing density functional theory (DFT). It was found that the formation of heterostructures could enhance the adsorption energies and increase the amount of electron transfer between NO2 molecules and the hybrid materials. Furthermore, the electron redistribution driven by orbital hybridization between O and Bi atoms improves the capacity of NO2 molecules to capture additional electrons from the Bi2S3/Bi2O3 heterostructures. The content of this work supplies an innovative design strategy for constructing NO2 sensor with high performance and low energy consumption at RT.

1. Introduction

The threat of air pollution to environmental protection and human health has increased substantially in recent decades, which is caused by the rapid development of industrial automation, leading to the dramatic enhancement in the consumption of traditional energy. As one of the most vital air pollution sources, nitrogen dioxide (NO2) is mainly produced by the exhaust emission of fuel vehicles and the combustion of fossil fuels in the industrial scopes [1,2,3]. NO2 generates ozone by participating in photochemical reactions, promotes the formation of fine particulate matter (PM2.5), and reacts with water vapor to produce acid rain, which has long-term corrosive effects on vegetation, soil, and water [4,5,6,7]. Moreover, NO2 has evident irritation and oxidation. As reported by the World Health Organization, frequent exposure to excessive concentrations of NO2 (more than 410 ppb per hour) can induce chronic bronchitis, asthma, pulmonary dysfunction, and even increase the risk of respiratory and cardiovascular diseases [8,9,10]. Therefore, the development of reliable NO2 sensors for detecting sub-ppm concentrations of NO2 is urgently needed in improving environmental protection and public health.
Elevated operating temperatures are necessary for traditional metal oxide semiconductors (MOSs) due to their insufficient surface activity and sluggish carrier mobility [11]. Long-term exposure to elevated temperatures will accelerate the deterioration of the performance of MOSs-based sensors and decrease the reliability and service lifetimes of the sensors. Furthermore, the supplementary energies will increase the complexity and energy consumption of the sensing system, which inhibits the practical application in special occasions such as flammable and explosive [12,13]. Compared to the MOSs, two-dimensional (2D) materials have garnered extensive attention from researchers and have gradually emerged as the research hotspot within the scope of gas sensors owing to their large specific surface area, adjustable band structures, high carrier mobility, and rich active sites [14,15,16,17]. Meanwhile, quasi-2D sulfides, including TiS3, SnS2, and WS2, have also been proven to possess excellent gas detection capabilities due to their layered structure and chemical properties similar to those of 2D materials [18,19,20]. As a typical 2D material, Bi2S3 has been widely investigated for gas detection because of its unique physicochemical, optical, and electrical properties [21,22,23,24,25]. For instance, Harke et al. proved that the Bi2S3 nanostructured film-based sensor fabricated by the solvothermal process exhibited outstanding selectivity toward 100 ppm NO2 in the interfering gases of NH3, CO2, C2H5OH, and H2S [21]. Russ et al. revealed that the synthesized Bi2S3 nanowire-based sensor showed reliable long-term stability to 3 ppm NO2 within 42 days, and the average relative error of the sensor was less than 10.5% [22]. Yang et al. demonstrated that the Bi2S3-based sensor with a nanorod structure has superior sensitivity for detecting 1 ppm NO2 at RT, making it a potential material for developing high-performance gas sensors [23]. Nonetheless, the inherent limitations of Bi2S3, involving the inadequate sensitivity and the poor signal repeatability, restrict its application in detecting sub-ppm concentrations of NO2 at RT.
Constructing hybrid materials is regarded as a productive strategy to address the limitations of single-component sensitive materials, which has been thoroughly explored to enhance the gas-sensitive characteristics of the sensors. The formed heterostructures could improve the transmission and accumulation of carriers, promote the adsorption and desorption of gas molecules, and increase the interaction efficiency between gas molecules and hybrid materials due to the electron sensitization caused by the built-in electric field [26,27]. Thus, compared to the single-component sensitive materials such as pristine Bi2S3, the hybrid materials could exhibit excellent sensitivity and signal repeatability [28,29,30]. Moreover, the accurate regulation of the morphological structure of hybrid materials could further increase the sensitivity of the sensor due to the enlargement of specific surface area, which is conducive to the adsorption of abundant gas molecules [31,32]. However, the improvement of gas-sensitive performance of the hybrid Bi2S3-based sensor with controllable morphological structure has rarely been reported, and the sensitization mechanisms still need further exploration.
In this work, we report a novel chemiresistive-type sensor capable of detecting sub-ppm levels of NO2 at RT, which employed the Bi2S3/Bi2O3 heterostructures with mesoporous as the sensitive materials. Due to the influence of electron sensitization, the transfer of carriers on the surface of Bi2S3/Bi2O3 hybrid materials and the adsorption/desorption process of NO2 molecules are substantially improved. Moreover, the unique mesoporous structure offers massive active sites and facilitates the deep diffusion behavior of NO2 molecules; thus, the gas-sensitive performance of the prepared sensor could be substantially enhanced.

2. Experimental Section

2.1. Fabrication of Mesoporous Bi2S3/Bi2O3 Hybrid Materials

The Bi2S3/Bi2O3 hybrid materials with mesopores were fabricated by one-step in situ elevated temperature thermal oxidation in air, as illustrated in Figure 1. Firstly, 200 mg Bi2S3 powders (Chengdu Alfa Metal Materials Co., Ltd., Chengdu, China, 99.999%) were evenly dispersed in an alumina ceramic boat inside a tube furnace (OTF-1200X-S, Hefei Kejing of MTI Corp, Hefei, China). Subsequently, the synthetic air (200 sccm/min) was introduced into the tube furnace, and the thermal oxidation was performed at temperatures ranging from 300~500 °C. The powders were heated to the desired temperature at the rate of 3 °C/min, and the temperature was maintained for 0~2 h, respectively. After the oxidation procedure, the powders were naturally cooled down to RT. Finally, the Bi2S3/Bi2O3 hybrid materials prepared at different temperatures and oxidation times were obtained.

2.2. Material Characterizations

The thermostability of the Bi2S3 powders was characterized using a thermogravimetric analyzer (TGA, STA200, Hitachi Ltd., Tokyo, Japan) at a thermal ramp rate of 5 °C/min in air. The crystallographic information of the synthesized materials was identified utilizing X-ray diffraction (XRD, XRD-6000, PANalytical B.V., Almelo, Netherlands) with Cu Kα radiation. The morphological structures of the hybrid materials were recorded by scanning electron microscopy (SEM, Regulus8100, Hitachi Ltd., Tokyo, Japan) and transmission electron microscopy (TEM, Talos F200X G2, FEI Company, Hillsboro, OR, USA). The elemental states were investigated by X-ray photoelectron spectroscopy (XPS, Escalab250Xi, Thermo Fisher Scientific, Waltham, MA, USA). The pore diameter distributions and specific surface area of the hybrid materials were verified using Brunauer–Emmett–Teller (BET, NOVA2000e, Quantachrome Instruments, Boynton Beach, FL, USA). The elemental defects in the hybrid materials were measured through an electron paramagnetic resonance spectrometer (EPR, JES-FA200, JEOL Ltd., Tokyo, Japan). The work functions of the hybrid materials were confirmed by Kelvin probe force microscopy (KPFM, SKP5050, KP Technology, Wick, UK).

2.3. Sensor Preparation and Measurements

In a typical synthesis procedure, 5 mg of the as-prepared Bi2S3/Bi2O3 hybrid materials were placed into an agate mortar. Subsequently, a small amount of deionized water (~0.1 mL) was added, and the mixture was thoroughly ground for 10 min to form a printable paste. The paste was then uniformly dispensed onto the alumina substrate precoated with Au interdigitated electrodes (Changchun Beirun Electronics Technology Co., Ltd., Changchun, China, 10 × 5 × 0.635 mm, 8 pairs of electrodes, finger spacing of 200 μm, finger width of 200 μm) using a pipette gun (DLAB Scientific Co., Ltd., Beijing, China), achieving a thickness of approximately 3 μm.
The Bi2S3/Bi2O3 heterostructure-based sensors were developed by transferring the interdigital electrodes into a vacuum oven (DZF-6020, Hefei Kejing of MTI Corp, Hefei, China) and drying them at 60 °C for 12 h.
The measurement platform consisted of a digital signal detection system (Agilent 34410A, Santa Clara, CA, USA) and a dynamic gas allocation system (Beijing Sevenstar Huachuang Meter Co., Ltd., Beijing, China). The digital signal detection system used the Agilent 34410A electrometer to register the resistance of the sensors. The dynamic gas allocation system employed the Mass Flow Controllers (MFCs) to dilute the purchased 10 ppm NO2 to different concentrations by accurately controlling the flow rate of synthetic air. The humidity in the reaction chamber was regulated by humidifying synthetic air with different saturated salt solutions. The sensor response ( S ) was determined by the ratio of the saturated resistance in the measured gases ( R g a s ) to that in the background gas ( R a i r , synthetic air: O2/N2 = 1/8), as expressed by the following formulas:
S = R g a s / R a i r   ( for   oxidizing   gases )
S = R a i r / R g a s   ( for   reducing   gases )
The response/recovery time was calculated by measuring when the sensor resistance reached/recovered 90% of the saturated value.

2.4. DFT Computation

Theoretical computations were conducted using the Vienna Ab-initio Simulation Package (VASP.5.4.4) to analyze the adsorption energy, electron transfer amount, and DOS for NO2 molecules interacting with Bi2S3/Bi2O3 hybrid materials. The Perdew-Burke-Ernzerhof (PBE) of generalized gradient approximation (GGA) was employed for the exchange-correlation pseudopotential [33,34,35]. The influence of van der Waals interactions was corrected employing the DFT-D3 method [36,37]. The Brillouin zones were sampled using 2 × 2 × 1 and 9 × 9 × 1 k-points for structure optimization and DOS calculation, respectively. A force convergence criterion of 0.05 eV/Å and a plane-wave cutoff energy of 450 eV were utilized. The energy convergence standard of structural self-consistency was set to 1 × 10−5 eV/atom. Additionally, a 15 Å vacuum layer was included to prevent interactions between adjacent units. The adsorption energy ( E a d s ) was described as given below.
E a d s = E s u b + g a s E s u b E g a s
where E s u b + g a s , E s u b , and E g a s represent the total energy of the gas adsorption configuration, the energy of the Bi2S3/Bi2O3 heterostructure, and the energy of the NO2 molecule, respectively. The negative value of E a d s indicates that the interaction is exothermic, and the greater the | E a d s |, the stronger the stability of NO2 adsorption.

3. Results and Discussion

3.1. Microstructure of Bi2S3/Bi2O3 Hybrid Materials

The thermal stability of the pristine Bi2S3 was investigated by TGA from 25 °C to 800 °C with a ramping rate of 10 °C/min in air. As shown in Figure 2, the pristine Bi2S3 was relatively stable below 300 °C with only about 0.07% weight loss, which could be assigned to the evaporation of pre-adsorbed water molecules. In the temperature range of 300 °C to 650 °C, the weight of Bi2S3 obviously decreased by approximately 9.06% (theoretical weight loss was calculated to be 9.37%), indicating that the redox reaction occurred between Bi2S3 and O2. The weight loss is caused by the direct volatilization of an abundance of SO2 produced during the following reaction:
2 B i 2 S 3 + 9 O 2 = 2 B i 2 O 3 + 6 S O 2
There was no obvious weight loss between 650 °C and 800 °C, illustrating that the Bi2S3 had been completely oxidized to Bi2O3 in this temperature range. Therefore, according to the analysis results of the thermal stability of pristine Bi2S3, the Bi2S3/Bi2O3 hybrid materials were fabricated utilizing 300 °C, 350 °C, 400 °C, 450 °C, and 500 °C as oxidation temperatures and controlling the oxidation times at 0 h, 0.5 h, 1 h, and 2 h, respectively.
The crystal structure and crystal-phase information of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 were characterized by XRD. The Bi2S3/Bi2O3 hybrid materials were prepared at 400 °C for 1 h, and the Bi2O3 was obtained at 700 °C for 1 h under the same fabrication method. In Figure 3a, the diffraction peaks located at 15.74°, 24.84°, 25.34°, 35.71°, and 48.36° are indexed to the (020), (130), (310), (240), and (060) lattice planes of Bi2S3 (JCPDS No. 17-0320), respectively, illustrating that the pristine Bi2S3 possesses an orthogonal phase [38,39]. Similarly, the diffraction peaks at 25.71°, 26.97°, 27.36°, 32.93°, and 33.31° are assigned to the (002), (111), (120), (121), and (200) crystal planes of Bi2O3 (JCPDS No. 71-0465) with a monoclinic structure [40,41]. As shown in Figure 3b, the three weak peaks of Bi2S3/Bi2O3 at 2θ of 25.84°, 27.05°, and 28.09°, which could be attributed to the (002), (111), and (012) planes of Bi2O3, demonstrate that the Bi2O3 nanoparticles were formed during the elevated temperature thermal oxidation process. It should be noted that, compared to the Bi2O3, the positions of the three diffraction peaks are slightly shifted by approximately 0.1° towards the large-angle direction, implying that the introduction of the O element causes the lattice distortion, and the lattice spacing of Bi2S3 becomes smaller [42,43].
The morphological evolution procedure of the as-synthesized materials prepared at various temperatures was investigated by SEM and TEM, as described in Figure 4. In Figure 4a, the surface of the pristine Bi2S3 microflake without elevated-temperature thermal oxidation treatment was smooth, showing the typical layered structure characteristic of 2D materials, and the length was in the range of 4~6 μm. As the temperature increased, the surface of the materials became rough gradually. Notably, compared to the pristine Bi2S3, the surface of Bi2S3/Bi2O3 hybrid materials prepared at 400 °C for 1 h exhibits abundant mesoporous structures, as shown in Figure 4b. This structural feature is beneficial to enhance the specific surface area of the hybrid materials and promote the deep diffusion of gas molecules, which is required by chemiresistive-type semiconductor gas sensors. Figure 4c displays the Bi2O3 obtained at 700 °C for 1 h. It could be observed that the previously formed mesoporous structures had disappeared, and the surface of Bi2O3 had changed to be smoother and denser. This could be due to the decrease in surface energy with the progress of redox reaction [44]. In accordance with the principle of surface energy minimization, the redistribution of surface atoms reduces the number of surface defects and dangling bonds with higher energy, thus altering the morphology of Bi2O3. Figure 4d illustrates the high-resolution TEM image of the Bi2S3/Bi2O3 hybrid materials fabricated at 400 °C for 1 h. It can be seen that a Bi2O3 nanoparticle with a diameter of about 10 nm was formed at the edge of the Bi2S3 microflake. The 0.312 nm lattice spacing corresponds to the (211) crystalline plane of Bi2S3, while the 0.269 nm lattice spacing is consistent with the (200) crystalline plane of Bi2O3, implying the development of a Bi2S3/Bi2O3 heterostructure. The elemental mapping images of the Bi2S3/Bi2O3 hybrid materials are presented in Figure 4e–i. The elements of Bi (red), S (yellow), and O (green) are homogeneously and continuously distributed, demonstrating that the Bi2O3 nanoparticles are uniformly spread on the surface of Bi2S3 microflakes without agglomeration. This makes the active sites completely exposed, which improves the sensitivity and response speed of the Bi2S3/Bi2O3 heterostructure-based sensors.
In Figure 5, the N2 adsorption/desorption isotherms and pore size distribution were characterized to determine the specific surface area and porosity of the hybrid materials synthesized with the introduction of Bi2O3 nanoparticles. The curves of the pristine Bi2S3 and the Bi2S3/Bi2O3 hybrid materials prepared at 400 °C for 1 h both exhibit the typical type-IV isotherms. In Figure 5a, in the range of relative pressure ( P / P 0 ) from 0.0 to 1.0, no hysteresis loop could be observed in the pristine Bi2S3. In contrast, a hysteresis loop appeared obviously within the range of 0.4~1.0 ( P / P 0 ) for the hybrid materials, which could be assigned to the unique mesoporous structure on the surface of Bi2S3/Bi2O3, as illustrated in Figure 5b. The pore size distribution curve of Bi2S3/Bi2O3 demonstrates that the average pore size is about 17.4 nm. Furthermore, the specific surface areas of pristine Bi2S3 and the Bi2S3/Bi2O3 hybrid materials were recorded to be 3.31 m2g−1 and 26.92 m2g−1, respectively. It could be found that the mesoporous structure was formed on the surface of hybrid materials through in situ elevated temperature thermal oxidation, and the specific surface area was significantly improved by approximately 8.1-fold. This desired structure supplies additional adsorption sites, which enhances the ability to attract gas molecules.
The chemical states of surface elements of pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 were identified by XPS measurement. The survey spectrums (see Figure S1) demonstrate that Bi, S, and O elements coexist in the Bi2S3/Bi2O3 hybrid materials prepared at 400 °C for 1 h, consistent with the elemental mapping results. After elevated-temperature thermal oxidation at 700 °C for 1 h, no S 2s elemental peak could be detected in Bi2O3, illustrating that the pristine Bi2S3 had been completely oxidized to Bi2O3. Figure 6a describes the XPS spectra of Bi 4f and S 2p in the synthesized materials. In the pristine Bi2S3, the elemental peaks at 158.59 eV and 163.89 eV correspond to Bi 4f7/2 and Bi 4f5/2, respectively [28]. Similarly, in Bi2O3, these peaks are found at 159.51 eV and 164.81 eV [45]. It reveals that the valence of bismuth atoms is Bi3+ in both pristine Bi2S3 and Bi2O3. In Bi2S3/Bi2O3 hybrid materials, the elemental peaks of Bi 4f are split into two groups, and the peaks centered at 158.45 eV and 163.76 eV correspond to Bi-S bonds, while those at 159.73 eV and 165.03 eV are considered to be Bi-O bonds [46]. It could be observed that, compared to the pristine Bi2S3 and Bi2O3, the peaks of Bi-S bonds shifted upward by about 0.14 eV toward the lower binding energy, while the peaks of Bi-O bonds shifted downward by approximately 0.22 eV in the opposite direction. This could be assigned to the transfer and accumulation of carriers at the heterostructure interfaces due to the built-in electric field [47,48]. Moreover, the elemental peaks of S 2p occurred at 161.25 eV and 162.52 eV and are assigned to the S 2p3/2 and S 2p1/2, implying that the valence of sulfur atoms is S2− in both pristine Bi2S3 and Bi2S3/Bi2O3 hybrid materials [49].
Figure 6b shows the O 1s spectrums of the as-synthesized materials. The elemental peak at 531.54 eV in the pristine Bi2S3 corresponds to the chemisorbed oxygen ions (OC), and the species should be O 2 at RT [50,51]. The O 1s spectrum in the Bi2S3/Bi2O3 hybrid materials displays three peaks at 531.22 eV, 530.76 eV, and 529.89 eV, which could be assigned to the OC, oxygen vacancies (OV), and lattice oxygen ions (OL) [52,53]. The binding energies of OC, OV, and OL in Bi2O3 shift to 531.43 eV, 530.71 eV, and 529.65 eV, respectively, as the oxidation temperature rises [54,55]. By calculating the integral area of each peak in the O 1s spectrum, the relative concentrations of OC, OV, and OL in the prepared materials can be determined and summarized (see Table S1). It could be found that compared to the pristine Bi2S3, the content of OC in Bi2S3/Bi2O3 hybrid materials is significantly enhanced by approximately 3.05-fold. However, the improvement of the content of OC in Bi2O3 is only 1.13-fold. This could be due to the formation of heterostructures promoting the adsorption of oxygen molecules on the surface of hybrid materials. In addition, the proportion of OV in Bi2S3/Bi2O3 and Bi2O3 was estimated to be about 25.27% and 11.29%, respectively. These oxygen vacancies act as capture centers of O2 molecules, providing additional active sites for the gas-sensitive process.
EPR was employed to investigate the oxygen vacancies in the Bi2S3/Bi2O3 hybrid materials prepared at 400 °C for 1 h and in the Bi2O3 obtained at 700 °C for 1 h. In Figure 7, the microwave frequency was set to 9.45 GHz, and the EPR signal at g = 2.00037 confirmed oxygen vacancies in both Bi2S3/Bi2O3 and Bi2O3 [56]. Compared to the Bi2O3, the EPS signal intensity of Bi2S3/Bi2O3 is higher, demonstrating that the unpaired electrons in the hybrid materials present more pronounced interaction with the external magnetic field. It reveals that the oxygen vacancy content in Bi2S3/Bi2O3 hybrid materials exceeds that in Bi2O3, consistent with XPS analysis.

3.2. Gas-Sensitive Properties

To determine the gas-sensitive performance of the sensors under various preparation conditions, the response curves of the pristine Bi2S3-based sensor and the Bi2S3/Bi2O3 heterostructure-based sensors fabricated at different thermal oxidation temperatures (from 300 °C to 500 °C) and oxidation times (from 0 h to 2 h) to 8 ppm NO2 are shown in Figure 8a,b. In Figure 8a (with oxidation time fixed at 1 h), the response curves of all synthesized sensors exhibited an upward trend when NO2 was introduced into the reaction chamber, and the curves showed the opposite behavior after NO2 was removed. Since NO2 is a typical oxidizing gas species, combined with Formula (1), it could be found that the prepared Bi2S3/Bi2O3 hybrid materials take electrons as the majority carriers, indicating the characteristics of N-type semiconductors. Furthermore, with the oxidation temperature increased from 300 °C to 350 °C, the responses rose from 3.08 to 4.52, which illustrated that the sensitivity of the sensors to NO2 had been enhanced gradually. When the oxidation temperature reached 400 °C, the Bi2S3/Bi2O3 heterostructure-based sensor exhibited a response of approximately 7.85 to 8 ppm NO2, achieving the maximum response value. This response surpassed that of the pristine Bi2S3-based sensor by 3.5-fold. However, the sensing characteristics of the sensors started to deteriorate as the oxidation temperature rose from 450 °C to 500 °C, with the responses decreasing to 6.29 and 5.31, respectively. The decline of gas-sensitive performance could be related to the reduction in heterostructure content. Thus, the optimal oxidation temperature was determined to be 400 °C. Notably, the resistance of the Bi2O3-based sensor prepared at 700 °C exceeded the measurement range of the digital multimeter at RT (~1200 MΩ), rendering it impossible to obtain its response curve to NO2. The I-V curves (see Figure S2) of the pristine Bi2S3, the Bi2S3/Bi2O3 heterostructure, and the Bi2O3-based sensors present a linear relationship (Ohmic contact) at RT, indicating adequate contact between the sensitive materials and the interdigital electrodes.
Figure 8b depicts the impact of various oxidation times on the characteristics of the prepared sensors at a constant oxidation temperature of 400 °C. It could be observed that the response of Bi2S3/Bi2O3 heterostructure-based sensors increased initially and then decreased as the oxidation time rose from 0 h to 2 h. The optimal oxidation time is distinctly identified as 1 h. In Figure 8c,d, the response and recovery times of the Bi2S3/Bi2O3 heterostructure-based sensor and the pristine Bi2S3-based sensor have been calculated. In comparison to the pristine Bi2S3-based sensor, the Bi2S3/Bi2O3 heterostructure-based sensor produced at 400 °C for 1 h exhibited response and recovery times of 71 s and 238 s, which were reduced by approximately 32.4% and 24.2%, respectively. Obviously, the construction of the heterostructure could significantly enhance the sensitivity and the response/recovery speed of the pristine Bi2S3-based sensor.
Figure 9a displays the resistance curve of the Bi2S3/Bi2O3 heterostructure-based sensor obtained utilizing the optimal oxidation temperature (400 °C) and oxidation time (1 h) for detecting NO2 at concentrations ranging from 0.1 ppm to 8 ppm. As the concentration of NO2 increased continuously, the resistance variation scope of the sensor expanded gradually. The initial resistance was approximately 36 MΩ, which could be recovered completely after the termination of NO2; no drift of the baseline resistance was observed. The response fitting curve in the illustration (inset) reveals that the response of the Bi2S3/Bi2O3 heterostructure-based sensor had a linear correlation with the change of NO2 concentration, and the sensitivity of the sensor was calculated to be 0.81418 ± 0.01238 ppm−1. Figure 9b shows the signal repeatability verification of the Bi2S3/Bi2O3 heterostructure-based sensor. The results of 10 cyclic measurements of 8 ppm NO2 confirmed that the response deviation of the synthesized sensor was less than 3.4%, and it exhibited excellent signal repeatability. As we know, environmental humidity has an outstanding influence on the performance of chemiresistive-type gas sensors. Thus, the sensing properties of the Bi2S3/Bi2O3 heterostructure-based sensor were examined at varying humidities, as depicted in Figure 9c. The response and resistance of the Bi2S3/Bi2O3 heterostructure-based sensor decreased gradually as the ambient humidity rose (corresponding curves see Figures S3 and S4). At 85% ± 2% RH, the response of the sensor was only 39.14% (~3.08) of that in the low-humidity environment (≤10% ± 2% RH). This could be assigned to water molecules in high-humidity environments occupying the active sites on the surface of Bi2S3/Bi2O3 hybrid materials, which weakens the contact and reaction opportunities between gas molecules and hybrid materials, resulting in the deterioration of the surface activity of the sensor. Meanwhile, the reduction in chemisorbed oxygen ions decreases the surface potential of Bi2S3/Bi2O3 hybrid materials. This increase in conductivity covered the resistance change caused by NO2 molecules, thus decreasing the response of the sensor.
Figure 9d describes the response of the Bi2S3/Bi2O3 heterostructure-based sensor to 8 ppm NO2 and 100 ppm concentrations of various interfering gases, including ammonia, formaldehyde, methylbenzene, benzene, acetone, methanol, and ethanol. Clearly, compared to the interfering gas species, the as-prepared sensor exhibited a significantly stronger interaction with low concentrations of NO2, proving its superior gas selectivity. The long-term stability of the Bi2S3/Bi2O3 heterostructure-based sensor was evaluated as described in Figure 9e. During the entire verification process, the sensor was assessed every two weeks, and it was stored in the natural environment for the rest of the time. The slight decrease in sensor response in the initial stage could be related to the aging of Bi2S3/Bi2O3 hybrid materials. This procedure could partially passivate the dangling bonds on the surface of the hybrid materials, thus reducing the number of defect states. In spite of this, after 11 weeks of measurement, the response deviation of the sensor to 8 ppm NO2 remained less than 6.6%, and the responses tended to stabilize gradually, demonstrating its outstanding long-term stability.
The endurance of the sensor is a vital criterion for assessing its stable operation in complex environments and plays a crucial role in ensuring the accuracy and reliability of gas detection. According to the above experimental results, compared to its anti-humidity interference ability, the Bi2S3/Bi2O3 heterostructure-based sensor fabricated in this paper exhibits remarkable signal repeatability, gas selectivity, and long-term stability. This makes it comply with the practical application requirements in various harsh scenarios under moderate and low-humidity conditions. Thus, the developed Bi2S3/Bi2O3 hybrid materials are expected to be a high-performance candidate for the detection of trace NO2.
The gas-sensitive characteristics of the Bi2S3/Bi2O3 heterostructure-based sensor developed in this work for detecting NO2 at RT were compared to those of Bi2S3-based sensors reported in the literature, as summarized in Table 1. It can be concluded that the specific surface area of the synthesized Bi2S3/Bi2O3 hybrid materials was increased by constructing the heterostructure and combining it with the unique mesoporous structure. This provided more adsorption sites for NO2 molecules, significantly improving the detection performance of Bi2S3-based sensors. Therefore, the as-prepared sensor possesses superior sensitivity, rapid response and recovery speed, and a lower detection limit for sensing low-concentration NO2.

3.3. Sensitization Mechanisms

3.3.1. Roles of Bi2S3/Bi2O3 Heterostructure

The enhanced gas-sensitive characteristics of the Bi2S3/Bi2O3 heterostructure-based sensor to NO2 are primarily assigned to the effect of electron sensitization. The contact potential differences (CPDs) of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 were measured using the single-point mode of KPFM. A gold-plated cantilever with a diameter of 2 mm was adopted as the probe tip. The vibration frequency, data acquisition rate, and bias voltage were set at 79 Hz, 13,500 Hz, and ± 7 V, respectively. The work function of the tip and the measured CPDs of the materials were utilized to calculate their corresponding work functions, which were found to be 4.83 eV, 4.89 eV, and 5.04 eV, respectively (see Figure S5). Since both Bi2S3 and Bi2O3 are N-type semiconductors, an N-N homojunction with the type-II staggered bandgap is formed as shown in Figure 10a [49,59]. The main carrier electrons are transferred from the conduction band of Bi2S3 with a relatively narrow bandgap to that of Bi2O3, which is caused by the concentration gradient. Meanwhile, the holes migrate to the valence band of Bi2S3 in the opposite direction until the Fermi level reaches dynamic equilibrium. The accumulation of electrons and holes on each side of the heterostructure generates the built-in electric field, thus achieving effective separation of the carriers. The increased electron content at the interface of the heterostructure facilitates greater oxygen molecule ( O 2 ( a d s ) ) adsorption, increasing the content of chemisorbed oxygen ions ( O 2 ( a d s ) ) on the surface of the Bi2S3/Bi2O3 hybrid materials. The XPS analysis findings support this observation. Subsequently, the introduced NO2 molecules directly capture surface electrons of hybrid materials through physisorption [60]. The redox interaction between NO2 molecule and O 2 ( a d s ) generates N O 2 ( a d s ) and O ( a d s ) ions, which can further trap additional electrons through chemisorption [61,62]. According to Equations (7) and (8), the combination of both sensing processes leads to the narrowing of the conductive channels in the hybrid materials, increases the barrier height, and aggravates the degree of energy band bending. As a result, a broader range of resistance changes is produced, and the sensitivity of the Bi2S3/Bi2O3 heterostructure-based sensor is improved. Figure 10b describes the schematic diagram of gas-sensitive procedure, and the equations of related reactions are as follows:
O 2 ( g a s ) O 2 ( a d s )
O 2 ( a d s ) + e O 2 ( a d s )
N O 2 ( g a s ) + e N O 2 ( a d s )
N O 2 ( g a s ) + O 2 ( a d s ) + 2 e N O 2 ( a d s ) + 2 O ( a d s )

3.3.2. Roles of Morphological Structure

The particular morphological structure also contributes significantly to improving the gas-sensitive properties of the Bi2S3/Bi2O3 heterostructure-based sensor. In the SEM image (see Figure 4b), the surface of the Bi2S3/Bi2O3 hybrid materials exhibits the unique mesoporous structure, which could be attributed to two aspects. Since the radius of the O atom is smaller than that of the S atom, the primitive cell volume of Bi2O3 (~81.07 Å3) formed after thermal oxidation is only about 16.05% of that of Bi2S3 (~505.06 Å3), resulting in the lattice distortion and the local collapse of hybrid materials [63,64]. On the other hand, the Bi2O3 nanoparticles were found to be rich in oxygen vacancies, and the proportion of oxygen vacancies was estimated to be approximately 25.27% by EPR and XPS measurements. This is also the decisive factor for the formation of the mesoporous structure. Owing to this morphological characteristic, the specific surface area of the Bi2S3/Bi2O3 hybrid materials increased by approximately 8.1-fold. Moreover, the mesoporous structure facilitates the thorough permeation of O2 and NO2 molecules, while the oxygen vacancies supply more active sites for gas molecule attachment. Compared to the pristine Bi2S3, the content of O 2 ( a d s ) on the surface of hybrid materials was increased by about 3.05-fold. Thus, the effectiveness of the Bi2S3/Bi2O3 heterostructure-based sensor in detecting low-concentration NO2 was significantly improved.

3.3.3. DFT Theoretical Calculation

To deeply reveal the sensitization mechanisms, the gas-sensitive processes and electron transfer characteristics of pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 in interaction with NO2 molecules were investigated utilizing DFT calculations. The adsorption sites of NO2 molecules on various structures were explored, and the optimized structures for NO2 adsorption are shown in Figure 11a–c. In Figure 11b, it could be observed that a new Bi-O bond with the length of 2.41 Å was formed between the NO2 molecule and the Bi2S3/Bi2O3 heterostructure, which constitutes a stable adsorption structure. The corresponding binding energies of the adsorption sites are shown in Table S2. The adsorption energy was calculated to be approximately −1.26 eV according to Equation (3) mentioned above, indicating that the Bi2S3/Bi2O3 heterostructure has strong adsorption capacity for NO2. In contrast, the interaction of NO2 molecules with the surfaces of Bi2S3 and Bi2O3 is primarily governed by relatively weak van der Waals forces, exhibiting typical characteristics of physisorption. The corresponding adsorption energies are approximately −0.21 eV and −0.39 eV, while the distances between NO2 molecules and the adsorption sites are measured to be 2.97 Å and 2.81 Å, respectively. Moreover, the electron transfer properties were analyzed, and the related charge density differences are shown in Figure 11d–f. The yellow electron cloud represented electron aggregation, and the cyan area reflected electron depletion. The Bader charge calculations confirm that the electron transfer amounts in the adsorption structures of Bi2S3, Bi2S3/Bi2O3 heterostructure, and Bi2O3 are −0.17 e, −0.67 e, and −0.08 e, respectively. Evidently, NO2 molecules capture numerous electrons from the heterostructure, resulting in the greatest change in resistance. This illustrates that the Bi2S3/Bi2O3 heterostructure-based sensor possesses the optimal sensitivity to NO2, as demonstrated by the experimental results. The adsorption energies, electron transfer amounts, and the distances between NO2 molecules and the adsorption sites are summarized in Table 2.
Figure 12 displays the DOS for the adsorption structures of Bi2S3, the Bi2S3/Bi2O3 heterostructure, and Bi2O3 in the presence of NO2 molecules. It can be found that all the structures retain the semiconductor properties with bandgaps after the adsorption of NO2 molecules. The total density of states (TDOS) of the three different structures reflects that a new characteristic peak occurs around the Fermi level (0 eV) after NO2 molecule adsorption. In particular, compared to the adsorption structures of Bi2S3 and Bi2O3, the TDOS of the Bi2S3/Bi2O3 heterostructure exhibits the most pronounced peak intensity in the range of −0.1 to −0.4 eV (the position of the NO2 characteristic peak) as depicted in Figure 12c, suggesting that the number of electron orbits in this region is the largest. It implies that the heterostructure could provide more electrons to participate in the interaction with NO2 molecules. Meanwhile, the comparison of the projected density of states (PDOS) describes that the overlapping areas of the p orbits of O atoms in NO2 and the p orbits of Bi atoms on the Bi2S3/Bi2O3 heterostructure surface are enlarged in both the conduction band (from 1.9 eV to 5.0 eV) and the valence band (from −0.7 eV to −3.7 eV) as shown in Figure 12d, indicating a strong orbital hybridization effect. The enhanced electron delocalization due to orbital convergence redistributes electrons on the Bi2S3/Bi2O3 heterostructure surface. As a conclusion, the NO2 molecules could capture more electrons from the hybrid materials, thus boosting the gas-sensitive properties of the Bi2S3/Bi2O3 heterostructure-based sensor.

4. Conclusions

The Bi2S3/Bi2O3 hybrid materials with mesoporous structure were prepared utilizing a facile in situ elevated temperature thermal oxidation method. The influence of various oxidation temperatures (300~500 °C) and oxidation times (0~2 h) on the gas sensitivity of the as-fabricated sensor was explored, and the formation mechanism of the mesoporous structure was analyzed in detail. The Bi2S3/Bi2O3 heterostructure-based sensor synthesized at 400 °C for 1 h exhibits a response of around 7.85 to 8 ppm NO2 at RT, reflecting a 3.5-fold boost compared to the pristine Bi2S3-based sensor. The response and recovery times are 71 s and 238 s, and the detection limit for low-concentration NO2 can reach 0.1 ppm. The sensor was also proven to possess exceptional signal repeatability, gas selectivity, and long-term stability. The response deviation of the sensor remained below 6.6% after 11 weeks of measurement. The significant improvement of the Bi2S3/Bi2O3 heterostructure-based sensor can be assigned to electronic sensitization and distinct morphological structure. Furthermore, DFT was employed to calculate the adsorption energies, electron transfer amounts, and DOS for various NO2 adsorption structures. It reveals that the orbital hybridization between O atoms in NO2 molecules and Bi atoms on the heterostructure surface promotes the electron delocalization, resulting in the redistribution of electrons. Thus, NO2 molecules can capture more electrons. This work demonstrates from both theoretical and experimental perspectives that the Bi2S3/Bi2O3 hybrid materials with the mesoporous structure have excellent NO2 detectability and could serve as a potential material for high-performance and low-energy-consumption gas sensors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/s25123612/s1; Figure S1: XPS full survey spectrums of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials and Bi2O3; Figure S2: I–V polarization curves of the pristine Bi2S3, Bi2S3/Bi2O3 heterostructure, and Bi2O3-based sensors at RT; Figure S3: The response curves of the Bi2S3/Bi2O3 heterostructure-based sensor under various humidities to 8 ppm NO2; Figure S4: The resistance curves of the Bi2S3/Bi2O3 heterostructure-based sensor under various humidities to 8 ppm NO2; Figure S5: The work functions of the (a) Bi2S3, (b) Bi2S3/Bi2O3 hybrid materials and (c) Bi2O3; Table S1: The relative contents of the OC (chemisorbed oxygen), OL (lattice oxygen), and OV (oxygen vacancy) in the synthesized materials; Table S2: The binding energies ( E b ) of NO2 molecule at various adsorption sites on the surface of Bi2S3/Bi2O3 heterostructure.

Author Contributions

Conceptualization, W.L.; Formal analysis, X.L. (Xinlei Li); Investigation, J.C.; Data curation, S.S.; Writing—original draft, W.L.; Supervision, D.G. and X.L. (Xiaogan Li); Funding acquisition, W.L. and D.G. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support provided by the Startup Foundation for Introducing Talent of Nanjing University of Information Science and Technology (No. 2023r020) and the Natural Science Foundation of Jiangsu Province (No. BK20220620) is appreciated.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhao, X.; Xu, Z.; Zhang, Z.; Liu, J.; Yan, X.; Zhu, Y.; Attfield, J.P.; Yang, M. Titanium Nitride Sensor for Selective NO2 Detection. Nat. Commun. 2025, 16, 182. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Jiang, Y.; Yuan, Z.; Liu, B.; Zhao, Q.; Huang, Q.; Li, Z.; Zeng, W.; Duan, Z.; Tai, H. Synergistic Effect of Electron Scattering and Space Charge Transfer Enabled Unprecedented Room Temperature NO2 Sensing Response of SnO2. Small 2023, 19, 535–546. [Google Scholar] [CrossRef]
  3. Chen, Y.; Li, Z.; Tang, T.; Cheng, Y.; Cheng, L.; Wang, X.; Haidry, A.A.; Jannat, A.; Ou, J.Z. Room-Temperature Optoelectronic NO2 Sensing Using Two-Dimensional Gallium Oxyselenides. ACS Appl. Nano Mater. 2024, 7, 3229–3238. [Google Scholar] [CrossRef]
  4. Sun, S.; Li, X.; Sun, Y.; Wang, N.; Huang, B.; Li, X. Fabrication of TeNT/TeO2 Heterojunction Based Sensor for Ultrasensitive Detection of NO2. J. Hazard. Mater. 2025, 487, 137229. [Google Scholar] [CrossRef]
  5. Wang, B.; Gao, X.; He, J.; Xiao, Y.; Liu, Y.; Jia, X.; Zhang, K.; Wang, C.; Sun, P.; Liu, F.; et al. Room-Temperature Ppb-Level NO2 Sensor Based on Three-Dimensional Mo2CTx Nano-Crumpled Spheres. Sens. Actuators B Chem. 2024, 399, 134790. [Google Scholar] [CrossRef]
  6. Du, H.; Zhang, Z.; Jiang, X.; Wang, J.; Yi, W.; Li, X.; Chu, J. Enhancement of NO2 Gas Sensing Properties of Polypyrrole by Polarization Doping with DBS: Experimental and DFT Studies. ACS Appl. Mater. Interfaces 2023, 15, 52961–52970. [Google Scholar] [CrossRef]
  7. Hu, Y.; Li, T.; Zhang, J.; Guo, J.; Wang, W.; Zhang, D. High-Sensitive NO2 Sensor Based on p-NiCo2O4/n-WO3 Heterojunctions. Sens. Actuators B Chem. 2022, 352, 130912. [Google Scholar] [CrossRef]
  8. Ding, Y.; Guo, X.; Kuang, D.; Hu, X.; Zhou, Y.; He, Y.; Zang, Z. Hollow Cu2O Nanospheres Loaded with MoS2/Reduced Graphene Oxide Nanosheets for Ppb-Level NO2 Detection at Room Temperature. J. Hazard. Mater. 2021, 416, 126218. [Google Scholar] [CrossRef]
  9. Zhang, W.; Wang, W.; Ge, Y.; Sun, L.; Zhou, C.; Sun, Y.; Hu, J. SnS2/Mo4/3B2 MBene Microcomposites for Highly Sensitive NO2 Sensor at Room Temperature. Sens. Actuators B Chem. 2024, 420, 136496. [Google Scholar] [CrossRef]
  10. Li, Q.; Zeng, W.; Li, Y. Metal Oxide Gas Sensors for Detecting NO2 in Industrial Exhaust Gas: Recent Developments. Sens. Actuators B Chem. 2022, 359, 131579. [Google Scholar] [CrossRef]
  11. Li, Z.; Liu, X.; Zhou, M.; Zhang, S.; Cao, S.; Lei, G.; Lou, C.; Zhang, J. Plasma-Induced Oxygen Vacancies Enabled Ultrathin ZnO Films for Highly Sensitive Detection of Triethylamine. J. Hazard. Mater. 2021, 415, 125757. [Google Scholar] [CrossRef] [PubMed]
  12. Lim, K.; Jo, Y.M.; Yoon, J.W.; Kim, J.S.; Lee, D.J.; Moon, Y.K.; Yoon, J.W.; Kim, J.H.; Choi, H.J.; Lee, J.H. A Transparent Nanopatterned Chemiresistor: Visible-Light Plasmonic Sensor for Trace-Level NO2 Detection at Room Temperature. Small 2021, 17, e2100438. [Google Scholar] [CrossRef]
  13. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Kim, T.W. Recent Advances in Energy-Saving Chemiresistive Gas Sensors: A Review. Nano Energy 2021, 79, 105369. [Google Scholar] [CrossRef] [PubMed]
  14. Goswami, P.; Gupta, G. Recent Progress of Flexible NO2 and NH3 Gas Sensors Based on Transition Metal Dichalcogenides for Room Temperature Sensing. Mater. Today Chem. 2022, 23, 100726. [Google Scholar] [CrossRef]
  15. Tyagi, D.; Wang, H.; Huang, W.; Hu, L.; Tang, Y.; Guo, Z.; Ouyang, Z.; Zhang, H. Recent Advances in Two-Dimensional-Material-Based Sensing Technology toward Health and Environmental Monitoring Applications. Nanoscale 2020, 12, 3535–3559. [Google Scholar] [CrossRef]
  16. Galstyan, V.; Moumen, A.; Kumarage, G.W.C.; Comini, E. Progress towards Chemical Gas Sensors: Nanowires and 2D Semiconductors. Sens. Actuators B Chem. 2022, 357, 131466. [Google Scholar] [CrossRef]
  17. Rohaizad, N.; Mayorga-Martinez, C.C.; Fojtů, M.; Latiff, N.M.; Pumera, M. Two-Dimensional Materials in Biomedical, Biosensing and Sensing Applications. Chem. Soc. Rev. 2021, 50, 619–657. [Google Scholar] [CrossRef]
  18. Sysoev, V.V.; Lashkov, A.V.; Lipatov, A.; Plugin, I.A.; Bruns, M.; Fuchs, D.; Varezhnikov, A.S.; Adib, M.; Sommer, M.; Sinitskii, A. UV-Light-Tunable p-/n-Type Chemiresistive Gas Sensors Based on Quasi-1D TiS3 Nanoribbons: Detection of Isopropanol at Ppm Concentrations. Sensors 2022, 22, 9815. [Google Scholar] [CrossRef]
  19. Zhu, L.; Zhang, J.; Wang, J.; Liu, J.; Yan, W. Hierarchical Heterojunctions of Metal Sulfide WS2 Nanosheets/Metal Oxide In2O3 Nanofibers for an Efficient Detection of Formaldehyde. Nanomaterials 2024, 14, 1702. [Google Scholar] [CrossRef]
  20. Meng, T.Y.; Hu, K.S.; Hsueh, T.J. A 2D Tin Disulfide Nanosheet Gas Sensor for Hydrogen Sulfide. IEEE Sens. J. 2025, 25, 3559181. [Google Scholar] [CrossRef]
  21. Harke, S.S.; Jadhav, Y.; Patil, V.B.; Gurnani, C. Facile Solution-Processed Deposition of Bi2S3 Nanostructures for a Highly Sensitive and Selective Room-Temperature NO2 Sensor. ACS Appl. Electron. Mater. 2025, 7, 1291–1304. [Google Scholar] [CrossRef]
  22. Russ, T.; Hu, Z.; Li, L.; Zhou, L.; Liu, H.; Weimar, U.; Barsan, N. In Operando Investigation of the Concentration Dependent NO2 Sensing Mechanism of Bi2S3 Nanorods at Low Temperatures and the Interference of O3. ACS Sens. 2022, 7, 3023–3031. [Google Scholar] [CrossRef]
  23. Yang, Y.; Liu, C.; Wang, Y.; Hao, J. Nanorods Assembled Hierarchical Bi2S3 for Highly Sensitive Detection of Trace NO2 at Room Temperature. Chemosensors 2024, 12, 8. [Google Scholar] [CrossRef]
  24. Yang, Y.; Xin, T.; Liu, C.; Zhang, T.; Hao, W.; Wang, Y.; Hao, J. Urchin-like Bi2S3 Nanostructures with Rich Sulfur Vacancies for Ppb-Level NO2 Sensing. J. Alloys Compd. 2023, 930, 167467. [Google Scholar] [CrossRef]
  25. Kan, H.; Yang, W.; Guo, Z.; Li, M. Highly Sensitive Room-Temperture NO2 Gas Sensor Based on Bi2S3 Nanorods. J. Mater. Sci. Mater. Electron. 2024, 35, 331. [Google Scholar] [CrossRef]
  26. Li, H.; Huang, H.; Huang, W.; Zhang, X.; Hai, G.; Lai, F.; Zhu, T.; Bai, S.; Zhang, N.; Liu, T. Interfacial Accumulation and Stability Enhancement Effects Triggered by Built-in Electric Field of SnO2/LaOCl Nanofibers Boost Carbon Dioxide Electroreduction. Small 2024, 2402654, e2402654. [Google Scholar] [CrossRef]
  27. Kim, Y.; Lee, S.; Song, J.G.; Ko, K.Y.; Woo, W.J.; Lee, S.W.; Park, M.; Lee, H.; Lee, Z.; Choi, H.; et al. 2D Transition Metal Dichalcogenide Heterostructures for P- and n-Type Photovoltaic Self-Powered Gas Sensor. Adv. Funct. Mater. 2020, 30, 2003360. [Google Scholar] [CrossRef]
  28. Yu, M.; Li, J.; Yin, D.; Zhou, Z.; Wei, C.; Wang, Y.; Hao, J. Enhanced Oxygen Anions Generation on Bi2S3/Sb2S3 Heterostructure by Visible Light for Trace H2S Detection at Room Temperature. J. Hazard. Mater. 2024, 476, 134932. [Google Scholar] [CrossRef]
  29. Chen, X.; Wang, T.; Shi, J.; Lv, W.; Han, Y.; Zeng, M.; Yang, J.; Hu, N.; Su, Y.; Wei, H.; et al. A Novel Artificial Neuron-Like Gas Sensor Constructed from CuS Quantum Dots/Bi2S3 Nanosheets. Nano-Micro Lett. 2022, 14, 8. [Google Scholar] [CrossRef]
  30. Liu, L.; Gao, Y.; Hu, Q. Controllable Epitaxial-like Growth of Bi2S3/MoS2 Hybrid Aerogel Nanostructures for Sensitive NO2 Detection at Room Temperature. ACS Appl. Nano Mater. 2023, 6, 8260–8269. [Google Scholar] [CrossRef]
  31. Yang, Y.; Zhu, M.; Zhang, H.; Wang, B.; Chen, C.; Li, J.; Wang, Y.; Hao, J. Room Temperature Gas Sensor Based on rGO/Bi2S3 Heterostructures for Ultrasensitive and Rapid NO2 Detection. Chem. Eng. J. 2024, 490, 151872. [Google Scholar] [CrossRef]
  32. Zhao, X.; Zhi, M.; Hang, D.; Ren, Q.; Zhang, P.; Chen, C.; Chen, Q.; Li, Q.; Zhang, Z.; Yan, J.; et al. Ultrasensitive NO2 Gas Sensor Based on MoS2 Modified Urchin-like Bi2S3 Heterojunction. Phys. E Low-Dimens. Syst. Nanostructures. 2023, 147, 115575. [Google Scholar] [CrossRef]
  33. Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  34. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B Condens. Matter Mater. Phys. 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  36. Li, X.; Sun, S.; Wang, N.; Huang, B.; Li, X. SnTe/SnSe Heterojunction Based Ammonia Sensors with Excellent Withstand to Ambient Humidities. Small 2024, 20, 2309831. [Google Scholar] [CrossRef]
  37. Xing, H.; Li, X.; Sun, S.; Huang, B.; Li, X. Low-Temperature NO2 Sensor Based on γ-In2Se3/In2O3 Nanoflower Heterojunction. Sens. Actuators B Chem. 2024, 415, 136034. [Google Scholar] [CrossRef]
  38. Lan, M.; Dong, X.; Zheng, N.; Zhang, X.; Wang, Y.; Zhang, X. In-Situ Construction of Novel Sulfur-Vacancy-Rich Bi/Bi2S3/SnS2 Z-Scheme Heterostructure Photocatalysts for Efficient Cr(VI) Reduction and Nitrogen Fixation. J. Mater. Sci. Technol. 2023, 167, 237–247. [Google Scholar] [CrossRef]
  39. Chen, D.N.; Wang, G.Q.; Mei, L.P.; Feng, J.J.; Wang, A.J. Dual II-Scheme Nanosheet-like Bi2S3/Bi2O3/Ag2S Heterostructures for Ultrasensitive PEC Aptasensing of Aflatoxin B1 Coupled with Catalytic Signal Amplification by Dendritic Nanorod-like Au@Pd@Pt Nanozyme. Biosens. Bioelectron. 2023, 223, 115038. [Google Scholar] [CrossRef]
  40. Wang, X.; Du, C.; Yu, Y.; Li, W.; Li, T.; Wang, S.; Mao, S.; Wang, Y.; Zhao, J.; Xiong, C. Efficient Photocatalytic Degradation with a Lattice-Matched α-Bi2O3/Co3O4 Z-Scheme Heterojunction: An Integrated Experimental and DFT Study. J. Water Process Eng. 2024, 65, 105829. [Google Scholar] [CrossRef]
  41. Cheng, X.; Xiao, X.; Wang, F.; Lu, T.; Zhang, Y. Heterojunctions Based on BiOBr Nanosheets Decorated on α-Bi2O3 for Photodegradation of Rhodamine B. ACS Appl. Nano Mater. 2024, 7, 4413–4422. [Google Scholar] [CrossRef]
  42. Tang, Y.; Gan, S.; Zhong, L.; Sun, Z.; Xu, L.; Liao, C.; Lin, K.; Cui, X.; He, D.; Ma, Y.; et al. Lattice Proton Intercalation to Regulate WO3-Based Solid-Contact Wearable PH Sensor for Sweat Analysis. Adv. Funct. Mater. 2022, 32, 2107653. [Google Scholar] [CrossRef]
  43. Bai, Y.; Fu, H.; Yang, X.; Xiong, S.; Li, S.; An, X. Conductometric Isopropanol Gas Sensor: Ce-Doped In2O3 Nanosheet-Assembled Hierarchical Microstructure. Sens. Actuators B Chem. 2023, 377, 133007. [Google Scholar] [CrossRef]
  44. Zhang, Q.; Xie, G.; Duan, M.; Liu, Y.; Cai, Y.; Xu, M.; Zhao, K.; Tai, H.; Jiang, Y.; Su, Y. Zinc Oxide Nanorods for Light-Activated Gas Sensing and Photocatalytic Applications. ACS Appl. Nano Mater. 2023, 6, 17445–17456. [Google Scholar] [CrossRef]
  45. Cheng, L.; Li, Y.; Sun, G.; Cao, J.; Wang, Y. Modification of Bi2O3 on ZnO Porous Nanosheets-Assembled Architecture for Ultrafast Detection of TEA with High Sensitivity. Sens. Actuators B Chem. 2023, 376, 132986. [Google Scholar] [CrossRef]
  46. Kashmery, H.A.; El-Hout, S.I. Bi2S3/Bi2O3 Nanocomposites as Effective Photocatalysts for Photocatalytic Degradation of Tetracycline under Visible-Light Exposure. Opt. Mater. 2023, 135, 113231. [Google Scholar] [CrossRef]
  47. Wang, J.; Chen, J.; Liu, F.; Jia, M.; Zhang, Z.; Liu, M.; Jiang, L. Renewable WO3/Bi2O3 Heterojunction for Photoelectrochemical and Visual Dual-Mode Detection of Hydrogen Sulfide. Sens. Actuators B Chem. 2022, 369, 132274. [Google Scholar] [CrossRef]
  48. Guang, Q.; Huang, B.; Yu, J.; Bonyani, M.; Moaddeli, M.; Kanani, M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Li, X. PtS-Decorated WS2 Microflakes Based Sensors for Selective Ammonia Detection at Room Temperature. Sens. Actuators B Chem. 2023, 394, 134399. [Google Scholar] [CrossRef]
  49. Ma, Y.; Jiang, X.; Sun, R.; Yang, J.; Jiang, X.; Liu, Z.; Xie, M.; Xie, E.; Han, W. Z-Scheme Bi2O2.33/Bi2S3 Heterojunction Nanostructures for Photocatalytic Overall Water Splitting. Chem. Eng. J. 2020, 382, 123020. [Google Scholar] [CrossRef]
  50. Cheng, D.; Wu, H.; Feng, C.; Ding, Y.; Mei, H. Bifunctional Photoelectrochemical Sensor Based on Bi/Bi2S3/BiVO4 for Detecting Hexavalent Chromium and Hydrogen Peroxide. Sens. Actuators B Chem. 2022, 353, 131108. [Google Scholar] [CrossRef]
  51. Ou, L.X.; Liu, M.Y.; Zhu, L.Y.; Zhang, D.W.; Lu, H.L. Recent Progress on Flexible Room-Temperature Gas Sensors Based on Metal Oxide Semiconductor. Nano-Micro Lett. 2022, 14, 206. [Google Scholar] [CrossRef] [PubMed]
  52. Zheng, Y.; Cui, X.; Zhou, Y.; Zhang, H.; Cao, L.; Gao, L.; Yin, H.; Ai, S. MXene Enhanced Photoactivity of Bi2O3/Bi2S3 Heterojunction with G-Wire Superstructure for Photoelectrochemical Detection of TET1 Protein. ACS Sens. 2022, 7, 3116–3125. [Google Scholar] [CrossRef] [PubMed]
  53. Ikram, M.; Liu, L.; Lv, H.; Liu, Y.; Ur Rehman, A.; Kan, K.; Zhang, W.J.; He, L.; Wang, Y.; Wang, R.; et al. Intercalation of Bi2O3/Bi2S3 Nanoparticles into Highly Expanded MoS2 Nanosheets for Greatly Enhanced Gas Sensing Performance at Room Temperature. J. Hazard. Mater. 2019, 363, 335–345. [Google Scholar] [CrossRef]
  54. Li, H.; Jing, P.; He, C.; Pan, Z.; Liu, J.; Cui, Y.; Wang, J. Interfacial Configuration and Mechanism Insights of an All-Solid-State Z-Scheme BaTiO3/Bi/Bi2O3 Heterojunctions for Rapid Removal of Tetracycline Antibiotics. Appl. Surf. Sci. 2023, 615, 156416. [Google Scholar] [CrossRef]
  55. Chang, J.; Liang, W.; Wang, W.; Wu, D.; Jiang, K.; Wang, G.; Xu, F.; Gao, Z. Oxygen Vacancies Enriched Bi2O3 as High Capacity and High Rate Negative Material for Aqueous Alkali Battery. Appl. Surf. Sci. 2022, 601, 154296. [Google Scholar] [CrossRef]
  56. Li, X.; Zhang, Z.; Sun, S.; Wang, N.; Huang, B.; Li, X. Hierarchically In2S3@In2O3 Nanorods Heterojunctions for Enhanced NO2 Sensing at Lower Operating Temperature. Sens. Actuators B Chem. 2024, 419, 136360. [Google Scholar] [CrossRef]
  57. Chen, X.; Shi, J.; Wang, T.; Zheng, S.; Lv, W.; Chen, X.; Yang, J.; Zeng, M.; Hu, N.; Su, Y.; et al. High-Performance Wearable Sensor Inspired by the Neuron Conduction Mechanism through Gold-Induced Sulfur Vacancies. ACS Sens. 2022, 7, 816–826. [Google Scholar] [CrossRef]
  58. Wang, X.; Chen, Y.; Zeng, Z.; Yan, M.; Jia, X.; Hu, P.; Xu, J.; Xue, Z.; Xu, J. Tailoring the Injection Action of Oxygen over Top-Surface of Bismuth Sulfide to Change Reactive Electron Transfer Path for Flexible NO2 Sensors. Mater. Sci. Eng. R. Rep. 2024, 160, 100819. [Google Scholar] [CrossRef]
  59. Chang, F.; Peng, S.; Yan, W.; Yang, C.; Li, S.; Liu, X. A Novel and Facile Procedure to Decorate Bi2O3 with Bi2S3 Nanocrystals: Composites Synthesis, Analyses, and Photocatalytic Performance Assessment. Colloids Surf. A Physicochem. Eng. Asp. 2021, 610, 125640. [Google Scholar] [CrossRef]
  60. Gasso, S.; Sohal, M.K.; Mahajan, A. MXene Modulated SnO2 Gas Sensor for Ultra-Responsive Room-Temperature Detection of NO2. Sens. Actuators B Chem. 2022, 357, 131427. [Google Scholar] [CrossRef]
  61. Bai, H.; Guo, H.; Wang, J.; Dong, Y.; Liu, B.; Xie, Z.; Guo, F.; Chen, D.; Zhang, R.; Zheng, Y. A Room-Temperature NO2 Gas Sensor Based on CuO Nanoflakes Modified with rGO Nanosheets. Sens. Actuators B Chem. 2021, 337, 129783. [Google Scholar] [CrossRef]
  62. Hau, H.H.; Duong, T.T.H.; Man, N.K.; Thi Viet Nga, T.; Thi Xuan, C.; Thi Thanh Le, D.; Van Toan, N.; Hung, C.M.; Van Duy, N.; Van Hieu, N.; et al. Enhanced NO2 Gas-Sensing Performance at Room Temperature Using Exfoliated MoS2 Nanosheets. Sens. Actuators A Phys. 2021, 332, 113137. [Google Scholar] [CrossRef]
  63. Yakout, S.M. New Efficient Sunlight Photocatalysts Based on Gd, Nb, V and Mn Doped Alpha-Bi2O3 Phase. J. Environ. Chem. Eng. 2020, 8, 103644. [Google Scholar] [CrossRef]
  64. Motaung, M.P.; Onwudiwe, D.C.; Lei, W. Microwave-Assisted Synthesis of Bi2S3 and Sb2S3 Nanoparticles and Their Photoelectrochemical Properties. ACS Omega 2021, 6, 18975–18987. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of the fabrication process of Bi2S3/Bi2O3 hybrid materials with mesoporous structure.
Figure 1. Schematic diagram of the fabrication process of Bi2S3/Bi2O3 hybrid materials with mesoporous structure.
Sensors 25 03612 g001
Figure 2. TGA curve of the pristine Bi2S3 powders from 25 °C to 800 °C in air.
Figure 2. TGA curve of the pristine Bi2S3 powders from 25 °C to 800 °C in air.
Sensors 25 03612 g002
Figure 3. (a) XRD patterns of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3. (b) An enlarged view of the selected region of XRD.
Figure 3. (a) XRD patterns of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3. (b) An enlarged view of the selected region of XRD.
Sensors 25 03612 g003
Figure 4. (ac) SEM images of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3; (d) HRTEM image of Bi2S3/Bi2O3; and (ei) elemental mapping of Bi2S3/Bi2O3.
Figure 4. (ac) SEM images of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3; (d) HRTEM image of Bi2S3/Bi2O3; and (ei) elemental mapping of Bi2S3/Bi2O3.
Sensors 25 03612 g004
Figure 5. N2 adsorption/desorption isotherms and pore size distribution of (a) pristine Bi2S3 and (b) Bi2S3/Bi2O3 hybrid materials.
Figure 5. N2 adsorption/desorption isotherms and pore size distribution of (a) pristine Bi2S3 and (b) Bi2S3/Bi2O3 hybrid materials.
Sensors 25 03612 g005
Figure 6. XPS spectra of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 in the (a) Bi 4f and S 2p regions and (b) O 1s region.
Figure 6. XPS spectra of the pristine Bi2S3, Bi2S3/Bi2O3 hybrid materials, and Bi2O3 in the (a) Bi 4f and S 2p regions and (b) O 1s region.
Sensors 25 03612 g006
Figure 7. EPR spectra of the Bi2S3/Bi2O3 hybrid materials and Bi2O3.
Figure 7. EPR spectra of the Bi2S3/Bi2O3 hybrid materials and Bi2O3.
Sensors 25 03612 g007
Figure 8. Response curves of the Bi2S3-based sensor and Bi2S3/Bi2O3 heterostructure-based sensors to 8 ppm NO2 under various (a) oxidation temperatures and (b) oxidation times, and the response/recovery time of (c) the Bi2S3/Bi2O3 heterostructure-based sensor and (d) the Bi2S3-based sensor.
Figure 8. Response curves of the Bi2S3-based sensor and Bi2S3/Bi2O3 heterostructure-based sensors to 8 ppm NO2 under various (a) oxidation temperatures and (b) oxidation times, and the response/recovery time of (c) the Bi2S3/Bi2O3 heterostructure-based sensor and (d) the Bi2S3-based sensor.
Sensors 25 03612 g008
Figure 9. (a) Resistance curve of the Bi2S3/Bi2O3 heterostructure-based sensor to NO2 concentrations from 0.1 ppm to 8 ppm (inset describes the corresponding response fitting curve), (b) signal repeatability, (c) response and resistance of the sensor under various humidity conditions, (d) selectivity, and (e) long-term stability to 8 ppm NO2.
Figure 9. (a) Resistance curve of the Bi2S3/Bi2O3 heterostructure-based sensor to NO2 concentrations from 0.1 ppm to 8 ppm (inset describes the corresponding response fitting curve), (b) signal repeatability, (c) response and resistance of the sensor under various humidity conditions, (d) selectivity, and (e) long-term stability to 8 ppm NO2.
Sensors 25 03612 g009
Figure 10. Schematic diagrams of the sensitization mechanisms: (a) bandgap structure of Bi2S3, Bi2O3, and Bi2S3/Bi2O3 hybrid materials and (b) the related gas-sensitive procedure.
Figure 10. Schematic diagrams of the sensitization mechanisms: (a) bandgap structure of Bi2S3, Bi2O3, and Bi2S3/Bi2O3 hybrid materials and (b) the related gas-sensitive procedure.
Sensors 25 03612 g010
Figure 11. Adsorption structures of NO2 adsorbed on (a) Bi2S3, (b) Bi2S3/Bi2O3 heterostructure, and (c) Bi2O3, and the charge density difference in (d) Bi2S3, (e) Bi2S3/Bi2O3 heterostructure, and (f) Bi2O3. (the yellow region represents electron accumulation, while the cyan region stands for the electron depletion).
Figure 11. Adsorption structures of NO2 adsorbed on (a) Bi2S3, (b) Bi2S3/Bi2O3 heterostructure, and (c) Bi2O3, and the charge density difference in (d) Bi2S3, (e) Bi2S3/Bi2O3 heterostructure, and (f) Bi2O3. (the yellow region represents electron accumulation, while the cyan region stands for the electron depletion).
Sensors 25 03612 g011
Figure 12. The TDOS and PDOS of NO2 adsorbed on (a,b) Bi2S3, (c,d) Bi2S3/Bi2O3 heterostructure, and (e,f) Bi2O3.
Figure 12. The TDOS and PDOS of NO2 adsorbed on (a,b) Bi2S3, (c,d) Bi2S3/Bi2O3 heterostructure, and (e,f) Bi2O3.
Sensors 25 03612 g012
Table 1. The comparison of NO2 sensing performance of the prepared Bi2S3/Bi2O3 heterostructure-based sensor with the previously reported Bi2S3-based sensors at RT.
Table 1. The comparison of NO2 sensing performance of the prepared Bi2S3/Bi2O3 heterostructure-based sensor with the previously reported Bi2S3-based sensors at RT.
MaterialsGasTemp. (°C)Conc. (ppm) Responses   ( R g a s / R a i r ) Tres/Tres (s)LOD (ppm)Refs.
Bi2S3NO2RT1001.3914/2571[21]
Bi2S3NO2RT1012.239/6962[25]
Au/Bi2S3NO2RT55.621/3710.25[57]
CuS/Bi2S3NO2RT103.418/3380.5[29]
Bi2S3/Bi2O3NO2RT101.2826.6/69.90.1[58]
Bi2S3/Bi2O3NO2RT87.8571/2380.1this work
Tres/Trec = response/recovery time; LOD= limit of detection; Temp. = temperature; Conc. = concentration.
Table 2. The adsorption energies ( E a d s ), electron transfer amounts ( Q ), and the distances ( d ) between NO2 and corresponding adsorption sites.
Table 2. The adsorption energies ( E a d s ), electron transfer amounts ( Q ), and the distances ( d ) between NO2 and corresponding adsorption sites.
Configurations E a d s (eV) Q (e) d (Å)
Bi2S3/Bi2O3−1.26−0.672.41
Bi2S3−0.21−0.172.97
Bi2O3−0.39−0.082.81
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

Liu, W.; Chen, J.; Gu, D.; Sun, S.; Li, X.; Li, X. Mesoporous Bi2S3/Bi2O3 Heterostructure-Based Sensors for Sub-ppm NO2 Detection at Room Temperature. Sensors 2025, 25, 3612. https://doi.org/10.3390/s25123612

AMA Style

Liu W, Chen J, Gu D, Sun S, Li X, Li X. Mesoporous Bi2S3/Bi2O3 Heterostructure-Based Sensors for Sub-ppm NO2 Detection at Room Temperature. Sensors. 2025; 25(12):3612. https://doi.org/10.3390/s25123612

Chicago/Turabian Style

Liu, Wei, Jiashuo Chen, Ding Gu, Shupeng Sun, Xinlei Li, and Xiaogan Li. 2025. "Mesoporous Bi2S3/Bi2O3 Heterostructure-Based Sensors for Sub-ppm NO2 Detection at Room Temperature" Sensors 25, no. 12: 3612. https://doi.org/10.3390/s25123612

APA Style

Liu, W., Chen, J., Gu, D., Sun, S., Li, X., & Li, X. (2025). Mesoporous Bi2S3/Bi2O3 Heterostructure-Based Sensors for Sub-ppm NO2 Detection at Room Temperature. Sensors, 25(12), 3612. https://doi.org/10.3390/s25123612

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