Next Article in Journal / Special Issue
Flexible Humidity Sensor Based on Chemically Reduced Graphene Oxide
Previous Article in Journal
A Nanodiamond-Based Electrochemical Sensor for the Determination of Paracetamol in Pharmaceutical Samples
Previous Article in Special Issue
Optimization by Hydrogen Plasma Treatment of a-CH and Hydrogen/Nitrogen-Assisted a-CH Layers for SAW Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Metal Oxide Semiconductor Heterojunctions for the Detection of Volatile Organic Compounds

School of Material Science and Engineering, Shanghai University of Engineering Science, Shanghai 201620, China
*
Authors to whom correspondence should be addressed.
Chemosensors 2024, 12(12), 244; https://doi.org/10.3390/chemosensors12120244
Submission received: 10 October 2024 / Revised: 19 November 2024 / Accepted: 20 November 2024 / Published: 22 November 2024
(This article belongs to the Special Issue Functional Nanomaterial-Based Gas Sensors and Humidity Sensors)

Abstract

The efficient detection of volatile organic compounds (VOCs) is critically important in the domains of environmental protection, healthcare, and industrial safety. The development of metal oxide semiconductor (MOS) heterojunction gas-sensing materials is considered one of the most effective strategies to enhance sensor performance. This review summarizes and discusses the types of heterojunctions and their working principles, enhancement strategies, preparation methodologies, and applications in acetone and ethanol detection. To address the constraints pertaining to low sensitivity, sluggish response/recovery times, and elevated operating temperatures that are inherent in VOC sensors, several improvement methods are proposed, including doping with metals like Ag and Pd, incorporating additives such as MXene and polyoxometalates, optimizing morphologies through a fine design, and self-doping via oxygen vacancies. Furthermore, this work provides insights into the challenges faced by MOSs heterojunction-based gas sensors and outlines future research directions in this field. This review will contribute to foundational theories to overcome existing bottlenecks in MOS heterojunction technology while promoting its large-scale application in disease screening or agricultural food quality assessments.

1. Introduction

Volatile organic compounds (VOCs) generated from combustion emissions, industrial production, and biological metabolism, such as ketones, alcohols, methane, benzene, etc., have potential impacts on the environment and human health [1,2,3,4,5]. Therefore, monitoring VOCs has become one of the important tasks for environmental protection and health protection today. A variety of VOC sensors based on different working mechanisms and technologies are emerging, including metal oxide semiconductor (MOS) sensors, nanomaterials sensors, functionalized polymer sensors, and intelligent sensing systems based on machine learning and artificial intelligence [6,7,8,9,10,11,12,13,14,15,16,17].
Among these sensors, MOS sensors are widely used in various fields due to their advantages, such as simple manufacturing process, low energy consumption, and cost-effectiveness. In particular, zinc oxide (ZnO)- [18,19,20,21,22,23], copper oxide (CuO)- [24,25,26,27,28,29,30,31,32], tin dioxide (SnO2)- [33,34,35,36,37,38,39,40], nickel oxide (NiO)- [41,42,43,44,45,46], and cobalt tetroxide (Co3O4)-based [47,48] materials are extensively employed as sensing materials in gas sensors. However, MOSs sensors still face some key challenges in sensing performance that need to be addressed, such as high detection limit, high operating temperature, high baseline resistance, and poor moisture resistance. Consequently, various strategies have been developed to enhance sensing performances, including constructing special morphologies, introducing oxygen vacancies, and doping impurities. Among these, constructing heterojunctions and combining them with other strategies are widely adopted approaches.
This review aims to explore the improvement of VOC detection by chemiresistive sensors through the construction of MOS heterojunctions. Specifically addressed will be the types of heterojunctions, underlying working principles, improvement measures, and preparation methodologies as well as applications in acetone and ethanol detection.

2. Types of Heterojunctions and Preparation Methods

2.1. Types of Heterojunctions

Heterojunctions can generally be defined as interface structures composed of semiconductor contacts with two different band structures. According to the relationships between the energy bands of MOSs, heterojunctions can be categorized into three types: straddling-gap-like, staggered-gap-like, and broken-gap-like (Figure 1). For a type I heterojunction with a straddling gap, the bottom of the conduction band and the top of the valence band of the semiconductor material with a small band gap (Egap1) are in the band gap of the semiconductor material with a wide band gap (Egap2). For a type II heterojunction with a staggered gap, the band gaps of two semiconductor materials are interlaced with each other. For a type III heterojunction with a broken gap, the bandgap of two semiconductor materials is completely staggered. Note that the band gap energy of MOSs, defined as the energetic separation between the valence band maximum and the conduction band minimum within a material, varies based on factors such as crystalline phase, purity levels, and other pertinent parameters. In the context of MOS heterojunctions, the band gap of frequently utilized materials like α-phase iron(III) oxide (α-Fe2O3) is approximately 2.0 to 2.4 eV, whereas titanium dioxide (TiO2) demonstrates a band gap in the range of 3.0 to 3.5 eV [49]. As an illustrative example, ZnO with a band gap of 3.4 eV and indium oxide (In2O3) with a band gap of 3.6 eV can potentially form a type II heterojunction [50], showcasing a distinctive alignment of energy bands conducive to specific electronic and optoelectronic applications.

2.2. Working Principle of Heterojunction Sensors

Fabricated MOS-heterojunction-based materials can be used in the sensing materials in VOC detection. The working principle of heterojunction sensors for VOC detection is complex and diverse.
This review focuses on chemiresistive sensors, which operate based on the resistance changes that result from the adsorption and desorption processes of gas molecules, alongside the chemical reactions that occur at the surfaces of sensing materials. The sensing mechanism of chemiresistive gas sensors relies on a fluctuation in the resistance of sensor materials [51,52]. When exposed to air, oxygen molecules initially adsorb onto the surface of the sensing material (Equation (1)). After reaching an equilibrium between chemisorbed sites and the oxygen molecules, the adsorbed oxygen molecules draw electrons from the conduction band of the material and involve into various oxygen species (O2−, O, and O2−; Equations (2)–(4)) [53]. This electron-trapping process increases the concentration of holes and thus reduces/increases the resistance of n-type/p-type semiconductors (Ra).
When the sensor is transferred to the reducing gas environment at the optimal operating temperature, the target gas will undergo a reaction with oxygen ions, resulting in the liberation of free electrons into the conduction band. This process facilitates the recombination process between holes and electrons, resulting in fluctuations in the sensor resistance (Rg). Specifically, ethanol or acetone reacts with oxygen ions to produce CO2 and H2O and release electrons (Equations (5)–(8)) [54,55]. As indicated in Equations (2)–(4), the key oxygen species depend on the optimal operating temperature [54,55]. Within the common range of optimal operating temperatures for acetone and ethanol detection (100–300 °C), the dominant oxygen species is O. For ethanol reaction, the intermediates depend on the surface properties of the materials, which could be CH2CH2 or CH3CHO [56,57].
O 2 gas O 2 ads ,
O 2 ads + e O 2 ads T   < 100   ° C ,
O 2 ads + e   2 O ads 100   ° C T   < 300   ° C ,
O 2 ads + 4 e   2 O 2 ads T < 300   ° C ,
C H 3 COC H 3 + 8 O ads 3 C O 2 gas + 3 H 2 O vap + 8 e ,
C H 3 COC H 3 + 8 O 2 ads 3 C O 2 gas + 3 H 2 O vap + 16 e ,
C H 3 C H 2 OH + 6 O ads 2 C O 2 gas + 3 H 2 O vap + 6 e ,
C H 3 C H 2 OH + 6 O 2 ads   2 C O 2 gas + 3 H 2 O vap + 12 e

2.3. Strategies for Improving the VOC Sensing Performances of MOSs Heterojunctions

The above working mechanism provides hints for developing highly sensitive and selective VOC sensors. To improve the sensing performance of MOSs heterojunctions, the following strategies can be adopted (Figure 2):
(I) Adjust the Fermi level of the heterojunction components. The Fermi level control of MOSs can be achieved through various methods, mainly based on changing the electronic structure of the semiconductor material or environmental conditions. Specifically, this involves the following strategies:
Changing the concentration of doping elements. It is widely accepted that doping can modify the electronic properties of materials. For n-type doping, the doping element provides additional electrons. Consequently, reducing the concentration of these doping elements can lower the concentration of free electrons, thereby shifting the Fermi level towards the valence band and lowering the Fermi level. Conversely, increasing the doping concentration of doping elements can raise the Fermi level. For p-type doping, this approach is usually used to increase the hole concentration of p-type semiconductors. Changing the p-type doping concentration may indirectly affect the Fermi level distribution within MOSs structures.
Temperature control. Temperature is an important factor affecting the Fermi level. Lowering the operating temperature can shift the Fermi level of a semiconductor towards the valence band, as the thermal motion of electrons weakens at low temperatures, resulting in more states occupying lower energy levels. When the temperature is increased, the thermal motion of electrons is enhanced, which can cause the Fermi level of the semiconductor to move towards the conduction band. However, it should be noted that excessively high temperatures may lead to performance degradation in, or even the failure of MOSs materials.
External energy field regulation. By applying an external electric field, the distribution of electrons and holes in MOSs can be altered, thereby affecting the position of the Fermi level. The light irradiation may also change the Fermi level of MOSs by shifting the band structure of MOSs. For example, Liu et al. found that under 370-nanometer light illumination, the responses of a p-SmFeO3/p-YFeO3 planar-electrode sensor to 30 ppm of ethanol, acetone, and methanol gases were 1.35, 1.28, and 1.59 times higher than those without light, respectively, [58]. However, the effectiveness of this method may be limited by light intensity, wavelength, and MOSs’ intrinsic properties.
Interface engineering. By optimizing the interface structure, such as through interface passivation, surface treatment, etc., the interface state density and Fermi level pinning effect can be altered to regulate the Fermi level [59,60,61,62,63,64].
The band structure of MOS heterojunctions can be regulated through the above strategies, thereby influencing its gas sensing performance.
(II) Morphology design. For the sensing performances of chemiresistive sensors based on MOSs, one of the widely employed strategies is via designing structures with a high specific surface area and porosity through precise micromorphology control. A large surface area can offer rich active sites for the target gas, thereby enhancing the intensity of the response [65,66,67,68]. The high porosity guarantees a high diffusion rate of the gas molecules, shortening the response/recovery time. In addition to the common nanosheets and microspheres, recently developed inverse opal photonic crystals (IOPCs) with a 3D periodic macroporous array spherical symmetry structure and fractal-like structures with interconnected and cross-connected connections are also emerging as good sensing morphologies [6,69]. Some detailed examples about the morphology design of MOS heterojunction materials for acetone and ethanol detection will be provided in Section 3 and Section 4.
(III) Construction of surface defects. Creating an appropriate amount of oxygen vacancies or other defects on the surface of the material is also widely employed to improve sensing performances [55,70,71,72,73]. These defect sites can serve as active centers, promoting the adsorption and reaction of gas molecules. A suitable amount of defect sites can significantly improve the sensing performance of materials. However, excessive defects may also lead to a decrease in material stability. Therefore, it is necessary to identify the optimal equilibrium concentrations for defects.

2.4. Preparation Methods

In addition to the strategies mentioned above to improve the intrinsic properties of MOSs, optimizing the preparation process of MOSs is also an important strategy to enhance the VOC sensing performance of MOS heterojunctions.
Common preparation methods for semiconductor heterojunction materials for VOCs sensors include the hydrothermal/solvothermal method, the co-precipitation method, the sol–gel method, the calcination method, etc. (Figure 3).
Among the various synthesis methods, the co-precipitation reaction stands out as an efficient and gentle approach, which is attributed to its low energy consumption, facilitated large-scale production, abbreviated reaction duration, and ability to yield highly uniform materials. For example, CuO-ZnO heterojunction nanocomposites have been successfully synthesized by a co-precipitation method [74]. Shruthi et al. fabricated Y2O3-In2O3 heterostructure nanocrystals using a combined co-precipitation and sol–gel method for methanol sensing [75].
The hydrothermal/solvothermal method is also a common preparation method. This method uses solvation, e.g., water as the reaction medium under high temperatures and high pressures. By adjusting parameters such as reaction temperature, pressure, and material ratio, precise control of the product formation process can be achieved. Moreover, solvothermal reactions occur in a closed system, resulting in a pollution-free reaction process that meets the requirements of green synthesis. The sensing materials fabricated by this method usually present good crystallinity. For example, Sm2O3/ZnO/SmFeO3 microspheres [76], cubic rhombic In2O3 microspheres [77], TiO2 nanofilm [78], and α-Fe2O3/NiO nanorods used for VOC detection were successfully prepared by the hydrothermal/solvothermal method [42]. However, there are also disadvantages, such as high equipment costs, harsh process conditions, and difficulty in product separation and purification.
The calcination/annealing method is also commonly used for synthesizing semiconductor heterojunction materials for VOC sensors, but this method usually requires high temperatures. Typically, various precursors such as layered double hydroxides (LDHs), are prepared first, and then semiconductor heterojunction materials are obtained through high-temperature calcination/annealing (usually at 300–1000 °C) [79,80,81]. For instance, a highly sensitive acetone sensing performance based on p-SmFeO3/n-ZnO nanocomposites was successfully synthesized by a thermal treatment method [82]. Notably, the annealing temperature may affect the phase of as-prepared nanomaterials [83].
Other than these, electrospray-assisted fabrication or the spray-drying method may also be employed [84]. For example, Li et al. fabricated ZnO-SnO2 heterojunction IOPCs via the spray-drying approach [69]. Spray drying is liquid-waste-free. It possesses the ability to form spherical or sphere-like particles in a single-step process, exhibiting excellent scalability and a remarkable production speed (within milliseconds), thereby presenting a viable option for the rapid and scalable fabrication of IOPBs [69].
In the following sections, we will focus on the discussion of the detection of two representative VOCs, acetone and ethanol, based on MOS heterojunctions.

3. P-n Heterojunctions Sensors for VOC Detection

In the past few decades, n-type MOSs (n-MOSs) have drawn significant attention for VOC detections. N-MOSs usually present higher sensitivity to a given gas than p-type MOSs (p-MOSs). That reason for this is that p-MOS and n-MOS sensors work in a hole-accumulation layer and an electron-depletion layer, respectively. This difference yields a smaller resistance change in p-MOSs than that in n-MOSs under identical band-bending conditions. The gas response follows the equation S = exp(−qΔVs/mkT), where m, qΔVs, k, and T denote constants, band bending, the Boltzmann constant, and absolute temperature, respectively. For p-MOSs sensors, m ≥ 2, and its value rises with an increase in the degree of upward band bending, whereas m = 1 for n-MOSs sensors. However, n-MOSs also show the disadvantages of high operating temperature and sensitivity to humidity. By contrast, p-MOSs could show lower operating temperatures due to their superior redox catalytic properties to those of n-MOSs. This is because p-MOSs are able to yield a hole accumulation layer, which can chemisorb the oxygen molecules of a high concentration. In addition, investigations imply that the p-MOS sensors also exhibit low humidity dependence.
Apparently, the fabrication of p-n heterojunction structures using suitable components shows the potential to reach a low detection limit, low working temperature, low base line resistance, and strong moisture tolerance by taking advantage of each component. In addition, the formed interface at the heterojunction also provides new reactive sites and adjusts the electronic properties of the sensor materials.

3.1. P-n Heterojunctions Sensors for Acetone Detection

Studies have shown that the fabrication of p-n heterojunction structures can efficiently regulate band bending and hole concentration and, consequently, promote resistance for acetone detection [85]. Zhao et al. found that compared with Co3O4, the formation of a n-p heterojunction for a ZnO-Co3O4 heterostructure reduced both the degree of upward band bending and the surface hole concentration under the influence of electron transfer from ZnO to Co3O4 [85]. The reduction in the surface hole concentration demonstrated that even minimal sensing reactions (corresponding to less surface charge transfer) can cause significant resistance changes, which, along with the excellent performance of the p-MOS, resulted in high sensitivity to and low detection limit of acetone gas [85].
Introducing oxygen vacancy is an effective strategy to improve the sensitivity of reducing gas. In general, intrinsic oxygen vacancies formed at the heterojunction interface induced by lattice mismatches are thought to play an important role in improving charge transport between heterojunction structures. Three types of oxygen species may exist on the surfaces of sensitive materials: chemisorption oxygen species (OC), vacancy oxygen species (OV), and lattice oxygen species (OL). Among these, OL is regarded as having low activity and is not involved in the sensing reaction. OC is associated with chemisorption and the presence of ionized oxygen on the surface of the sensing material. Ovs facilitate efficient gas molecular adsorption and sensing reactions by providing more active sites. Therefore, the content of OV is regarded as an indicator for obtaining better gas-sensitive characteristics. Hu et al. found that Fe2O3-Co3O4, with a higher content of OV than Fe2O3 and Co3O4, presented better sensing characterizations (Figure 4a–d) [86]. As demonstrated in Figure 4c, after fabricating a heterojunction structure by combining Fe2O3 and Co3O4, the sensor exhibited an improved response to acetone, accompanied by a significant enhancement in selectivity.
In another work, Wu et al. found that during a series of the oxide cage/nanofiber heterostructure composed of ZIF-67-derived Co3O4 and In2O3, the one with the highest OV content and Co2+/Co3+ ratio delivered the best sensing performances, demonstrating an ultrasensitive response of 954 at 50 ppm acetone concentration at 300 °C, coupled with a low detection limit of 18.8 ppb (as shown in Table 1). Additionally, it exhibited a swift response time of 4 s and favorable selectivity and repeatability, as well as robust long-term stability [55]. Besides the factor of the OV content, the formation of the type I heterojunction results in an augmentation of the degree of upward energy band bending of In2O3 in the air and a larger change in the electron concentration induced by acetone, leading to a higher gas response than pure In2O3 (Figure 5) [55]. In addition to intrinsic Ovs induced by the lattice mismatches of heterojunction components, common methods for generating oxygen vacancies include reduction treatment methods (using reducing agents such as H2, NaBH4, imidazole, L-ascorbic acid, etc.), the atmosphere deoxidation method (for the atmospheres of vacuum or inert gases, e.g., Ar, N2, He), and the ion doping method. Doping elements may react with oxygen atoms or occupy the positions of oxygen atoms, thereby indirectly introducing oxygen vacancies.
To create oxygen vacancies, another promising approach is to introduce noble metals into the sensing layer. Moreover, the introduced noble metals, such as Ag, Au, and Pd, can also facilitate oxygen dissociation. To date, various synthesis routes for decorating noble-metal nanoparticles (NPs) into MOSs heterojunction structures for gas-sensing applications have been developed. Yang et al. proposed an easy and general route for the synthesis of Ag-NiO/SnO2 nanotubes (NTs) using a combined method of electro-spinning–coating–calcination–wet chemical reduction. In the last step, Ag NPs were introduced via a wet chemical reduction method (Figure 6a). The morphologies of one-dimensional (1D) NTs and the internal hollow structures of Ag-NiO/SnO2 (Figure 6b,c) promoted the adsorption and the desorption of gas molecules, facilitating the availability of more active sites. The incorporation of p-n heterojunction structures into the 1D nanotubes (NTs) further expedited the charge transfer process. Subsequent decoration with Ag nanoparticles (NPs) augmented the number of surface oxygen vacancies, contributing to exceptional sensing performance (Figure 6d). The Ag-NiO/SnO2 NT sensors exhibited remarkable gas-sensing properties towards acetone, featuring a superior response (Ra/Rg = 15.7 at 1 ppm), excellent selectivity (the response to 1 ppm of acetone exceeded that to 5 ppm of interfering gases), swift response (12 s) and recovery (25 s) times at 1 ppm of acetone, and a minimal detection limit of 50 ppb at a low operating temperature of 190 °C [87].
The preparation of structurally stable heterojunctions is conducive to enhancing the performance of the sensor [93,94]. Generally, diverse in situ strategies, such as in situ growth and in situ oxidation, can be employed to fabricate stable structures. This stable structure can offer tight p-n contact surfaces and robust chemical bonds, increase the carrier density, and facilitate charge transfer. Utilizing WS2 as the primary phase, WS2/WO3 nanosheets with a p-n heterojunction were formed through in situ growth of WO3. The assembled gas sensor based on the type II WS2/WO3 heterojunction nanosheets demonstrated acetone detection with satisfactory repeatability, selectivity, and long-term stability [95]. A similar type II heterojunction of ZnO-CuO nanomaterial with a core–hollow cube morphology was fabricated through the formation of ZnO-Cu2O core–cube nanostructures and thermal oxidation in air (Figure 6e). The ZnO-CuO core–hollow cube nanostructure was employed as the acetone sensing material, yielding a response of 11.14 at 1 ppm acetone and 200 °C and a detection limit of approximately 9 ppb, with the detection range spanning from 40 ppb to 10 ppm. Its outstanding performance was primarily attributed to its uniform and distinctive morphology. The core–shell structure, featuring ZnO as the core and CuO as the shell, maximizes the p-n heterojunction effect and increases resistance, thereby resulting in high sensitivity and low detection limits. Additional structural factors, such as small grain size and large surface area, can also cause significant changes in resistance. Techniques such as precious metal decoration and surface functionalization can further improve sensitivity and selectivity [90]. Wang et al. directly grew a TiO2 nano ribbon (NR) array on a ceramic electrode, and the α-Fe2O3 nanoscale branches were epitaxially grown on the TiO2 NRs to obtain TiO2/α-Fe2O3 hierarchical heterostructures through a simple hydrothermal method. The length of the α-Fe2O3 branch can be precisely controlled by the reaction synthesis time. Compared with pure TiO2 NRs, the response and selectivity of the TiO2/α-Fe2O3 heterostructures to acetone were significantly enhanced within a wide temperature range (175–275 °C) [91].

3.2. P-n Heterojunctions Sensors for Ethanol Detection

Ethanol, as a typical species of VOCs, has a wide range of applications in various fields. Monitoring the concentration of ethanol in the environment could provide important information for alcohol-driving detection, industrial production, and medical diagnosis. Hence, the study of ethanol sensors has been widely explored.
A series of MOS heterojunction composites have been constructed for ethanol detection, including n-n heterojunctions (Table 2), such as ZnO/SnO2 [69,88], ZnO/α-Fe2O3 [96], and In2O3/SnO2 [97], and p-n heterojunctions, such as α-Fe2O3/NiO [98].
To overcome the high resistance of p–n heterojunctions, some suitable additives can be added, such as doping additives with moderate catalytic activity, covering polydimethylsiloxane (PDMS), carbonaceous material, and loading noble metals [104]. Two-dimensional (2D) metal materials, such as MXene, are ideal additive materials owing to their distinctive layered morphology, high electrical conductivity, substantial surface area, superior hydrophilicity, robust thermal stability, and environmentally benign characteristics [104]. Consequently, the integration of MXene with TiO2/CuO composites led to a reduction in the sensor’s resistance and a subsequent enhancement of its gas-sensing performance. As illustrated in Figure 7, the MXene was added during the preparation process, and it transferred electrons to the bonded p-type CuO, resulting in reduced resistance of the heterojunction. The sensor based on the p-type CuO/TiO2/MXene composite exhibited an exceptional response value of 95% at 1 ppm of ethanol, coupled with swift response and recovery times of 16 and 13 s, respectively, a low detection limit of 0.3 ppm, and outstanding long-term stability, which were observed for the trace detection of ethanol gas at room temperature [102].
In order to obtain enough reactive sites for the enhanced resistive ability of heterojunction sensing materials, researchers have been engaged in constructing diverse morphologies with high specific surface areas. Finely designed and prepared three-dimensional (3D) shapes such as hollow structures, nanoflowers, mesh-like and sponge-like structures, etc., could maintain a high specific surface area while also exhibiting excellent mechanical properties and stability. The hollowed-out Fe2O3-loaded NiO heterojunction nanorods assembled by porous ultrathin nanosheets (Figure 8a,b) prepared by Li et al. showed superior humidity adaptability during the process of ethanol monitoring. They exhibited an outstanding response of 51.2 toward 10 ppm of ethanol at 80% relative humidity (RH) (Figure 8c), along with exceptional selectivity for ethanol (Figure 8d) under conditions of high relative humidity. This was achieved at a lower operating temperature of 150 °C and with a low detection limit of 0.5 ppb. The excellent selectivity could be attributed to the favorable specificity of the NiO material towards ethanol and the hollowed-out morphology, which allowed massive ethanol gas molecules to adhere to the water molecules to spread into the inside of the material. Moreover, Fe2O3@NiO had a higher content of adsorbed oxygen than its components, which was also beneficial to gas-sensing reactions [98].
Moreover, the porous structure is also well-designed, since it not only presents a large specific surface area, but also owns rich channels to facilitate species transport. To acquire porous heterostructures, one of the most extensively employed approaches is to use porous precursors, such as metal–organic frameworks (MOFs) [105,106,107,108,109,110,111,112]. MOFs contain metal centers, which are linked by organic ligands. MOFs feature high porosity and specific surface areas, and their pore size and structure can be regulated by adjusting the types of metal ions and organic ligands and synthesis conditions. When MOFs are calcined at an appropriate temperature, oxides can be obtained, which can inherit the porous structure with rich channels from MOFs. For example, the commonly used n-ZnO, n-type α-Fe2O3, p-Co3O4, and p-CuO can be obtained from the calcination of zeolitic imidazolate frameworks (ZIF) ZIF-8, ZIF-67, MIL-101(Fe), and HKUST-1 [55,113]. Doan et al. constructed n-ZnO/p-Co3O4 heterojunctions derived from ZIF-8/ZIF-67 for the detection of low concentrations of ethanol. The assembled sensor showed great ethanol selectivity in eight interfering gases and a low detection limit [113]. Li et al. constructed Co3O4-TiO2 porous heterojunction nanosheets by calcining ZIF-67/MIL-125 precursors. The sensing results showed that the fabricated Co3O4-TiO2 porous nanosheets exhibited superior ethanol-sensing performances compared to MOF-derived TiO2 nanotablets, owing to the formation of p–n heterojunctions and unique porous nanosheet nanostructures, which possess a high specific surface area [114].

4. N-n Heterojunction Sensors for VOC Detection

4.1. N-n Heterojunction Sensors for Acetone Detection

In addition to the construction of p-n heterojunction, a few works suggested that the fabrication of n-n heterojunction composites could also improve the sensing properties. The n-n heterojunctions can modulate the energy band and carrier concentration to enhance sensing performances.
Using MOSs as precursors can not only prepare the above-mentioned p-n materials, but also n-n materials. Zhang et al. fabricated a hollow polyhedral SnO2-ZnO composite from MOFs (Figure 9a) [115]. The optimized 6% SnO2-ZnO NPs showed remarkable acetone sensing with a 0.82 ppb detection limit and linear response up to 10 ppm [115]. ZnO-SnO2 heterojunction inverse opal photonic bandgap materials, characterized by a regular three-dimensional ordered macroporous structure, were successfully synthesized through a straightforward spray drying method followed by an impregnation–annealing process. The uniformly distributed ZnO and SnO2 formed the connected skeleton. The special structure resulted in a five-times-higher response (Ra/Rg) of IOPBs (40.3 to 50 ppm acetone) in comparison to the performance of the pure SnO2 sensor [69] (Figure 9c). He et al. synthesized porous ZnFe2O4/SnO2 core–shell spheres for high-performance acetone by combining SnO2 nanospheres with ZnFe2O4. The optimal porous ZFO/SNO composite exhibited a high response value of 120 to 100 ppm of acetone at a temperature of 210 °C, accompanied by excellent stability. Furthermore, its detection limit could reach as low as 0.1 ppm, which was attributed to its abundant heterojunctions, porous structure, and small nanoparticle size [92].
As discussed in Section 3.2, in addition to being additive in p-n type heterojunctions, MXene can also act as the support for n-n type heterojunctions. Su et al. fabricated 1D ZnSnO3/ZnO nanofibers via the electrospinning technique and subsequently attached these nanofibers to 2D Ti3C2Tx MXene sheets using electrostatic self-assembly technology, resulting in the formation of 1D/2D heterostructured ZnSnO3/ZnO/Ti3C2Tx MXene composites. The accordion-like MXene slices were uniformly and tightly cross-covered with ZnSnO3/ZnO nanofibers, forming a strong bond. As n-type semiconductor materials, ZnO and ZnSnO3 established an n-n heterojunction, which subsequently formed a p-n heterojunction between the ZnSnO3/ZnO component and the p-type semiconductor material Ti3C2Tx MXene within the ZnSnO3/ZnO/Ti3C2Tx MXene composite. The successful construction of multiple heterojunctions and the synergistic interactions among these three components facilitated real-time and efficient detection of acetone gas. Compared to the ZnSnO3/ZnO samples, the ZnSnO3/ZnO/Ti3C2Tx MXene composite exhibited a response value of 15.68 in 100 ppm acetone gas, representing a 3.5-fold increase. Additionally, the optimal operating temperature was notably decreased to 120 °C [54].

4.2. N-n Heterojunction Sensors for Ethanol Detection

Compared to single-oxide sensing materials, MOS composites typically demonstrate superior sensitivity, improved stability, and a reduced detection limit. Zhang et al. developed hierarchical ZnO-In2O3 heterostructures assembled from nanosheets for the purpose of ethanol detection [116]. Yan et al. synthesized ZnO-In2O3 porous nanosheets for the application of ethanol sensing; it is noteworthy that the optimal operating temperature required for these nanosheets was relatively high (350 °C) [117]. Recently, Jiang et al. successfully designed and synthesized a distinctive hierarchical In2O3-ZnO compound through a hydrothermal calcination process, which exhibited high reproducibility for the detection of ethanol (Figure 10). The gas sensor containing 15 at. % In2O3-ZnO achieved a lower detection limit for ethanol (200 ppb) and a reduced baseline resistance (~1 Mohm at 225 °C) compared to pure ZnO. The improved moisture tolerance suggested that an appropriate doping concentration of In3+ in ZnO could effectively enhance gas sensing performance [50]. This enhancement is primarily attributed to the smaller size of In2O3 nanoparticle clusters, which reduce baseline resistance via electron transfer and lower the detection limit by optimizing humidity tolerance. Additionally, Gao et al. linked the improved gas sensing performance of ZnO@In2O3-2 heterostructures to their unique heterogeneous structure construction, abundant Ovs, substantial specific surface area, and internal electric field [100].
Similar results have also been observed in other n-n MOSs heterojunctions, such as MoO3 (1D) @SnO2 (2D) core–shell heterostructures [56], an irregular particle-shaped SnO2/In2O3 composite [101], and In2O3/Fe2O3 shuttle-like structures [118]. Wang et al. found that an ethanol gas sensor using the optimal MoO3 (1D) @SnO2 (2D) core–shell heterostructures as the sensing materials exhibited a higher response than its components with 48.64 at the optimum operating temperature (200 °C) to 100 ppm ethanol. The enhanced gas sensing mechanism of MoO3@SnO2 can be attributed to its abundance of high-activity sites, narrow band gap, and rich concentration of surface vacancies [56]. Similarly, the formation of n-n heterojunctions, an augmentation in oxygen vacancies, and its distinctive structural characteristics also induced In2O3/Fe2O3 to show excellent ethanol sensing performances [118]. The response to 100 ppm of ethanol gas achieved a value of 67.5 at an optimal operating temperature of 200 °C, while the response/recovery time was 9 s/236 s [118].
In addition to the commonly used n-type MOSs, such as ZnO, SnO2, In2O3, etc., some other novel sensing materials have also been explored. Spinel oxides with a formula of AB2O4 show great potential for gas sensing application [119]. Zn2SnO4 is a promising candidate due to its heightened chemical sensitivity, minimal visible light absorption, and outstanding optical electronic properties. Compared with the nanoparticles of SnO2/Zn2SnO4, Zn2SnO4 and SnO2, hierarchical porous SnO2/Zn2SnO4 nanospheres, synthesized via a one-step hydrothermal method, exhibited superior gas sensing properties towards ethanol. At the optimal operating temperature of 250 °C, the sensor fabricated from these porous SnO2/Zn2SnO4 nanospheres demonstrated a peak response of 30.5 to 100 ppm of ethanol [120]. Recently, Ma et al. fabricated hollow Zn2SnO4/SnO2 nanocubes with the n-n heterojunction between Zn2SnO4 and SnO2 grains and found that the Zn2SnO4/SnO2 sensor possessed good responses (Ra/Rg) toward ethanol and formaldehyde compared to other VOC gases [121]. Following these works, superior ethanol-sensing performances were achieved by the in situ synthesis of a porous Zn2SnO4/CdSnO3-nanocube-based n-n heterostructure [57]. Taking Zn2SnO4/CdSnO3 nanocubes as an example, the sensing mechanism is as follows: CdSnO3 and Zn2SnO4 are ubiquitous n-type semiconductor perovskite materials, with electrons serving as the primary charge carriers. Consequently, the oxygen adsorption/desorption model theory can be employed to elucidate the underlying mechanisms of their gas=sensing mechanism for Zn2SnO4/CdSnO3, through Equations (1)–(4), (7), and (8). Firstly, the pore volume and specific surface area of Zn2SnO4/CdSnO3 were higher than those of Zn2SnO4 and CdSnO3, which facilitated the adsorption of reactive oxygen species from the atmosphere, enabling their participation in the gas-sensing reaction and enhancing the overall gas sensing performance. Secondly, the abundant Ovs in the Zn2SnO4/CdSnO3 heterostructure promoted oxygen adsorption, resulting in a thicker oxygen space charge layer and a more pronounced electron depletion layer and electron depletion layer based on the XPS and electron paramagnetic resonance (EPR) spectra. As depicted in Figure 11b, the intensity of the Zn2SnO4/CdSnO3 was much higher than those of the other two samples, revealing the highest content of Ovs in Zn2SnO4/CdSnO3. In n-type MOSs, oxygen vacancies act as electron donors, which is favorable for the adsorption of oxygen molecules onto the surface and the subsequent formation of a space charge layer. Zn2SnO4/CdSnO3 has higher Ovs than pure CdSnO3, causing greater electron depletion during oxygen adsorption, leading to higher Ra and Ra/Rg. The n-n heterojunction between Zn2SnO4 and CdSnO3 induces electron transfer due to work function differences, creating an electron depletion layer on Zn2SnO4. This increases CdSnO3 carrier concentration, facilitating oxygen adsorption and oxygen species space charge generation [57].
Note that in the above work, the annealing temperature of the CdSn(OH)6 precursor was 700 °C and produced Zn2SnO4. When the annealing temperature decreased to 500 °C, it would produce ZnSnO3 [57]. This material is also a good candidate for ethanol sensing [122]. In addition to obtaining ZnSnO3 through calcination, it can also be fabricated through other methods, such as magnetic sputtering. Cao et al. prepared a single-layer SiO2 microsphere template using the self-assembly method, and used this template to prepare a porous thin film of SnO2-and-ZnSnO3 composite by magnetron sputtering. Next, a double-layer porous-film ethanol gas sensor was prepared, and a detection limit of 50 ppm was reached at 290 °C, with a response value of 11.5 [123].
To overcome the limit of the high electron–hole pair recombination rate of n-type MOSs, which restricts their performances in gas sensors, the incorporation of additives, such as MXene and polyoxometalates (POMs), into the fabricated heterojunctions has been demonstrated to be an effective strategy. POMs, which are metal–oxygen cluster ions rich in transition elements like W, Mo, and V within a specific skeletal structure, can function as electron acceptors. They participate in the synthesis of composite materials to acquire and store electrons, thereby enhancing the gas-sensing response [97]. Taking into account the distinct energy levels of In2O3 and SnO2, along with the electron capture capabilities of POMs [97], Zhang et al. designed and successfully synthesized In2O3@PW12@SnO2 core–middle-shell nanofibers (NFs) through a one-step coaxial electrospinning technique. The interfaces between In2O3/PW12 and PW12/SnO2 are arranged in a tandem fashion, resulting in the formation of 1D continuously distributed tandem heterojunctions (Figure 12b). The addition of an appropriate amount of PW12 electron acceptor and the tandem heterogeneous interfaces between 1D POM and MOS accelerated electrons transferred and reduced the recombination of electron–hole pairs. Consequently, the ethanol sensing performances were enhanced. The optimal sensitivity response to 100 ppm of ethanol gas was approximately 4-fold higher than that observed in the control samples (In2O3@SnO2 and In2O3/3%PW12/SnO2 sensors) (Figure 12c) [97].
On the other hand, MXene can act as an electron doner in n-type composites. Mei et al. discovered that the work function of SnO2 exceeds that of Ti3C2Tx MXene, resulting in the transfer of electrons from Ti3C2Tx MXene to SnO2 upon contact. This electron transfer induced a downward bending of the SnO2 band until Fermi level equilibrium was achieved. In the process, Schottky junctions were formed at the interface, and the Ti3C2Tx MXene became electron depletion centers [103]. The assembled sensor based on the n-type SnO2/ZnO/Ti3C2Tx MXene composites with the optimal Ti3C2Tx quantity of 3 wt% exhibited a superior response value of 121.1, a faster response time of 3 s, and favorable selectivity characteristics, as compared to pure Ti3C2Tx or SnO2/ZnO nanomaterials (Figure 13b,c). Note that the SnO2/ZnO/Ti3C2Tx MXene composite also presented a response toward acetone, but the response was much lower under the same detection conditions. However, the syntheses of In2O3@PW12@SnO2 and SnO2/ZnO/Ti3C2Tx both required multiple steps (Figure 12a and Figure 13a).

5. Conclusions and Outlook

As summarized and extensively discussed in this review, a substantial number of studies have been conducted on the detection of acetone and ethanol by various heterojunction materials.
To enhance the sensing performance of MOS-heterojunction-based sensing materials, various strengthening methods have been proposed, including doping (Ag, Pd, etc.), the incorporation of additives (MXene, POMs, etc.), novel morphology design, and self-doping via oxygen vacancies. The content of OV was strongly related to the sensing performances. The heterojunction MOSs with higher OV contents commonly presented improved resistant ability. The aforementioned strategies have significantly enhanced the sensing capabilities of MOS heterojunction composites. Nevertheless, improving the selectivity of MOS sensors remains a critical challenge that constrains their practical applications. Specifically, differentiating between ethanol and acetone poses difficulties due to their analogous chemical properties. To date, only limited research has addressed this particular issue [88].
To enhance sensitivity, the integration of multiple sensors is employed through sensor array technology [4,124]. This approach amalgamates the response patterns from various sensors to improve sensitivity toward low-concentration target compounds. Simultaneously, it enhances selectivity for complex VOCs and diminishes responses to interfering substances. Furthermore, due to its rapid response time, this technology facilitates real-time monitoring [125].
Additionally, in order to fabricate MOS heterojunctions with superior performances, current synthesis methods predominantly involve multi-step processes that allow for precise control over product structure and morphology. However, these intricate reaction steps complicate the regulation of reaction parameters. Consequently, developing a universal method that is easy to operate, controllable, and suitable for large-scale production presents a significant challenge that must be addressed for the industrial application of such sensors. The implementation of data-driven material design methodologies and the optimization of synthesis parameters may serve as crucial strategies in overcoming this issue [126,127,128,129,130].
In summary, MOS heterojunction materials utilized as gas sensors are expected to evolve in the following directions: enhancing sensitivity and selectivity while minimizing operating temperatures; advancing intelligence and integration; achieving miniaturization and portability; ensuring environmental protection and sustainability; and broadening their application domains in the future. These developments will propel continuous advancements and innovations within gas sensor technology, providing reliable and efficient monitoring solutions across fields such as environmental protection, medical health care, and industrial safety.

Author Contributions

Conceptualization, X.S. and S.X.; validation, H.Z, P.W., H.Y. and M.Z.; formal analysis, S.Z., P.W., H.Y. and M.Z.; investigation, S.Z., M.Z., H.Z., P.W. and H.Y.; resources, S.Z., P.W., H.Y., M.Z. and X.S.; data curation, S.Z., H.Z., P.W. and H.Y.; writing—original draft preparation, S.Z.; writing—review and editing, X.S. and S.X.; supervision, X.S. and S.X.; project administration, X.S.; funding acquisition, X.S. and S.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Shanghai (23ZR1425500) and Class III Peak Discipline of Shanghai—Materials Science and Engineering (High-Energy Beam Intelligent Processing and Green Manufacturing).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article material. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors are very grateful to Shanghai Science and Technology Commission for the support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Belluomo, I.; Boshier, P.R.; Myridakis, A.; Vadhwana, B.; Markar, S.R.; Spanel, P.; Hanna, G.B. Selected Ion Flow Tube Mass Spectrometry for Targeted Analysis of Volatile Organic Compounds in Human Breath. Nat. Protoc. 2021, 16, 3419–3438. [Google Scholar] [CrossRef]
  2. Nah, S.H.; Kim, J.B.; Chui, H.N.T.; Suh, Y.; Yang, S. Enhanced Colorimetric Detection of Volatile Organic Compounds Using a Dye-Incorporated Photonic Crystal-Based Sensor Array. Adv. Mater. 2024, 36, 2409297. [Google Scholar] [CrossRef]
  3. Kaur, N.; Zappa, D.; Maraloiu, V.-A.; Comini, E. Novel Christmas Branched like NiO/NiWO4/WO3 (p–p–n) Nanowire Heterostructures for Chemical Sensing. Adv. Funct. Mater. 2021, 31, 2104416. [Google Scholar] [CrossRef]
  4. Chen, M.; Zhang, M.; Yang, Z.; Zhou, C.; Cui, D.; Haick, H.; Tang, N. AI-Driven Wearable Mask-Inspired Self-Healing Sensor Array for Detection and Identification of Volatile Organic Compounds. Adv. Funct. Mater. 2024, 34, 2309732. [Google Scholar] [CrossRef]
  5. Yang, X.Y.; Deng, Y.; Yang, H.T.; Liao, Y.Z.; Cheng, X.W.; Zou, Y.D.; Wu, L.M.; Deng, Y.H. Functionalization of Mesoporous Semiconductor Metal Oxides for Gas Sensing: Recent Advances and Emerging Challenges. Adv. Sci. 2023, 10, 2204810. [Google Scholar] [CrossRef]
  6. Kamathe, V.; Nagar, R. Designing Sensing Material with Fractal Geometry: A Gateway for Enhanced Ethanol Sensing under Ambient Conditions. Sens. Actuators B Chem. 2024, 410, 135642. [Google Scholar] [CrossRef]
  7. Khatib, M.; Haick, H. Sensors for Volatile Organic Compounds. ACS Nano 2022, 16, 7080–7115. [Google Scholar] [CrossRef]
  8. Zhao, Y.; Liu, Y.X.; Han, B.; Wang, M.Y.; Wang, Q.; Zhang, Y.N. Fiber Optic Volatile Organic Compound Gas Sensors: A Review. Coord. Chem. Rev. 2023, 493, 215297. [Google Scholar] [CrossRef]
  9. Gaur, S.; Singh, S.; Deb, J.; Bhutani, V.; Mondal, R.; Pareek, V.; Gupta, R. Site-Selective MoS2-Based Sensor for Detection and Discrimination of Triethylamine from Volatile Amines Using Kinetic Analysis and Machine Learning. Adv. Funct. Mater. 2024, 34, 2405232. [Google Scholar] [CrossRef]
  10. Singh, S.; Sajana, S.; Varma, P.; Sreelekha, G.; Adak, C.; Shukla, R.P.; Kamble, V.B. Metal Oxide-Based Gas Sensor Array for VOCs Determination in Complex Mixtures Using Machine Learning. Microchim. Acta 2024, 191, 196. [Google Scholar] [CrossRef]
  11. Kulkarni, S.; Ghosh, R. CuO-ZnO p-n Junctions for Accurate Prediction of Multiple Volatile Organic Compounds Aided by Machine Learning Algorithms. Anal. Chim. Acta 2023, 1253, 341084. [Google Scholar] [CrossRef]
  12. Khan, M.; Abid, K.; Ferlazzo, A.; Bressi, V.; Espro, C.; Hussain, M.; Foti, A.; Gucciardi, P.G.; Neri, G. A Sensitive and Selective Non-Enzymatic Dopamine Sensor Based on Nanostructured Co3O4-Fe2O3 Heterojunctions. Chemosensors 2023, 11, 379. [Google Scholar] [CrossRef]
  13. Zhang, X.; Sun, C.; Xu, S.; Huang, M.; Wen, Y.; Shi, X.-R. DFT-Assisted Rational Design of CoMxP/CC (M = Fe, Mn, and Ni) as Efficient Electrocatalyst for Wide PH Range Hydrogen Evolution and Oxygen Evolution. Nano Res. 2022, 15, 8897–8907. [Google Scholar] [CrossRef]
  14. Mohammadzadeh, M.R.; Hasani, A.; Hussain, T.; Ghanbari, H.; Fawzy, M.; Abnavi, A.; Ahmadi, R.; Kabir, F.; De Silva, T.; Rajapakse, R.K.N.D.; et al. Enhanced Sensitivity in Photovoltaic 2D MoS2/Te Heterojunction VOC Sensors. Small 2024. [Google Scholar] [CrossRef]
  15. Maruthapandi, M.; Durairaj, A.; Saravanan, A.; Luong, J.H.T.; Bakandritsos, A.; Gedanken, A.; Zboril, R. The Innovative Design of Carbon Dots on Polymer Texture for Highly Selective Detection of Amino Compounds. Carbon 2024, 228, 119414. [Google Scholar] [CrossRef]
  16. Mohammadzadeh, M.R.; Hasani, A.; Jaferzadeh, K.; Fawzy, M.; De Silva, T.; Abnavi, A.; Ahmadi, R.; Ghanbari, H.; Askar, A.; Kabir, F.; et al. Unique Photoactivated Time-Resolved Response in 2D GeS for Selective Detection of Volatile Organic Compounds. Adv. Sci. 2023, 10, 2205458. [Google Scholar] [CrossRef]
  17. Capman, N.S.S.; Zhen, X.V.; Nelson, J.T.; Chaganti, V.R.S.K.; Finc, R.C.; Lyden, M.J.; Williams, T.L.; Freking, M.; Sherwood, G.J.; Bühlmann, P.; et al. Machine Learning-Based Rapid Detection of Volatile Organic Compounds in a Graphene Electronic Nose. ACS Nano 2022, 16, 19567–19583. [Google Scholar] [CrossRef]
  18. Gong, Y.; Wu, X.; Li, X.; Wang, A.; Zhang, M.; Chen, Y. Enhanced Acetone Sensing Properties of Pt@Al-Doped ZnO Core-Shell Nanoparticles. Sens. Actuators B Chem. 2021, 329, 129153. [Google Scholar] [CrossRef]
  19. Zhang, J.; Jia, X.; Liu, T.; Yang, J.; Wang, S.; Li, Y.; Shao, D.; Feng, L.; Song, H. Facile Strategy to Synthesize Porous GO/ZnO Heterostructure for Enhanced Acetone Gas Sensing Properties. Sens. Actuators B Chem. 2022, 359, 131601. [Google Scholar] [CrossRef]
  20. Nemufulwi, M.I.; Swart, H.C.; Shingange, K.; Mhlongo, G.H. ZnO/ZnFe2O4 Heterostructure for Conductometric Acetone Gas Sensors. Sens. Actuators B Chem. 2023, 377, 133027. [Google Scholar] [CrossRef]
  21. Park, J.Y.; Kwak, Y.; Lim, H.-R.; Park, S.-W.; Lim, M.S.; Cho, H.-B.; Myung, N.V.; Choa, Y.-H. Tuning the Sensing Responses towards Room-Temperature Hypersensitive Methanol Gas Sensor Using Exfoliated Graphene-Enhanced ZnO Quantum Dot Nanostructures. J. Hazard. Mater. 2022, 438, 129412. [Google Scholar] [CrossRef]
  22. Zhao, L.J.; Sun, S.P.; Zhang, S.; Liu, Z.; Huang, B.Y.; Wang, N.; Zhang, J.W.; Li, X.G. UV-Activated Hollow ZnO@TiO2 Heterostructured Nanaospheres for Detecting Formaldehyde at Room Temperature. Sens. Actuators B-Chem. 2023, 394, 134306. [Google Scholar] [CrossRef]
  23. Amouzesh, S.P.; Khodadadi, A.A.; Mortazavi, Y.; Saris, S.; Asgari, M. MIL-100(Fe) /ZnO Nanocomposite Sensors: An Enhanced Ammonia Selectivity and Low Operating Temperature. Sens. Actuators B-Chem. 2024, 399, 134791. [Google Scholar] [CrossRef]
  24. Umar, A.; Ibrahim, A.A.; Nakate, U.T.; Albargi, H.; Alsaiari, M.A.; Ahmed, F.; Alharthi, F.A.; Ali Alghamdi, A.; Al-Zaqri, N. Fabrication and Characterization of CuO Nanoplates Based Sensor Device for Ethanol Gas Sensing Application. Chem. Phys. Lett. 2021, 763, 138204. [Google Scholar] [CrossRef]
  25. Hermawan, A.; Zhang, B.; Taufik, A.; Asakura, Y.; Hasegawa, T.; Zhu, J.; Shi, P.; Yin, S. CuO Nanoparticles/Ti3C2Tx MXene Hybrid Nanocomposites for Detection of Toluene Gas. ACS Appl. Nano Mater. 2020, 3, 4755–4766. [Google Scholar] [CrossRef]
  26. Guo, M.M.; Luo, N.; Chen, Y.; Fan, Y.; Wang, X.H.; Xu, J.Q. Fast-Response MEMS Xylene Gas Sensor Based on CuO/WO3 Hierarchical Structure. J. Hazard. Mater. 2022, 429, 127471. [Google Scholar] [CrossRef]
  27. Wu, W.; Long, J.Y.; Guo, Y.J.; Zu, X.T.; Li, S.; Xiang, X. P-CuO/n-TiO2 Heterojunction Nanostructure-Based Surface Acoustic Wave Sensor with Strong Electric Loading Effect for Highly Sensitive H2 S Gas Sensing. Sens. Actuators B-Chem. 2023, 394, 134380. [Google Scholar] [CrossRef]
  28. Govind, A.; Bharathi, P.; Mohan, M.K.; Archana, J.; Harish, S.; Navaneethan, M. Highly Sensitive near Room Temperature Operable NO2 Gas-Sensor for Enhanced Selectivity via Nanoporous CuO@ZnO Heterostructures. J. Environ. Chem. Eng. 2023, 11, 110056. [Google Scholar] [CrossRef]
  29. Seekaew, Y.; Kamlue, S.; Wongchoosuk, C. Room-Temperature Ammonia Gas Sensor Based on Ti3C2Tx MXene/Graphene Oxide/CuO/ZnO Nanocomposite. ACS Appl. Nano Mater. 2023, 6, 9008–9020. [Google Scholar] [CrossRef]
  30. Su, C.; Li, M.Y.; Zhang, Y.F.; Liu, T.Q.; Ren, C.; Li, P.P.; Yin, X.Q.; Zhang, L.; Zhang, M.; Wu, W.W. Boosting Ethylene Glycol Sensing Performance with Dendritic Hierarchical CuO/Co3O4 Heterojunction Nanowire. ACS Appl. Nano Mater. 2023, 6, 19249–19256. [Google Scholar] [CrossRef]
  31. Wang, J.H.; Zhang, D.Z.; Gao, Y.H.; Chen, F.J.; Wang, T.; Xia, H.; Sui, X.X.; Wang, Z.H. Fast-Response Hydrogen Sulfide Gas Sensor Based on Electrospinning Co3O4 Nanofibers-Modified CuO Nanoflowers: Experimental and DFT Calculation. Sens. Actuators B-Chem. 2023, 396, 134579. [Google Scholar] [CrossRef]
  32. Oosthuizen, D.N.; Weber, I.C. A Strategy to Enhance Humidity Robustness of P-Type CuO Sensors for Breath Acetone Quantification. Small Sci. 2023, 3, 2200096. [Google Scholar] [CrossRef]
  33. Wang, Y.; Li, H.; Huang, D.; Wang, X.; Cai, L.; Chen, Y.; Wang, W.; Song, Y.; Han, G.; Zhen, B. A High-Performance Ethanol Gas Sensor Based on Ce-Doped SnO2 Nanomaterials Prepared by the Pechini Method. Mater. Sci. Semicond. Process 2022, 137, 106188. [Google Scholar] [CrossRef]
  34. Chen, L.; Song, Y.; Liu, W.; Dong, H.; Wang, D.; Liu, J.; Liu, Q.; Chen, X. MOF-Based Nanoscale Pt Catalyst Decorated SnO2 Porous Nanofibers for Acetone Gas Detection. J. Alloys Compd. 2022, 893, 162322. [Google Scholar] [CrossRef]
  35. Li, C.; Kim, K.; Fuchigami, T.; Asaka, T.; Kakimoto, K.; Masuda, Y. Acetone Gas Sensor Based on Nb2O5@SnO2 Hybrid Structure with High Selectivity and Ppt-Level Sensitivity. Sens. Actuators B Chem. 2023, 393, 134144. [Google Scholar] [CrossRef]
  36. Hu, J.; Xiong, X.; Guan, W.; Tan, C. Hollow Mesoporous SnO2/Zn2 SnO 4 Heterojunction and RGO Decoration for High-Performance Detection of Acetone. ACS Appl. Mater. Interfaces 2022, 14, 55249–55263. [Google Scholar] [CrossRef]
  37. Yang, Z.; Cao, W.; Peng, C.; Wang, T.; Li, B.; Ma, H.; Su, Y.; Zhou, Z.; Yang, J.; Zeng, M. Construction, Application and Verification of a Novel Formaldehyde Gas Sensor System Based on Ni-Doped SnO2 Nanoparticles. IEEE Sens. J. 2021, 21, 11023–11030. [Google Scholar] [CrossRef]
  38. Zou, Y.H.; Chen, S.; Sun, J.; Liu, J.Q.; Che, Y.K.; Liu, X.H.; Zhang, J.; Yang, D.J. Highly Efficient Gas Sensor Using a Hollow SnO2 Microfiber for Triethylamine Detection. ACS Sens. 2017, 2, 897–902. [Google Scholar] [CrossRef]
  39. 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]
  40. Zhou, Y.; Hu, Z.H.; Zhao, H.C.; Wang, Y.J.; Li, J.; Zou, C. Two-Dimensional Black Phosphorus/Tin Oxide Heterojunctions for High-Performance Chemiresistive H2S Sensing. Anal. Chim. Acta 2023, 1245, 340825. [Google Scholar] [CrossRef]
  41. Su, C.; Zhang, L.; Han, Y.; Ren, C.; Chen, X.; Hu, J.; Zeng, M.; Hu, N.; Su, Y.; Zhou, Z.; et al. Controllable Synthesis of Crescent-Shaped Porous NiO Nanoplates for Conductometric Ethanol Gas Sensors. Sens. Actuators B Chem. 2019, 296, 126642. [Google Scholar] [CrossRef]
  42. Wang, Z.; Zhang, K.; Fei, T.; Gu, F.; Han, D. α-Fe2O3/NiO Heterojunction Nanorods with Enhanced Gas Sensing Performance for Acetone. Sens. Actuators B Chem. 2020, 318, 128191. [Google Scholar] [CrossRef]
  43. Chen, Q.; Zhang, Y.; Ma, S.; Wang, Y.; Wang, P.; Zhang, G.; Gengzang, D.; Jiao, H.; Wang, M.; Chen, W. Multishelled NiO/NiCo2O4 Hollow Microspheres Derived from Bimetal-Organic Frameworks as High-Performance Sensing Material for Acetone Detection. J. Hazard. Mater. 2021, 415, 125662. [Google Scholar] [CrossRef]
  44. Li, C.; Choi, P.G.; Kim, K.; Masuda, Y. High Performance Acetone Gas Sensor Based on Ultrathin Porous NiO Nanosheet. Sens. Actuators B Chem. 2022, 367, 132143. [Google Scholar] [CrossRef]
  45. Liu, H.; Wang, Z.; Cao, G.; Pan, G.; Yang, X.; Qiu, M.; Sun, C.; Shao, J.; Li, Z.; Zhang, H. Construction of Hollow NiO/ZnO p-n Heterostructure for Ultrahigh Performance Toluene Gas Sensor. Mater. Sci. Semicond. Process 2022, 141, 106435. [Google Scholar] [CrossRef]
  46. Nakate, U.T.; Yu, Y.T.; Park, S. High Performance Acetaldehyde Gas Sensor Based on P-n Heterojunction Interface of NiO Nanosheets and WO3 Nanorods. Sens. Actuators B Chem. 2021, 344, 130264. [Google Scholar] [CrossRef]
  47. Wang, W.; Li, F.; Zhang, N.; Liu, C.; Zhou, J.; Liu, D.; Ruan, S. Self-Assembled Co3O4@WO3 Hollow Microspheres with Oxygen Vacancy Defects for Fast and Selective Detection of Toluene. Sens. Actuators B Chem. 2022, 351, 130931. [Google Scholar] [CrossRef]
  48. Liu, H.; Wang, Z.; Sun, C.; Shao, J.; Li, Z.; Zhang, H.; Qiu, M.; Pan, G.; Yang, X. Construction of Co3O4/Fe3O4 Heterojunctions from Metal Organic Framework Derivatives for High Performance Toluene Sensor. Sens. Actuators B Chem. 2023, 375, 132863. [Google Scholar] [CrossRef]
  49. Yu, S.; Lin, H.; Liu, X.; Zhong, J.; Hu, L.; Lu, F.; Hu, Z.; Gao, W. Enhancing the Photoelectric Performance of TiO2-Based Heterojunction Using the Hydrothermal Decomposition of MIL-125 Strategy for Sensitive Detection of Bisphenol A. Sens. Actuators B Chem. 2024, 413, 135881. [Google Scholar] [CrossRef]
  50. Jiang, B.; Zhou, T.; Zhang, L.; Han, W.; Yang, J.; Wang, C.; Sun, Y.; Liu, F.; Sun, P.; Lu, G. Construction of Mesoporous In2O3-ZnO Hierarchical Structure Gas Sensor for Ethanol Detection. Sens. Actuators B Chem. 2023, 393, 134203. [Google Scholar] [CrossRef]
  51. Nguyen, T.T.; Patel, M.; Kim, J. Self-Powered Transparent Photodetectors for Broadband Applications. Surf. Interfaces 2021, 23, 100934. [Google Scholar] [CrossRef]
  52. 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. Nanomicro Lett. 2023, 15, 89. [Google Scholar] [CrossRef] [PubMed]
  53. Li, C.; Zhou, H.; Yang, S.; Wei, L.; Han, Z.; Zhang, Y.; Pan, H. Preadsorption of O2 on the Exposed (001) Facets of ZnO Nanostructures for Enhanced Sensing of Gaseous Acetone. ACS Appl. Nano Mater. 2019, 2, 6144–6151. [Google Scholar] [CrossRef]
  54. Sun, B.T.; Ding, Y.L.; Wang, Q.; Song, P. Rational Design of 1D/2D Heterostructured ZnSnO3/ZnO/Ti3C2Tx MXene Nanocomposites for Enhanced Acetone Gas Sensing Performance. Sens. Actuators B-Chem. 2024, 409, 135541. [Google Scholar] [CrossRef]
  55. Wu, J.; Zheng, Z.; Chi, H.; Jiang, J.; Zhu, L.; Ye, Z. Ultrasensitive and Exclusive Chemiresistors with a ZIF-67-Derived Oxide Cage/Nanofiber Co3O4/In2O3 Heterostructure for Acetone Detection. ACS Appl. Mater. Interfaces 2024, 16, 9126–9136. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, X.; Gao, Y.; Zhang, Q.; He, X.; Wang, X. Synthesis of MoO3 (1D) @SnO2 (2D) Core-Shell Heterostructures for Enhanced Ethanol Gas Sensing Performance. Sens. Actuators B Chem. 2023, 382, 133484. [Google Scholar] [CrossRef]
  57. Gao, Y.F.; Zhang, Q.B.; Mao, R.Z.; Duan, J.X.; Wang, H.Y.; Xu, G.G.; Wang, X.Z.; Xiong, Y.; Tian, J. Superior Ethanol Sensing Performance by In-Situ Construction of Porous Zn2SnO4/CdSnO3 Nanocubes n-n Heterostructure. Sens. Actuators B-Chem. 2024, 419, 136397. [Google Scholar] [CrossRef]
  58. Liu, H.Y.; Zhu, D.H.; Miao, T.T.; Liu, W.K.; Chen, J.; Cheng, B.; Qin, H.W.; Hu, J.F. Highly Sensitive P-SmFeO3/p-YFeO3 Planar-Electrode Sensor for Detection of Volatile Organic Compounds. Chemosensors 2023, 11, 187. [Google Scholar] [CrossRef]
  59. Zeng, C.; Lu, N.; Wen, Y.; Liu, G.; Zhang, R.; Zhang, J.; Wang, F.; Liu, X.; Li, Q.; Tang, Z.; et al. Engineering Nanozymes Using DNA for Catalytic Regulation. ACS Appl. Mater. Interfaces 2019, 11, 1790–1799. [Google Scholar] [CrossRef]
  60. Xiao, Y.; Gong, W.; Zhao, M.; Zhang, M.; Lu, N. Surface-Engineered Prussian Blue Nanozymes as Artificial Receptors for Universal Pattern Recognition of Metal Ions and Proteins. Sens. Actuators B Chem. 2023, 390, 134006. [Google Scholar] [CrossRef]
  61. Cao, F.; Shi, X.-R.; Wang, P.; Zhao, W.; Huang, M.; Hu, J.; Xu, S.; Zhao, G. Multistage Interface Engineered Cobalt Polysulfides Core-Shell Nanostructures for Dual Energy Storage Devices and Hydrogen Evolution. Vacuum 2023, 216, 112461. [Google Scholar] [CrossRef]
  62. Huang, M.; Yao, H.; Cao, F.; Wang, P.; Shi, X.-R.; Zhang, M.; Xu, S. Structural Engineering Evoked Multifunctionality in Molybdate Nanosheets for Industrial Oxygen Evolution and Dual Energy Storage Devices Inspired by Multi-Method Calculations. J. Colloid. Interface Sci. 2024, 676, 471–484. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, M.; Shi, X.-R.; He, X.; Zhang, X.; Cao, F.; Wang, P.; Sun, C.; Xu, S.; Zhang, M. Electronically Regulated FeOOH/c-NiMoO4 with Hierarchical Sandwich Structure as Efficient Electrode for Oxygen Evolution and Hybrid Supercapacitors. Electrochim. Acta 2022, 427, 140884. [Google Scholar] [CrossRef]
  64. Duan, Z.; Shi, X.-R.; Sun, C.; Lin, W.; Huang, S.; Zhang, X.; Huang, M.; Yang, Z.; Xu, S. Interface Engineered Hollow Co3O4@CoNi2S4 Nanostructure for High Efficiency Supercapacitor and Hydrogen Evolution. Electrochim. Acta 2022, 412, 140139. [Google Scholar] [CrossRef]
  65. Xu, S.R.; Wang, M.M.; Chen, C.P.; Feng, S.L. Sea Urchin-like SnO2/Alpha-Fe2O3 Heterostructural Microspheres for Enhanced Acetone Gas Sensing: Materials Preparation, Performance Evaluation, and Mechanism Investigation. Sens. Actuators B-Chem. 2023, 379, 133288. [Google Scholar] [CrossRef]
  66. Li, X.X.; Zhao, C.X.; Wang, Y.N.; Yuan, Z.Y. Precise Optical Fiber-Based Ammonia Sensor Using CdS Quantum Dots Decorated with ZnO at Heterointerface. Chemosensors 2024, 12, 169. [Google Scholar] [CrossRef]
  67. Zhang, R.; Lu, N.; Zhang, J.; Yan, R.; Li, J.; Wang, L.; Wang, N.; Lv, M.; Zhang, M. Ultrasensitive Aptamer-Based Protein Assays Based on One-Dimensional Core-Shell Nanozymes. Biosens. Bioelectron. 2020, 150, 111881. [Google Scholar] [CrossRef]
  68. Guo, M.; Jin, Z.; Pan, J.; Xu, J.; Guo, L.; Yin, X.-B.; Lu, N.; Zhang, M. Construction of COFs@MoS2-Pd Hierarchical Tubular Heterostructures for Enhanced Catalytic Performance. Inorg. Chem. 2024, 63, 18263–18275. [Google Scholar] [CrossRef]
  69. Li, F.; Jing, J.; Li, J.; Li, S.; Ye, S.; Song, X.; Zhan, Z.; Zhang, Y. Fabrication of ZnO-SnO2 Heterojunction Inverse Opal Photonic Balls for Chemiresistive Acetone Sensing. Sens. Actuators B Chem. 2024, 400, 134887. [Google Scholar] [CrossRef]
  70. Shi, X.-R.; Wang, P.; Jing, C.; Wu, K.; Xu, S.; Klötzer, B. The Formation of O Vacancy on ZrO2/Pd and Its Effect on Methane Dry Reforming: Insights from DFT and Microkinetic Modeling. Appl. Surf. Sci. 2023, 619, 156679. [Google Scholar] [CrossRef]
  71. Meng, C.; Zhao, G.; Shi, X.-R.; Chen, P.; Liu, Y.; Lu, Y. Oxygen-Deficient Metal Oxides Supported Nano-Intermetallic InNi3C0.5 toward Efficient CO2 Hydrogenation to Methanol. Sci. Adv. 2021, 7, eabi6012. [Google Scholar] [CrossRef] [PubMed]
  72. Si, J.; Zhao, G.; Sun, W.; Liu, J.; Guan, C.; Yang, Y.; Shi, X.-R.; Lu, Y. Oxidative Coupling of Methane: Examining the Inactivity of the MnOx-Na2WO4/SiO2 Catalyst at Low Temperature. Angew. Chem. Int. Ed. 2022, 61, e202117201. [Google Scholar] [CrossRef] [PubMed]
  73. Kaur, N.; Singh, M.; Casotto, A.; Arachchige, H.M.M.M.; Sangaletti, L.; Comini, E. Quenching of Oxygen-Related Defects in Graphene Oxide Nanohybrid: Highly Selective Room-Temperature Ethanol Sensor. Appl. Phys. Rev. 2022, 9. [Google Scholar] [CrossRef]
  74. Abraham, N.; Aseena, S. Dielectric Studies of CuO-ZnO Heterojunction Nanocomposites Synthesized by Co-Precipitation Method. Mater. Today Proc. 2021, 43, 3698–3700. [Google Scholar] [CrossRef]
  75. Shruthi, J.; Jayababu, N.; Ghosal, P.; Ramana Reddy, M.V. Ultrasensitive Sensor Based on Y2O3-In2O3 Nanocomposites for the Detection of Methanol at Room Temperature. Ceram. Int. 2019, 45, 21497–21504. [Google Scholar] [CrossRef]
  76. Li, K.; Wu, Y.; Chen, M.; Rong, Q.; Zhu, Z.; Liu, Q.; Zhang, J. High Methanol Gas-Sensing Performance of Sm2O3/ZnO/SmFeO3 Microspheres Synthesized Via a Hydrothermal Method. Nanoscale Res. Lett. 2019, 14, 57. [Google Scholar] [CrossRef]
  77. Cao, E.; Wu, L.; Zhang, Y.; Sun, L.; Yu, Z.; Nie, Z. Hydrothermal Synthesis of Cubic-Rhombohedral-In2O3 Microspheres with Superior Acetone Sensing Performance. Appl. Surf. Sci. 2023, 613, 156045. [Google Scholar] [CrossRef]
  78. Yang, W.; Ou, Q.; Yan, X.; Liu, L.; Liu, S.; Chen, H.; Liu, Y. High Sensing Performance Toward Acetone Vapor Using TiO2 Flower-Like Nanomaterials. Nanoscale Res. Lett. 2022, 17, 82. [Google Scholar] [CrossRef]
  79. Qu, F.; Thomas, T.; Zhang, B.; Zhou, X.; Zhang, S.; Ruan, S.; Yang, M. Self-Sacrificing Templated Formation of Co3O4/ZnCo2O4 Composite Hollow Nanostructures for Highly Sensitive Detecting Acetone Vapor. Sens. Actuators B Chem. 2018, 273, 1202–1210. [Google Scholar] [CrossRef]
  80. Wang, X.; Cao, R.; Zhang, S.; Hou, P.; Han, R.; Shao, M.; Xu, X. Hierarchical Flowerlike Metal/Metal Oxide Nanostructures Derived from Layered Double Hydroxides for Catalysis and Gas Sensing. J. Mater. Chem. A Mater. 2017, 5, 23999–24010. [Google Scholar] [CrossRef]
  81. Nie, L.; Fan, G.; Wang, A.; Zhang, L.; Guan, J.; Han, N.; Chen, Y. Finely Dispersed and Highly Toluene Sensitive NiO/NiGa2O4 Heterostructures Prepared from Layered Double Hydroxides Precursors. Sens. Actuators B Chem. 2021, 345, 130412. [Google Scholar] [CrossRef]
  82. Anajafi, Z.; Naseri, M.; Neri, G. Acetone Sensing Behavior of P-SmFeO3/n-ZnO Nanocomposite Synthesized by Thermal Treatment Method. Sens. Actuators B Chem. 2020, 304, 127252. [Google Scholar] [CrossRef]
  83. Qin, Y.; Liang, Y.; Zhou, C.; Bai, Y. Homogeneous Heterophase Junction Based on Cubic/Orthomorphic SnS for Chemiresistive Room Temperature Trace-Ethanol Sensor. Sens. Actuators B Chem. 2024, 404, 135285. [Google Scholar] [CrossRef]
  84. Shen, Z.H.; Zhu, Y.; Wu, L.M.; You, B.; Zi, J. Fabrication of Robust Crystal Balls from the Electrospray of Soft Polymer Spheres/Silica Dispersion. Langmuir 2010, 26, 6604–6609. [Google Scholar] [CrossRef]
  85. Zhao, L.; Jin, R.; Wang, C.; Wang, T.; Sun, Y.; Sun, P.; Lu, G. Flower-like ZnO-Co3O4 Heterojunction Composites for Enhanced Acetone Sensing. Sens. Actuators B Chem. 2023, 390, 133964. [Google Scholar] [CrossRef]
  86. Hu, J.; Xiong, X.; Guan, W.; Chen, Y.; Long, H. Regulation of O-Vacancy and Heterojunction Structure in MOF-Derived Fe2O3-Co3O4 Enhancing Acetone Sensing Performance. Sens. Actuators B Chem. 2024, 401, 135082. [Google Scholar] [CrossRef]
  87. Yang, L.; Qin, J.; Cui, S.; Liu, W. Silver Nanoparticle Functionalized Heterojunction NiO/SnO2 Nanotubes for Comprehensive Sensitization of Acetone Sensor. Sens. Actuators B Chem. 2024, 417, 136208. [Google Scholar] [CrossRef]
  88. Jiang, B.; Zhou, T.; Zhang, L.; Yang, J.; Han, W.; Sun, Y.; Liu, F.; Sun, P.; Zhang, H.; Lu, G. Separated Detection of Ethanol and Acetone Based on SnO2-ZnO Gas Sensor with Improved Humidity Tolerance. Sens. Actuators B Chem. 2023, 393, 134257. [Google Scholar] [CrossRef]
  89. Cheng, P.; Lv, L.; Wang, Y.; Zhang, B.; Zhang, Y.; Zhang, Y.; Lei, Z.; Xu, L. SnO2/ZnSnO3 Double-Shelled Hollow Microspheres Based High-Performance Acetone Gas Sensor. Sens. Actuators B Chem. 2021, 332, 129212. [Google Scholar] [CrossRef]
  90. Lee, J.E.; Lim, C.K.; Park, H.J.; Song, H.; Choi, S.-Y.; Lee, D.-S. ZnO–CuO Core-Hollow Cube Nanostructures for Highly Sensitive Acetone Gas Sensors at the Ppb Level. ACS Appl. Mater. Interfaces 2020, 12, 35688–35697. [Google Scholar] [CrossRef]
  91. Wang, C.; Wang, Y.; Cheng, P.; Xu, L.; Dang, F.; Wang, T.; Lei, Z. In-Situ Generated TiO2/α-Fe2O3 Heterojunction Arrays for Batch Manufacturing of Conductometric Acetone Gas Sensors. Sens. Actuators B Chem. 2021, 340, 129926. [Google Scholar] [CrossRef]
  92. He, L.; Hu, J.; Yuan, Q.; Xia, Z.; Jin, L.; Gao, H.; Fan, L.; Chu, X.; Meng, F. Synthesis of Porous ZnFe2O4/SnO2 Core-Shell Spheres for High-Performance Acetone Gas Sensing. Sens. Actuators B Chem. 2023, 378, 133123. [Google Scholar] [CrossRef]
  93. Yan, R.; Lu, N.; Han, S.; Lu, Z.; Xiao, Y.; Zhao, Z.; Zhang, M. Simultaneous Detection of Dual Biomarkers Using Hierarchical MoS2 Nanostructuring and Nano-Signal Amplification-Based Electrochemical Aptasensor toward Accurate Diagnosis of Prostate Cancer. Biosens. Bioelectron. 2022, 197, 113797. [Google Scholar] [CrossRef] [PubMed]
  94. Zhao, Z.; Shi, X.; Shen, Z.; Gu, Y.; He, L.; Zhang, M.; Lu, N. Single-Atom Fe Nanozymes Coupling with Atomic Clusters as Superior Oxidase Mimics for Ratiometric Fluorescence Detection. Chem. Eng. J. 2023, 469, 143923. [Google Scholar] [CrossRef]
  95. Liu, X.; Xu, J.; Cheng, Z.; Yang, J.; Li, Y. A Sensitive Acetone Sensor Based on WS2/WO3 Nanosheets with p-n Heterojunctions. ACS Appl. Nano Mater. 2022, 5, 12592–12599. [Google Scholar] [CrossRef]
  96. Chen, Y.; Li, H.; Ma, Q.; Che, Q.; Wang, J.; Wang, G.; Yang, P. ZIF-8 Derived Hexagonal-like α-Fe2O3/ZnO/Au Nanoplates with Tunable Surface Heterostructures for Superior Ethanol Gas-Sensing Performance. Appl. Surf. Sci. 2018, 439, 649–659. [Google Scholar] [CrossRef]
  97. Zhang, L.; Tian, J.; Wang, Y.; Wang, T.; Wei, M.; Li, F.; Li, D.; Yang, Y.; Yu, H.; Dong, X. Polyoxometalates Electron Acceptor-Intercalated In2O3@SnO2 Nanofibers for Chemiresistive Ethanol Gas Sensors. Sens. Actuators B Chem. 2024, 410, 135728. [Google Scholar] [CrossRef]
  98. Li, L.; Zhou, L.; Hu, Z.; Li, T.; Chen, B.; Li, H.Y.; Liu, H. Hollow-Out Fe2O3-Loaded NiO Heterojunction Nanorods Enable Real-Time Exhaled Ethanol Monitoring under High Humidity. ACS Appl. Mater. Interfaces 2023, 15, 15707–15720. [Google Scholar] [CrossRef]
  99. Wang, S.; Wang, X.; Qiao, G.; Chen, X.; Wang, X.; Cui, H. Core-Double Shell ZnO@In2O3@ZnO Hollow Microspheres for Superior Ethanol Gas Sensing. Sens. Actuators B Chem. 2021, 341, 130002. [Google Scholar] [CrossRef]
  100. Gao, Y.; Wang, X.; Zhang, Z.; Li, J.; Wang, H.; Xu, G.; Wang, X.; Tian, J. Synthesis of ZnO Nanosheets @In2O3 Hollow Micro-Rods Heterostructures for Enhanced Ethanol Gas Sensing Performance. Sens. Actuators B Chem. 2024, 404, 135271. [Google Scholar] [CrossRef]
  101. Kou, H.R.; Shao, T.T.; Dong, J.T.; Cheng, Y.Y.; Zhang, F.C.; Guo, J.M.; Liu, X.X.; Wang, X.Y. Ethanol Sensor Built on a SnO2/In2O3 Composite Generated from MOF. Sens. Actuators B-Chem. 2023, 396, 134628. [Google Scholar] [CrossRef]
  102. He, S.; Gui, Y.; Wang, Y.; Cao, L.; He, G.; Tang, C. CuO/TiO2 /MXene-Based Sensor and SMS-TENG Array Integrated Inspection Robots for Self-Powered Ethanol Detection and Alarm at Room Temperature. ACS Sens. 2024, 9, 1188–1198. [Google Scholar] [CrossRef] [PubMed]
  103. Mei, Z.H.; Song, Y.; Jiang, G.J.; Wu, D.D.; Li, G.R. SnO2/ZnO/Ti3C2Tx MXene Nanocomposites for Highly Sensitive and Stable Ethanol Sensing at Low Temperature. Sens. Actuators B-Chem. 2024, 419, 136390. [Google Scholar] [CrossRef]
  104. Li, T.; Yin, W.; Gao, S.W.; Sun, Y.N.; Xu, P.L.; Wu, S.H.; Kong, H.; Yang, G.Z.; Wei, G. The Combination of Two-Dimensional Nanomaterials with Metal Oxide Nanoparticles for Gas Sensors: A Review. Nanomaterials 2022, 12, 982. [Google Scholar] [CrossRef]
  105. Huang, S.; Shi, X.-R.; Sun, C.; Zhang, X.; Huang, M.; Liu, R.; Wang, H.; Xu, S. Template-Controlled in-Situ Growing of NiCo-MOF Nanosheets on Ni Foam with Mixed Linkers for High Performance Asymmetric Supercapacitors. Appl. Surf. Sci. 2022, 572, 151344. [Google Scholar] [CrossRef]
  106. Huang, S.; Shi, X.-R.; Sun, C.; Duan, Z.; Ma, P.; Xu, S. The Application of Metal–Organic Frameworks and Their Derivatives for Supercapacitors. Nanomaterials 2020, 10, 2268. [Google Scholar] [CrossRef]
  107. Wen, Y.; Xu, S.; Wang, P.; Shao, X.; Sun, X.; Hu, J.; Shi, X.-R. Bimetallic FeCo Phosphide Nanoparticles Anchored on N-Doped Carbon Foam for Wide PH Hydrogen Evolution Reaction. J. Alloys Compd. 2023, 931, 167570. [Google Scholar] [CrossRef]
  108. Liu, R.; Shi, X.-R.; Wen, Y.; Shao, X.; Su, C.; Hu, J.; Xu, S. Trimetallic Synergistic Optimization of 0D NiCoFe-P QDs Anchoring on 2D Porous Carbon for Efficient Electrocatalysis and High-Energy Supercapacitor. J. Energy Chem. 2022, 74, 149–158. [Google Scholar] [CrossRef]
  109. Shao, X.; Xu, S.; Wang, P.; Wen, Y.; Sun, X.; Hong, M.; Wu, K.; Shi, X.-R. Carbon-Incorporated Bimetallic Phosphide Nanospheres Derived from MOFs as Superior Electrocatalysts for Hydrogen Evolution. Dalton Trans. 2022, 51, 14517–14525. [Google Scholar] [CrossRef]
  110. Ma, H.; Xu, S.; Wang, P.; Zhu, J.; Yang, C.; Zhang, S.; Shi, X.-R.; Yao, L. Synergistic Improvement the Zn Storage Performance of ZnMn2O4 Quantum Dots by Ni Doping and In-Situ Electrochemical Induction. Appl. Surf. Sci. 2024, 663, 160208. [Google Scholar] [CrossRef]
  111. Gonçalves, B.F.; Fernández, E.; Valverde, A.; Gaboardi, M.; Salazar, H.; Petrenko, V.; Porro, J.M.; Cavalcanti, L.P.; Urtiaga, K.; Esperança, J.M.S.S.; et al. Exploring the Compositional Space of a Metal–Organic Framework with Ionic Liquids to Develop Porous Ionic Conductors for Enhanced Signal and Selectivity in VOC Capacitive Sensors. J. Mater. Chem. A Mater. 2024, 12, 14595–14607. [Google Scholar] [CrossRef]
  112. Mollick, S.; Rai, S.; Frentzel-Beyme, L.; Kachwal, V.; Donà, L.; Schürmann, D.; Civalleri, B.; Henke, S.; Tan, J. Unlocking Diabetic Acetone Vapor Detection by A Portable Metal-Organic Framework-Based Turn-On Optical Sensor Device. Adv. Sci. 2024, 11, 2305070. [Google Scholar] [CrossRef] [PubMed]
  113. Doan, T.L.H.; Kim, J.-Y.; Lee, J.-H.; Nguyen, L.H.T.; Dang, Y.T.; Bui, K.-B.T.; Pham, A.T.T.; Mirzaei, A.; Phan, T.B.; Kim, S.S. Preparation of N-ZnO/p-Co3O4 Heterojunctions from Zeolitic Imidazolate Frameworks (ZIF-8/ZIF-67) for Sensing Low Ethanol Concentrations. Sens. Actuators B Chem. 2021, 348, 130684. [Google Scholar] [CrossRef]
  114. Li, G.; Zhang, Y.Y.; Liang, Q.F.; Zhang, J.N.; Liu, J.; Liu, Y.M.; Wang, C.L.; Gao, J.Z.; Lu, H.B. Nanoporous Co3O4-TiO2 Heterojunction Nanosheets for Ethanol Sensing. ACS Appl. Nano Mater. 2022, 5, 4779–4786. [Google Scholar] [CrossRef]
  115. Zhang, H.; Guo, S.; Zheng, W.; Wang, H.; Li, H.-Y.; Yu, M.-H.; Chang, Z.; Bu, X.-H.; Liu, H. Facile Engineering of Metal–Organic Framework Derived SnO2-ZnO Composite Based Gas Sensor toward Superior Acetone Sensing Performance. Chem. Eng. J. 2023, 469, 143927. [Google Scholar] [CrossRef]
  116. Zhang, K.; Qin, S.W.; Tang, P.G.; Feng, Y.J.; Li, D.Q. Ultra-Sensitive Ethanol Gas Sensors Based on Nanosheet-Assembled Hierarchical ZnO-In2O3 Heterostructures. J. Hazard. Mater. 2020, 391, 122191. [Google Scholar] [CrossRef]
  117. Yan, W.J.; Chen, Y.L.; Zeng, X.M.; Wu, G.; Jiang, W.; Wei, D.; Ling, M.; Ng, K.W.; Qin, Y.X. Ultrasensitive Ethanol Sensor Based on Segregated ZnO-In2O3 Porous Nanosheets. Appl. Surf. Sci. 2021, 535, 147697. [Google Scholar] [CrossRef]
  118. Liu, Z.Y.; Zhang, Z.K.; Yue, C.; Mu, Y.; Yang, Z.G.; Dastan, D.; Yin, X.T.; Ma, X.G. In-MIL-68 Derived In2O3/Fe2O3 Shuttle-like Structures with n-n Heterojunctions to Improve Ethanol Sensing Performance. Phys. Chem. Chem. Phys. 2024, 26, 4184–4193. [Google Scholar] [CrossRef]
  119. Ma, S.; Shen, L.Y.; Ma, S.H.; Wen, J.; Xu, J.Y. Emerging Zinc Stannate and Its Application in Volatile Organic Compounds Sensing. Coord. Chem. Rev. 2023, 490, 215217. [Google Scholar] [CrossRef]
  120. Yang, X.; Li, H.; Li, T.; Li, Z.; Wu, W.; Zhou, C.; Sun, P.; Liu, F.; Yan, X.; Gao, Y.; et al. Highly Efficient Ethanol Gas Sensor Based on Hierarchical SnO2/Zn2SnO4 Porous Spheres. Sens. Actuators B Chem. 2019, 282, 339–346. [Google Scholar] [CrossRef]
  121. Ma, D.P.; Zhang, L.; Hu, J.T.; Fu, Z.W.; Luo, T.; Yang, D.; Fang, D.; Li, J.; Peng, J.B.; Wang, Y.W. Preparation and Gas-Sensitive Properties of Hollow Zn2SnO4/SnO2 Nano-Cubes. Inorg. Chem. Commun. 2022, 141, 109507. [Google Scholar] [CrossRef]
  122. Guo, R.; Wang, H.; Tian, R.; Shi, D.; Li, H.; Li, Y.; Liu, H. The Enhanced Ethanol Sensing Properties of CNT@ZnSnO3 Hollow Boxes Derived from Zn-MOF(ZIF-8). Ceram. Int. 2020, 46, 7065–7073. [Google Scholar] [CrossRef]
  123. Cao, H.; Hu, Z.; Wei, X.; Wang, H.; Tian, X.; Ding, S. Conductometric Ethanol Gas Sensor Based on a Bilayer Film Consisting of SnO2 Film and SnO2/ZnSnO3 Porous Film Prepared by Magnetron Sputtering. Sens. Actuators B Chem. 2023, 382, 133562. [Google Scholar] [CrossRef]
  124. Lu, Y.Y.; Yang, X.W.; Bao, H.M.; Lei, B.; Chen, K.; Wei, Y.; Zhao, Q.; Zhang, H.W.; Cai, W.P. Vortex Engineering on Oxide Bowl-Coated Oxide/Gold Dual-Layer Array for Dual Electrical/Spectroscopic Monitoring of Volatile Organic Compounds. Adv. Funct. Mater. 2024, 34, 2402173. [Google Scholar] [CrossRef]
  125. Li, D.; Liu, G.; Zhang, Q.; Qu, M.; Fu, Y.Q.; Liu, Q.; Xie, J. Virtual Sensor Array Based on MXene for Selective Detections of VOCs. Sens. Actuators B Chem. 2021, 331, 129414. [Google Scholar] [CrossRef]
  126. Sun, C.; Zhang, S.; Wang, P.; Wei, M.; Wang, S.; Shi, X.-R. High-Throughput Screening of Efficient Graphdiyne Supported Transition Metal Single Atom toward Water Electrolysis and Oxygen Reduction. J. Catal. 2024, 439, 115773. [Google Scholar] [CrossRef]
  127. Ferko, N.; Djeziri, M.A.; Al Sheikh, H.; Moubayed, N.; Bendahan, M.; El Rafei, M.; Seguin, J.L. Methodology for Estimating Ethanol Concentration with Artificial Intelligence in the Presence of Interfering Gases and Measurement Delay. Sens. Actuators B-Chem. 2024, 421, 136502. [Google Scholar] [CrossRef]
  128. Li, J.W.; Weng, H.D.; Yang, Q.Q.; Shen, J. Data-Driven Diagnosis Method of High-Pressure Hydrogen Leakage Based on Actual Driving Conditions and Probabilistic Neutral Network. Int. J. Hydrog. Energy 2024, 71, 411–421. [Google Scholar] [CrossRef]
  129. Shao, S.F.; Yan, L.W.; Zhang, L.; Zhang, J.; Li, Z.X.; Kim, H.W.; Kim, S.S. Utilizing Data Mining for the Synthesis of Functionalized Tungsten Oxide with Enhanced Oxygen Vacancies for Highly Sensitive Detection of Triethylamine. ACS Appl. Mater. Interfaces 2024, 16, 6098–6112. [Google Scholar] [CrossRef]
  130. Nakano-Baker, O.; Fong, H.; Shukla, S.; Lee, R.V.; Cai, L.; Godin, D.; Hennig, T.; Rath, S.; Novosselov, I.; Dogan, S.; et al. Data-Driven Design of a Multiplexed, Peptide-Sensitized Transistor to Detect Breath VOC Markers of COVID-19. Biosens. Bioelectron. 2023, 229, 115237. [Google Scholar] [CrossRef]
Figure 1. Illustration of three types of heterojunction.
Figure 1. Illustration of three types of heterojunction.
Chemosensors 12 00244 g001
Figure 2. Illustration of strategies for promoting the sensing performance of MOS heterojunctions.
Figure 2. Illustration of strategies for promoting the sensing performance of MOS heterojunctions.
Chemosensors 12 00244 g002
Figure 3. Illustrations of preparation methods for MOSs heterojunction materials.
Figure 3. Illustrations of preparation methods for MOSs heterojunction materials.
Chemosensors 12 00244 g003
Figure 4. (a) The TEM image of hierarchical Fe2O3-Co3O4 heterojunction; (b) the contents of oxygen species based on the analysis of X-ray photoelectron spectroscopy (XPS); (c) the dynamic dot-line pattern of three sensors; (d) the selectivity of all gas sensors’ exposure to 100 ppm of different gases [86]. Reprinted with permission from Elsevier, copyright 2024.
Figure 4. (a) The TEM image of hierarchical Fe2O3-Co3O4 heterojunction; (b) the contents of oxygen species based on the analysis of X-ray photoelectron spectroscopy (XPS); (c) the dynamic dot-line pattern of three sensors; (d) the selectivity of all gas sensors’ exposure to 100 ppm of different gases [86]. Reprinted with permission from Elsevier, copyright 2024.
Chemosensors 12 00244 g004
Figure 5. Schematic illustration of the synthetic process of the ZIF-67-derived oxide cage/nanofiber Co3O4/In2O3 heterostructure for acetone gas sensing [55]. Reprinted with permission from [56]. Copyright {2024} American Chemical Society.
Figure 5. Schematic illustration of the synthetic process of the ZIF-67-derived oxide cage/nanofiber Co3O4/In2O3 heterostructure for acetone gas sensing [55]. Reprinted with permission from [56]. Copyright {2024} American Chemical Society.
Chemosensors 12 00244 g005
Figure 6. (a) Illustration of the preparation process for the Ag-NiO/SnO2 nanotubes (NTs); (b) scanning electron microscope (SEM) image; (c) transmission electron microscope (TEM) image; and (d) the diagram of the energy band structure of the Ag-NiO/SnO2 NTs [87]. Reprinted with permission from Elsevier, copyright 2024. (e) Synthetic scheme of ZnO-CuO core–hollow cube nanostructures [90]. Reprinted with permission from [91]. Copyright {2020} American Chemical Society.
Figure 6. (a) Illustration of the preparation process for the Ag-NiO/SnO2 nanotubes (NTs); (b) scanning electron microscope (SEM) image; (c) transmission electron microscope (TEM) image; and (d) the diagram of the energy band structure of the Ag-NiO/SnO2 NTs [87]. Reprinted with permission from Elsevier, copyright 2024. (e) Synthetic scheme of ZnO-CuO core–hollow cube nanostructures [90]. Reprinted with permission from [91]. Copyright {2020} American Chemical Society.
Chemosensors 12 00244 g006
Figure 7. (a) Preparation of the p-type CuO/TiO2/MXene gas-sensitive sensor; (b) energy band diagram of the gas-sensitive sensor in four different situations [102]. Reprinted with permission from [104]. Copyright {2024} American Chemical Society.
Figure 7. (a) Preparation of the p-type CuO/TiO2/MXene gas-sensitive sensor; (b) energy band diagram of the gas-sensitive sensor in four different situations [102]. Reprinted with permission from [104]. Copyright {2024} American Chemical Society.
Chemosensors 12 00244 g007
Figure 8. (a) Schematic illustration of the formation process of the 3% Fe2O3-loaded ultrathin nanosheet assembled hollowed-out hierarchical NiO nanorods (Fe2O3@NiO); (b) SEM and TEM images of Fe2O3@NiO; (c) response of the Fe2O3@NiO to 10 ppm ethanol at different operating temperatures and 80% RH; (d) response of the Fe2O3@NiO at 150 °C to 10 ppm of different target gases under different humidity conditions [98]. Reprinted with permission from [99]. Copyright {2023} American Chemical Society.
Figure 8. (a) Schematic illustration of the formation process of the 3% Fe2O3-loaded ultrathin nanosheet assembled hollowed-out hierarchical NiO nanorods (Fe2O3@NiO); (b) SEM and TEM images of Fe2O3@NiO; (c) response of the Fe2O3@NiO to 10 ppm ethanol at different operating temperatures and 80% RH; (d) response of the Fe2O3@NiO at 150 °C to 10 ppm of different target gases under different humidity conditions [98]. Reprinted with permission from [99]. Copyright {2023} American Chemical Society.
Chemosensors 12 00244 g008
Figure 9. (a) The schematic illustration of the construction and engineering of SnO2-ZnO based on MOF precursor and corresponding TEM and HRTEM images [115]. Reprinted with permission from Elsevier, copyright 2023.; (b) SEM image of ZnO-SnO2 heterojunction IOPBs; and (c) the response at 260 °C to 10, 20, 50, 80, and 100 ppm acetone [69]. Reprinted with permission from Elsevier, copyright 2024.
Figure 9. (a) The schematic illustration of the construction and engineering of SnO2-ZnO based on MOF precursor and corresponding TEM and HRTEM images [115]. Reprinted with permission from Elsevier, copyright 2023.; (b) SEM image of ZnO-SnO2 heterojunction IOPBs; and (c) the response at 260 °C to 10, 20, 50, 80, and 100 ppm acetone [69]. Reprinted with permission from Elsevier, copyright 2024.
Chemosensors 12 00244 g009
Figure 10. Schematic diagram of gas molecular adsorption, electron transfer, and energy band structure for the In2O3-ZnO material in (a) air and (b) ethanol [50]. Reprinted with permission from Elsevier, copyright 2023.
Figure 10. Schematic diagram of gas molecular adsorption, electron transfer, and energy band structure for the In2O3-ZnO material in (a) air and (b) ethanol [50]. Reprinted with permission from Elsevier, copyright 2023.
Chemosensors 12 00244 g010
Figure 11. (a) SEM image of Zn2SnO4/CdSnO3; (b) EPR spectra of CdSnO3, Zn2SnO4, and Zn2SnO4/CdSnO3; (c) schematic diagram of the energy band structure and ethanol gas sensing process of CdSnO3/Zn2SnO4 heterostructure [57]. Reprinted with permission from Elsevier, copyright 2024.
Figure 11. (a) SEM image of Zn2SnO4/CdSnO3; (b) EPR spectra of CdSnO3, Zn2SnO4, and Zn2SnO4/CdSnO3; (c) schematic diagram of the energy band structure and ethanol gas sensing process of CdSnO3/Zn2SnO4 heterostructure [57]. Reprinted with permission from Elsevier, copyright 2024.
Chemosensors 12 00244 g011
Figure 12. (a) Schematic illustration of the fabrication process of In2O3@PW12@SnO2 NFs [97]; (b) TEM image of In@3%P@Sn NFs after calcination and selectivity pattern of the gas sensors based on In@3%P@Sn [97]; (c) transient responses of different concentrations of ethanol at the optimal operating temperature [97]. Reprinted with permission from Elsevier, copyright 2024.
Figure 12. (a) Schematic illustration of the fabrication process of In2O3@PW12@SnO2 NFs [97]; (b) TEM image of In@3%P@Sn NFs after calcination and selectivity pattern of the gas sensors based on In@3%P@Sn [97]; (c) transient responses of different concentrations of ethanol at the optimal operating temperature [97]. Reprinted with permission from Elsevier, copyright 2024.
Chemosensors 12 00244 g012
Figure 13. (a) Formation process schematic for SnO2/ZnO/Ti3C2Tx MXene nanocomposites; (b) the curves for dynamic response and recovery of SnO2/ZnO/Ti3C2Tx MXene at 120 °C to 100 ppm ethanol [103]; (c) SnO2/ZnO/Ti3C2Tx MXene responses to various gases (100 ppm) at 120 °C [103]; (d) electron transfer of ZnO, SnO2, and Ti3C2Tx MXene in air and ethanol [103]. Reprinted with permission from Elsevier, copyright 2024.
Figure 13. (a) Formation process schematic for SnO2/ZnO/Ti3C2Tx MXene nanocomposites; (b) the curves for dynamic response and recovery of SnO2/ZnO/Ti3C2Tx MXene at 120 °C to 100 ppm ethanol [103]; (c) SnO2/ZnO/Ti3C2Tx MXene responses to various gases (100 ppm) at 120 °C [103]; (d) electron transfer of ZnO, SnO2, and Ti3C2Tx MXene in air and ethanol [103]. Reprinted with permission from Elsevier, copyright 2024.
Chemosensors 12 00244 g013
Table 1. Comparison of the responses of sensors for acetone detection. N.A. means that the data are not available. The same applies below.
Table 1. Comparison of the responses of sensors for acetone detection. N.A. means that the data are not available. The same applies below.
MaterialsTypeConcentration
(ppm)
Operating Temperature
(°C)
Response (Ra/Rg)Detection Limit
(ppm)
Response/Recovery Time
(s)
Reference
Fe2O3-Co3O4p-n10020091.5N.A.20/21[86]
Co3O4/In2O3p-n I503009540.01884/148[55]
Ag-NiO/SnO2p-n II119015.70.0512/25[87]
SnO2-ZnOn-n II1002500.2353/333[88]
SnO2/ZnSnO3n-n I1002900.045301/6[89]
ZnO-CuOp-n II120011.140.009N.A./N.A.[90]
TiO2/a-Fe2O3n-n II10022521.90.03613/10[91]
ZnFe2O4/SnO2n-n II1002101200.130/197.2[92]
ZnSnO3/ZnO/Ti3C2Tx MXeneN.A.10012015.68905 /12 [54]
Table 2. Comparison of the responses of sensors for ethanol detection.
Table 2. Comparison of the responses of sensors for ethanol detection.
MaterialsTypeConcentration
(ppm)
Operating Temperature
(°C)
Response (Ra/Rg)Detection Limit
(ppm)
Response/Recover Time
(s)
Reference
Mesoporous
In2O3-ZnO
n-n II100225350.24/90[50]
Core-double
shell
ZnO@In2O3
@ZnO
n-n100200453120/190[99]
MoO3 nanorods @SnO2 nanosheetsN.A.10020048.64N.A.65/230[56]
In2O3@PW12@SnO2n-n10032022.60.01391/132[97]
Porous Zn2SnO4/CdSnO3 nanocubesn-n II100300214.380.28530/55[57]
Hollowed-out Fe2O3-loaded NiO heterojunction nanorodsp-n I1015051.20.5N.A./N.A.[98]
ZnO nanosheets @In2O3 hollow microrodsn-n1002001269.118/35[100]
In2O3/SnO2n-n II100200N.A.49.1632/40[101]
CuO/TiO2/MXeneN.A.1300.950.316/13[102]
SnO2/ZnO/Ti3C2Tx MXeneN.A.100120121.1N.A.3/141[103]
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

Zhang, S.; Zhang, H.; Yao, H.; Wang, P.; Zhu, M.; Shi, X.; Xu, S. Recent Advances in Metal Oxide Semiconductor Heterojunctions for the Detection of Volatile Organic Compounds. Chemosensors 2024, 12, 244. https://doi.org/10.3390/chemosensors12120244

AMA Style

Zhang S, Zhang H, Yao H, Wang P, Zhu M, Shi X, Xu S. Recent Advances in Metal Oxide Semiconductor Heterojunctions for the Detection of Volatile Organic Compounds. Chemosensors. 2024; 12(12):244. https://doi.org/10.3390/chemosensors12120244

Chicago/Turabian Style

Zhang, Shengming, Heng Zhang, Haiyu Yao, Peijie Wang, Min Zhu, Xuerong Shi, and Shusheng Xu. 2024. "Recent Advances in Metal Oxide Semiconductor Heterojunctions for the Detection of Volatile Organic Compounds" Chemosensors 12, no. 12: 244. https://doi.org/10.3390/chemosensors12120244

APA Style

Zhang, S., Zhang, H., Yao, H., Wang, P., Zhu, M., Shi, X., & Xu, S. (2024). Recent Advances in Metal Oxide Semiconductor Heterojunctions for the Detection of Volatile Organic Compounds. Chemosensors, 12(12), 244. https://doi.org/10.3390/chemosensors12120244

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