Next Article in Journal
YOLO-BOS: An Emerging Approach for Vehicle Detection with a Novel BRSA Mechanism
Next Article in Special Issue
ZnO/MOx Nanofiber Heterostructures: MOx Receptor’s Role in Gas Detection
Previous Article in Journal
Bayesian Architecture for Predictive Monitoring of Unbalance Faults in a Turbine Rotor–Bearing System
Previous Article in Special Issue
ZnO Nanowires/Self-Assembled Monolayer Mediated Selective Detection of Hydrogen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of MEMS Gas Sensors: A Comprehensive Review of Sensing Materials

1
Ministry-of-Education Key Laboratory for the Green Preparation and Application of Functional Materials, School of Materials Science & Engineering, Hubei University, Wuhan 430062, China
2
Wuhan Micro & Nano Sensor Technology Co., Ltd., Xingye Building, Wuda Science and Technology Park, Donghu New Technology Development Zone, Wuhan 430223, China
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(24), 8125; https://doi.org/10.3390/s24248125
Submission received: 15 October 2024 / Revised: 2 December 2024 / Accepted: 4 December 2024 / Published: 19 December 2024
(This article belongs to the Special Issue Advanced Nanomaterials for Sensing)

Abstract

:
The MEMS gas sensor is one of the most promising gas sensors nowadays due to its advantage of small size, low power consumption, and easy integration. It has been widely applied in energy components, portable devices, smart living, etc. The performance of the gas sensor is largely determined by the sensing materials, as well as the fabrication methods. In this review, recent research progress on H2, CO, NO2, H2S, and NH3 MEMS sensors is surveyed, and sensing materials such as metal oxide semiconductors, organic materials, and carbon materials, modification methods like construction of heterostructures, doping, and surface modification of noble metals, and fabrication methods including chemical vapor deposition (CVD), sputtering deposition (SD), etc., are summarized. The effect of materials and technology on the performance of the MEMS gas sensors are compared.

1. Introduction

With the development of industry and technology, the potential danger of toxic or flammable and explosive gases leaking in the production and transportation process needs to be monitored and brought to attention as early as possible. The advance of high-performance, low-cost gas sensors is one of the most effective solutions. Great efforts have been put into the investigation of novel materials and facile processes with the aim of improving response sensitivity, reducing reaction temperatures, and minimizing device sizes. MEMS sensors are sensors based on microelectromechanical systems, usually with small sizes (1 μm and 1 mm) and unique manufacturing methods. They have the advantages of miniaturization, low power consumption, easily configurable multifunctionality, and easy integration. Various MEMS sensors, including pressure, acceleration, gyro, flow, gas, and temperature sensors, have been developed. The MEMS gas sensor is one of the most frequently used gas sensors at present. Compared with other types of gas sensors, MEMS gas sensors can be combined with temperature and humidity sensors more easily, and gas-sensing arrays or grids can also be easily constructed. Additionally, they have high sensitivity, fast response, lower power consumption, and a smaller size, enabling increased practicality in a wider range of applications. The detection of different gases can be realized through the development and improvement of sensitive materials with different properties [1].
The performance of the sensor can be evaluated according to the sensitivity, resolution, detection limit (LOD), response time, recovery time, and long-term stability. In this paper, recent progress on MEMS gas sensors including hydrogen, carbon monoxide, nitrogen dioxide, hydrogen sulfide, and ammonia sensors are reviewed in detail, and the above performance parameters of the gas sensor with different materials are compared. Firstly, we discuss the properties of different materials used in MEMS gas sensors. Then, the properties, application areas, and research progress regarding the five commonly used gas sensors are described in detail. Finally, we provide a summary of the performance of the various MEMS gas sensors, the remaining problems, and possible solutions to solve these problems in the future.

2. Materials for MEMS Gas Sensors

In recent years, research on different types of sensing materials for MEMS gas sensors has continued to expand, containing metal materials, metal oxide semiconductors (MOSs), metal–organic frameworks (MOFs), graphene and its derivatives, carbon nanotubes and their derivatives, nano-metal particles, transition metal dihalides (TMDs), etc. (Figure 1) [2,3]. Among them, MOS material is the most widely used MEMS gas-sensitive material because of its excellent performance and outstanding stability.

2.1. Metallic Material

Metal materials, including Pt, Pd, Au, Cr, and Al, are used in MEMS gas sensors in the form of metal nanoparticles, metal core–shell structures, and metal films. They meet the needs of MEMS sensing materials for suitable surface area, high diffusion rate, fast adsorption/desorption kinetics, and monitoring significant changes in material properties in the presence of gas molecules. In addition, alloys, metallic glass (MG), and other materials can be created through the combination of a variety of metal materials to achieve further improvement in performance.

2.2. Metal Oxide Semiconductor Materials

MOSs are the most widely used MEMS gas-sensing materials because of their excellent performance and good chemical and physical stability. The relationship between the resistance value of an MOS and the gas concentration on its surface is used to quantize the gas response. MOS-based materials can be divided into two types: N-type materials like TiO2, ZnO, SnO2, WO3, CeO2, In2O3, etc., and P-type materials such as Co2O3, CuO, and so on [4]. The difference between the two is that the charge carriers in N-type materials are electrons, while the charge carriers in P-type materials are holes [5]. The concentration of holes in P-type materials and electrons in N-type materials varies in different gas environments [5]. The relationship between the resistance value and gas concentration in an MOS gas sensor is related to its sensitivity characteristics. When an N-type MOS gas-sensitive material is exposed to air and heated to several hundred degrees Celsius by the heating wire, the oxygen in the air reacts at high temperature, trapping the electrons in the gas-sensitive material to generate active adsorbed oxygen negative ions (O2−, O, and O2), as shown in reaction Formulas (1)–(3):
O2(gas) ↔ O2(ads)
O2(ads) + e = O2(ads)
O2 + e = 2O(ads)
The oxygen species adsorb on the gas-sensitive material, attract free electrons, and reduce the flowing electrons in the semiconductor material, thereby increasing the sensor resistance. When the reducing gases, such as H2 and CO, appear, the reducing gases interact with the adsorbed negative oxygen ions, and the originally adsorbed electrons are released into the gas-sensitive material, which increases the sensor current and decreases the resistance [5]. With the increase in the reducing gas concentration, the resistance value of the sensor decreases gradually. Similarly, if the test gas is oxidizing, the resistance value will increase with the increase in concentration [5].
When a P-type semiconductor is exposed to air, oxygen molecules capture the electrons inside the gas-sensitive material and produce oxygen negative ions (O), forming a hole accumulation layer (HAL) on its surface, resulting in a reduction in the resistance of the P-type semiconductor material [5]. When reducing gas appears, oxygen anions react with it, the surface cavity accumulation layer becomes thinner, and the resistance value increases. On the contrary, when oxidizing gas is present, the resistance value decreases as its concentration increases.
Doping, noble metal modification, and heterostructure construction have been proposed to improve the gas-sensitive properties of MOS-based materials [4]. Doping refers to the introduction of heteroatoms into the MOS lattice, and it is a recognized as an effective way to change the sensing properties of materials, in which one or more metal atoms can be introduced. Structure defects and surface oxygen vacancies can be created through doping, thus promoting surface reactions and improving the gas-sensitive properties of materials [4]. Noble metal particles have electron sensitization and chemical catalysis; modification of MOS materials with noble metal particles can also improve their sensing performance [4]. The construction of heterogeneous structures refers to the realization of p-n, p-p, and n-n heterogeneous structures by the combination of two or more metal oxide semiconductor materials [5]. The heterojunction formed between the two MOS materials can often increase the mobility of charge carriers, thus increasing the gas-sensing response [4]. The heterogeneous materials are usually prepared by the hydrothermal method, electrospinning method, solvothermal method, chemical vapor deposition method, physical vapor deposition method, radio frequency sputtering method, etc. Sometimes it may be necessary to choose two or more ways to prepare materials to achieve changes in composition or structure to obtain better properties [6].

2.3. Carbon-Based Material

The commonly used carbon-based materials (CBMs) for MEMS gas sensors are graphene and carbon nanotubes (CNTs). Both single-walled carbon nanotubes (SWCNTs) and multi-walled carbon nanotubes (MWCNTs) have the potential to create high-performance gas sensors for the detection of a range of toxic gases, especially in the field of intelligent sensing and monitoring of large power equipment faults such as oil-immersed transformers [7]. However, these sensors still have the problem of the amount of detected gas being limited and the sensitivity needing to be further improved. Materials related to graphene include graphite sheets (Gt), graphene oxide (GO), reduced graphene oxide (RGO), etc. A GO sheet is the product of chemical oxidation and stripping of graphite powder. RGO is a two-dimensional layered carbonaceous material with a large number of active functional groups on its surface, which makes the sensing material more sensitive to gas molecules.

2.4. Silicon-Based Material

Silicon-based materials (SBMs), including silicon nanomembranes (SiNMs), silicon nanowires (SiNWs), and SiC, are also used in MEMS gas sensor preparation. Among them, SiNMs have a variety of electrical and material characteristics and can be used to make flexible RF thin-film transistors, microwave switches, and solar cells [8]. The silicon-based material gas sensor has the characteristics of strong stability, high reliability, and non-ageing and can be applied to the production of gas sensors requiring mobile operation in the future.

2.5. Polymer Material

Polymer gas-sensitive materials such as polypyrrole (PPy), polystyrene (PS), polythiophene (PTh), and other polymer materials are often used to develop MEMS gas sensors. Gas sensors made of polymer materials often have high flexibility and sensitivity. Functionalization of organic active materials with special functional groups and the preparation of composite films that provide more active sites and interfaces are effective means to improve the selectivity and sensitivity of polymer gas sensors.

2.6. Metal–Organic Framework Material

MOFs are a new generation of porous materials with large porosity and pore size variations, very high surface area (up to 10,000 m2/g), low density, and chemical tunability, and they have been widely studied and applied in many fields, including gas storage and separation, catalysis, biomedicine (such as drug storage and delivery), etc., [6]. In recent years, MOF materials have also often been used in the preparation of MEMS gas sensors.

3. Hydrogen Sensors

Hydrogen is a highly anticipated new energy gas. Due to its low molecular weight, it can easily leak during production and transportation, and it can easily explode in case of open fire. Therefore, it is necessary to develop MEMS sensors with high H2 sensitivity and good selectivity. In the field of new energy, these sensors can be used to ensure the safe operation of lithium batteries and fuel cells, etc.; in the industrial field, they can be used to detect hydrogen leaks in the process of steel manufacturing, and in the field of environmental protection, they can be used to ensure the safety and reliability of hydrogen energy use. In recent years, researchers have continued to develop hydrogen sensors with better performance and lower operating temperatures. Table 1 lists the performance of some hydrogen sensors developed in recent years.
Table 1. Performance comparison of MEMS hydrogen gas sensors and sensing materials.
Table 1. Performance comparison of MEMS hydrogen gas sensors and sensing materials.
Sensitive Material CategorySensitive MaterialFabrication TechniqueWorking Temperature (°C)Detection Range (ppm)Response Time (s)Recovery Time (s)SensitivityRef.
MetalPtSputter technique30–2001000–10,000N/aN/a3.4@1000 ppm d[9]
MetalPdPolyimide-based microfabrication process25–855000–40,000<5N/aN/a[10]
AlloyPd/AuDirect current (DC) magnetron SD605–30,000221603.3%@30,000 ppm c[11]
AlloyPd/MgDC magnetron sputtering100N/a160N/a[12]
AlloyPd/NiE-beam evaporation2580–500084967.1%@5000 ppm c[13]
MGPd78Cu5Si17Multi-target sputteringRoom
temperature
0.05–40,0002.1N/aN/a[14,15]
CBMPd/CNTElectrochemical deposition processRoom
temperature
303N/a75%@500 ppm c[16]
Metal/CBMPd cluster/graphene electrodesElectrodeposition methodRoom
temperature
N/aN/aN/aN/a[17]
SBMPd/SiNME-beam evaporationRoom
temperature
50–500022N/a764%@500 ppm c[7]
SBMPd/Si nanomeshNanosphere lithographyRoom
temperature
1000–10,00051327%@10,000 ppm c[18]
SBM3C–SiCCVDRoom
temperature
2000–20,000N/aN/aN/a[19]
SBMPd/SiNWE-beam evaporationRoom
temperature
100–1000N/aN/a154.5%@1000 ppm c[20]
MOSSnO2 with defectsDrop sintered2500.1–67122.3@6 ppm a[21]
MOSWO3Radio frequency (RF) magnetron method20–3505–1000120–180120–1801300@1000 ppm a[22]
MOSAmmonia plasma modification ZnO-NWsSputter techniqueRoom
temperature
500–2500N/aN/a27%@2500 ppm c[23]
Metal/MOSPd/ZnO-NRsSol–gel technologyRoom
temperature
0.2–100018.8N/a91%@1000 ppm c[24]
Metal/MOSPd nanotube/ZnOLow-temperature, wet-chemical process/hydrothermal methodRoom
temperature
N/aN/aN/a1500%@1000 ppm c[25]
Metal/MOSPt/TiO2Calcining method500–800N/a0.04 (H2)
0.02 (O2)
N/aN/a[26]
Metal/MOSPt/Nb/TiO2Electrostatic microspray method40200–10003127012.3@1000 ppm a[27]
Metal/MOSPd/SnO2 thin filmsSol–gel method75150–1000182108.9165@1000 ppm a[28]
MOS/MOSMOF-derived WO3-C/In2O3Drop coating2505–20001.99.210.11@1000 ppm a[29]
MOS/metal/MOFZnO/Pd/ZIF-8
nanowires
N/a20010–50N/a8.336.2@50 ppm b[30]
a R = Ra/Rg. b R = Rg/Ra. c R = |Rg − Ra|/Ra × 100%. d R = |Rg − Ra|/Rg × 100%.

3.1. Metallic Material

3.1.1. Metal

Sennik et al. used RF sputtering to develop an MEMS H2 sensor using a Pt film deposited on a glass substrate as the sensing material [9]. The induction mechanism is shown in Figure 2a: After exposure to hydrogen, the surface oxygen atoms of Pt film will be replaced by hydrogen atoms, without forming a bulk hydride phase [9]. And the replacement process is reversible [9]. Under dry airflow conditions, the optimal operating temperature of a 2 nm thick Pt film H2 sensor is 30 °C [9]. The H2 detection concentration range of the sensor is 0.1–1%, and the response value Ra/Rg = 3.4 when the H2 concentration is 1000 ppm [9]. Walewyns et al. introduce a Pd/Al-based 3D-MEMS capacitive H2 sensor. The Pd hydrogenation reaction gives it ultra-low power, a fast response (5 s), high sensitivity to the lower explosive limit (LEL), and high selectivity for hydrogen, which has higher safety in future low-power sensing applications with high selectivity and dynamics [10].

3.1.2. Alloy

Gong et al. reported an MEMS resistive H2 sensor based on a Pd-Au alloy thin film produced by DC magnetron SD [11]. Figure 2b illustrates the lattice expansion and diffusion properties of Pd-based alloy films [11]. After the hydrogen atom reacts with the Pd atom to form PdHx, lattice expansion occurs, and the expanded lattice increases the movement space and diffusion rate of the hydrogen atom in the alloy film, thus improving the response speed of the sensor [11]. The state of hydrogen atoms adsorbed and diffused into Pd-based films is affected by the partial pressure of hydrogen in the environment [11]. When the hydrogen partial pressure is stable, the physical and chemical properties of the film also remain stable [11]. The sensor operates at 60 °C, with a detection range of 5 ppm to 3% H2 and a response value of 3.3% when the hydrogen concentration is 3% [11]. The response time and recovery time are 22 s and 160 s, respectively [11]. The sensor has low power consumption, high production, high performance, good repeatability, long-term stability, and high selectivity for H2 [11]. Sanger et al. developed a Pd-coated Mg film deposited on an electrochemically etched porous silicon substrate using DC magnetron sputtering technology and used the material for the development of MEMS H2 sensing [12]. Figure 2c illustrates the reaction process of Pd/Mg films during hydrogenation and dehydrogenation [12]. The formation of MgH2 in the Mg layer leads to a change in the resistance of the Pd/Mg layer [12]. The porous structure of the silicon substrate allows hydrophobic and high-surface-area film deposition and prevents the material from delaminating from the surface under humid conditions [12]. The resistance of Pd/Mg films is reversible and mechanically stable during multiple hydrogenation/dehydrogenation cycles [12]. The Pd/Mg thin-film H2 sensor has the advantages of fast response (1 s), economy, and sensitivity at low temperatures [12]. Kondalkar et al. reported a Pd-Ni (3%) based bendable resistance MEMS H2 sensor for dissolved hydrogen analysis (DHGA) [13]. The manufacturing steps are shown in Figure 2d, including the deposition of Si3N4 and SiO2 by PECVD, the deposition of Ti/Pt/Ni by electron beam evaporator, the deposition of an aluminum wire, and the formation of Al2O3 by the atomic layer deposition process [13]. Atomic layer deposition (ALD) technology is also utilized to increase the surface area of the material to improve gas-sensitive properties [13]. It has a good response to 80–5000 ppm H2 concentration and has the highest response value at 25 °C with a response value (∆R/R0) of 7.1% at 5000 ppm [13]. The research results show that the developed sensor has stable performance and is very promising in the application of transformer monitoring [13].

3.1.3. Metallic Glass

Yamazaki et al. fabricated a capacitive MEMS H2 sensor based on Pd-based metallic glass (MG) by multi-target sputtering [14,15]. As shown in Figure 2e, a system with two DC sources and one RF source is employed for sputtering Pd, Si, and Cu [14]. The sputtering system is capable of controlling the compositional ratio through the modification of the sputtering intensity for both DC and RF sources, thereby elucidating the ease and precision inherent in RF sputtering processes [14]. The research shows that PdCuSi, as a PD-based MG, has a good application prospect in capacitive MEMS H2 sensors [14]. In addition, several PdCuSi compositions with different component contents were compared. Only Pd78Cu5Si17 was an amorphous metal with MG characteristics [14]. Under the condition of a low hydrogen concentration of 0.05 vol% to 4.0 vol%, its strain well followed Sieferts’s law, indicating that hydrogen exists in MG in a diffused state [15]. The sensor is manufactured by surface multi-target sputtering technology and has the characteristics of non-hysteresis and fast response (2.1 s) at room temperature, so it has obvious advantages in hydrogen leak detection applications [14].

3.2. MOS Material

3.2.1. MOS

A flexible H2 sensor also has good mechanical bending test results and can be applied to vehicles, aircraft, aviation, and portable electronic devices in the future [24]. Luo et al. synthesized SnO2 materials with oxygen vacancy defects used for MEMS H2 sensors [21]. As shown in Figure 2f, the flexible hydrothermal method was used to prepare ZnO nano-Petri dishes with oxygen-rich vacancies. Samples (SnO2-D3, SnO2-D4, SnO2-D5, SnO2-D6, and SnO2-D7) were obtained by annealing SnO2 at different temperatures (300 °C, 400 °C, 500 °C, 600 °C, and 700 °C), and the sample drops were coated on Pt digital electrodes, dried, and then sintered to complete the preparation [21]. When the SnO2-D sample is exposed in H2, lattice oxygen reacts with H2, forming an oxygen vacancy defect on the surface. The gas molecules react with chemisorbed oxygen, resulting in the release of more electrons after the oxidation–reduction reaction, making SnO2-D more sensitive and faster than the original SnO2 for gas sensing [21]. The original SnO2 has a response time/recovery time of 12 s/15 s for 6 ppm H2 at 250 °C, while the SnO2-D3, SnO2-D4, and SnO2-D5 sensors have a response time/recovery time of ~7 s/12 s under the same conditions [21]. Using XPS characterization, the results showed that the oxygen vacancy content of SnO2, SnO2-D3, SnO2-D4, and SnO2-D5 was 21.11%, 22.32%, 29.47%, and 23.65%, respectively [21]. Because SnO2-D4 had the highest oxygen vacancy content, its performance was also the best. When the relative humidity is 40%, the resistance values of SnO2, SnO2-D3, SnO2-D4, and SnO2-D5 to 3 ppm H2 are1.25 MΩ, 1.56 MΩ, 1.90 MΩ, and 1.45 MΩ, respectively [21]. Even with the lower specific surface area of SnO2-D4, gas-sensitive tests show that the SnO2-D4 MEMS sensor has a superior H2 response (Ra/Rg = 2.3@6 ppm) and a very low LOD (0.1 ppm) due to the higher oxygen vacancy on the surface of SnO2-D4 [21]. The sensor is expected to be widely used in elastic catalysis, photocatalysts, and other fields [21]. Mozalev et al. prepared a practical MEMS H2 sensor with a fast response (2–3 min) and high sensitivity (Ra/Rg = 1300@1000 ppm) using a nanostructured pore anodized aluminum oxide template WO3 layer [22]. The preparation process is shown in Figure 3a, including forming the Si3N4 layer on the silicon wafer, adding the polysilicon heater and SiO2 isolation layer to prepare the photoresist mask, sputtering the Al/Ti double layer on the photoresist mask to form porous anodic alumina (PAA) film, expanding the Al2O3 nanopore, and depositing the WO3 layer on the PAA film., annealing in the layer of SiO2 and forming a contact opening towards the heating end, forming a staggered electrode and an electrode contact pad, and etching the back wafer side to complete the gas-sensitive film [22]. The sensor developed can achieve low cost, low power consumption, and high capacity, which is conducive to the economic and environmental protection of hydrogen-based energy [22]. Ong et al. prepared a p-n junction diode-based ZnO nanowire MEMS H2 sensor by ammonia plasma modification [23]. ZnO nanowires are synthesized by the low-temperature hydrothermal method and transferred to a vinyl terephthalate (PET) substrate [23]. Three methods used in this process are shown in Figure 3b (slide transfer method, roll transfer method, and heat transfer method) [23]. The results show that the conductivity of ZnO nanowires synthesized by hydrothermal synthesis can be effectively adjusted by surface modification without heat treatment by using ammonia plasma, which guarantees the stability of the sensor at low temperatures [23]. At room temperature, when the H2 concentration is 2500 ppm, the sensor response value ∆R/R0 is 27% [23].

3.2.2. Metal/MOS

Rashid et al. reported that a metal nanotube array was synthesized by a new low-temperature wet chemical process, and the tubular Pd nanostructure was directly formed on the sensor device by in situ dissolution of the ZnO nanowire template grown on the electrode surface by hydrothermal growth [24]. Figure 3c displays that under different bending conditions, the response value of the flexible sensor changes with the concentration of H2 [24]. Due to the higher surface reactivity in the bending state, the sensor’s response value is improved (∆R/R0 = 93.1%@1000 ppm, at 90° bend), which can accept more H2 molecules for active adsorption [24]. Zhang et al. developed a Pt/TiO2 sensor, and the addition of Pt improved the response rate of the TiO2 sensor when exposed to H2/O2 [26]. The surface of the sensing layer was modified with a solution containing H2PtCl6 and treated at different temperatures and times [26]. The results show that the response time of the sensor exposed to H2 and O2 for 2 h at 900 °C is 40 ms and 20 ms at the operating temperature of 500–800 °C [26]. Experimental results show that sensors with more platinum particles dispersed on titanium dioxide exhibit higher response values at low temperatures [26]. Zhang et al. produced an MEMS H2 sensor based on a Pt-modified Nb-doped TiO2 sheet, aiming to solve the challenge of MEMS sensors working in room-temperature and low-oxygen environments [27]. As can be seen from Figure 3d, when hydrogen is adsorbed on a pure (001) TiO2 plate, H2 molecules are more inclined to be adsorbed and stay at the active site of Ti (4) [27]. After hydrogen is adsorbed in the heterogeneous structure of Pt/TiO2, it will be catalyzed by Pt to split into atom H, and the split atom H tends to bind with Ti (4), which makes the sensor respond faster and the performance more stable [27]. The sensor can be used to detect hydrogen released during the charge and discharge of lithium-ion batteries (LIBs), addressing potential safety risks [27]. The Pt/TiO2 sensor achieves the advantages of compact size (0.05 cm3), low power consumption (0.1 mW at room temperature), excellent sensing performance, and easy integration [27]. The 90% RH response value can reach 12.3, and the sensor’s response and recovery time can reach 31 s/270 s at room temperature (40 °C) and 1000 ppm H2. On the surface of Pt/TiO2, H is more inclined to combine with Ti(4), which leads to a more stable structure [27]. Due to the diffusion of H atoms into the TiO2 lattice, the energy band structure of TiO2 changes, making the sensor have excellent sensing performance for hydrogen [27]. Kadhim et al. prepared high-quality nanocrystalline SnO2/Pd metal film MEMS H2 by the sol–gel method [28]. Figure 3e shows palladium mesh contacts deposited on nanocrystalline tin oxide films [28]. Metal–semiconductor–metal gas sensor components are manufactured by RF sputtering palladium grids on nanocrystalline tin oxide films [28]. The mask contains two electrodes, each consisting of four fingers, with a space of 0.4 mm between the two adjacent fingers and a width of 0.35 mm for each finger [28]. The nanocrystalline SnO2 film produced by adding glycerol has high porosity and good sensitivity, and the addition of Pd further improves the sensing performance of the sensor [28]. At an operating temperature of 75 °C, the sensor has a sensitivity of 165 for 1000 ppm H2 [28].

3.2.3. MOS/MOS

Guo et al. synthesized MOF-derivative In2O3 by in situ coupling a carbon layer and WO3 and prepared an MEMS H2 sensor using the porous heterostructure WO3-C/In2O3 as a sensing material [29]. The self-assembly method is employed to make the c/h-In2O3 film composed of particles more compact, uniform, and continuous. The three-dimensional layered porous structure of WO3-C/In2O3 sensor material, which can promote H2 transmission and diffusion, is shown in Figure 3f [29]. The high sensitivity is mainly attributed to the high carrier mobility, low activation energy, and intrinsic noise brought by the C/MOS nanocomposites [29]. At the same temperature, when the mass ratio of WO3 to In2O3 is 9 wt% (WCI-9), compared with WCI-7 and WCI-11, the sensor has a higher response value (Ra/Rg = 10.11@1000 ppm), and the sensor reaches the best working state at 250 °C [29]. The sensor has a low LOD for H2 (5 ppm), and the response value increases with increasing H2 concentration and remains stable over time. In addition, the WCI-9 sensor has a fast response/recovery speed of 1.9/9.2 s@220 ppm [29].

3.3. Carbon-Based Material

Weber et al. reported a high-performance flexible MEMS H2 sensor based on decorated Pd nanoparticles and single-wall carbon nanotubes [16]. The sensor is highly selective to H2, and the film responds best to H2 at 200 °C operating temperature [16]. In the range of H2 concentrations from 10 ppm to 50 ppm, the response value increases with the concentration increase. When the H2 concentration in the air is 0.05%, the response value reaches 75% [16]. After detection, Tres and Trec are 3 s and 8.3 s, respectively [16]. Teleki et al. reported a flexible H2 sensor with an electrodeposition electrode based on a Pd-cluster-modified graphene electrode prepared by chemical vapor deposition [17]. The deposited palladium nanocluster (FPNC)–CG electrodes are sensitive to H2 at room temperature, and the performance will be improved with the increase in FPNC population [17].

3.4. Silicon-Based Material

Michaud et al. reported an MEMS H2 gas sensor based on a 3C-SiC microcantilever with a detection range of 0.2–2% [19]. The chemical adsorption of H2 molecules by the SiC layer causes a drastic change in the electronic properties of the SiC film [19]. This change is used to measure the concentration of H2 [19]. The 3C-SiC cantilever manufacturing process consists of five lithographic steps (Figure 4a): mode alignment intersections; define the cantilever geometry. An isolation layer is introduced to separate the contact between 3C-SiC and metal. For electromagnetic drive and induction detection of metal deposition, the cantilever is released from the rear by etching a silicon substrate with potassium hydroxide [19]. The sensor does not have a functionalized coating, which avoids a series of failure problems such as equipment aging, low reliability, and high response time that may be caused by the sensitive layer [19]. Cho et al. prepared a diode-type MEMS H2 sensor by releasing SiNMs from a rigid substrate, transferring them directly to a flexible substrate, and then performing metal deposition [7]. The flexible Pd/SiNM surface diagram and energy bend diagram reveal the sensing mechanism of H2 (Figure 4b) [7]. The cathode electrodes interlace; Pd and SiNMs form Schottky contact [7]. When the sensor is exposed to a H2 environment, the hydrogen atoms decomposed by H2 gas molecules diffuse into the Pd layer, forming PdHx at the Pd/SiNM interface, and the Schottky barrier decreases [7]. The manufacturing process is simple and suitable for the chip level, with a wide detection range (50–5000 ppm), high sensitivity (∆R/R0 = 764%@500 ppm), and very low power (nW range) [7]. Choi et al. prepared a chemically gated transistor gas sensor based on SiNW, which is topped with a SnO2 film [20]. Figure 4c shows a SiNW FET schematic with a bottom gate structure for H2 sensing and scanning electron microscopy (SEM) images of Pd nanoparticles deposited on SiNW [20]. Here, SiNW is used as the bottom gate of the field-effect transistor. The sensor enables the decoupling of chemically sensitive areas from conductive channels to reduce drive voltage and improve reliability and battery supply capacity for applications in mobile and wearable sensor platforms [20]. They demonstrated the sensor’s operation at 1 V for mobile applications. The sensor can selectively detect H2, H2S, NO2, and other gases [20].

3.5. MOF Material

Weber et al. developed an MEMS H2 sensor based on a combination of Pd nanoparticle-modified ZnO nanowire and molecular sieve metal–organic skeleton nanomembrane (ZIF-8) [30]. Pd-NPs enable the sensor to achieve maximum signal response, while the ZIF-8 coating has excellent selectivity for H2, and ZIF-8/Pd/ZnO nanostructured materials significantly enhance the sensor sensitivity (Rg/Ra = 6.2@50 ppm) [30].

4. Carbon Monoxide Sensors

Carbon monoxide is a typical toxic gas produced by incomplete combustion, and it can bind with hemoglobin and cause harm to human health. The working principle of CO sensors is usually based on chemical reactions such as electrocatalytic combustion or electrochemistry. The performance of some carbon monoxide sensors is shown in Table 2. MEMS CO sensors can be widely used in many fields, such as in the biomedical field to ensure human health, in the industrial field to monitor the petrochemical manufacturing process, and in the monitoring of atmospheric quality in the environment. Its good sensing performance is of key significance for safety and environmental protection applications, and the market size is also growing year by year. For example, Daly reported a new infrared MEMS sensor technology prepared by using a silicon-based, heat-isolated suspension bridge structure, which can be used for environmental monitoring of industrial pollutants (CO, CO2, NOx, etc.) [31].
The working mechanism of the CO sensor is based on the principle of a chemical reaction or physical reaction according to the different materials. The working principle of the chemical sensor is based on the oxidation reaction of CO with MOS, polymer, and other gas-sensitive materials. For example, when MOS material is exposed to CO, carbon monoxide molecules react with adsorbed oxygen ions on the surface of the material, as shown in reaction Formulas (4)–(6), electrons are re-released and migrate to the conduction band of the semiconductor material, and the energy band of the material is bent and the potential barrier is reduced, thus reducing the resistance [32].
O2(ads) + 2CO(gas) = 2CO2(gas) + e
O(ads) + CO(gas) = CO2(gas) + e
O2−(ads) + CO(gas) = CO2(gas) + 2e
A physical sensor performs measurements using the adsorption and desorption process of CO and a specific substance. Gas-sensitive materials such as MOS or CBM are used as adsorbents. When CO enters the sensor, it will be adsorbed on the surface of the adsorbent. When the concentration of CO increases, the concentration of CO on the adsorbent surface also increases. By measuring the change in CO concentration on the surface of the adsorbent, the CO concentration can be obtained indirectly.
Table 2. Performance comparison of carbon monoxide MEMS gas sensors and sensitive materials.
Table 2. Performance comparison of carbon monoxide MEMS gas sensors and sensitive materials.
Sensitive Material CategorySensitive MaterialFabrication TechniqueWorking Temperature (°C)Detection Range (ppm)Response Time (s)Recovery Time (s)SensitivityRef.
MOSTiO2 nanoparticleDrop-coating method5001–7530–60550N/a[33]
MOSNanocrystalline SnO2Sol–gel synthesis method450N/a106114N/a[34]
Metal/MOSAl/TiO2 nanopowderCombustion method600100–500N/aN/aN/a[35]
Metal/MOSPt/SnO2 nanoparticleIn situ deposition3508–50N/aN/aN/a[36]
Metal/MOSAl/ZnOSol–gel technique300507301.6@50 ppm b[37]
Metal/MOSCa/ZnO thin-film-coated langasite lanthanum gallium (LGS)Spin coated4001000871322.469 kHz/ppm d[38]
CBMSW-defect grapheneDrop coatingRoom
temperature
N/aN/aN/a35.25% c[39]
PolymerFerrocene–chitosanDrop-casting methodRoom
temperature
0–20003864108.85 Hz/ppm d[40]
PolymerCryptophane-AElectrospray method80N/aN/aN/a0.004 Hz/ppm d[41]
Metal/polymerFe-Al-doped PANI thin filmVacuum depositionRoom
temperature
10–150510800@150 ppm a[42]
Metal/polymerPDPP4T-T-Pd (II)Air–water interface coordination reactions of thymine groups with ionsRoom
temperature
0.01–100N/aN/aN/a[43]
Polymer/polymerPoly (styrenesulfonate)/polyvinylpyrrolidone (PEDOT/PSS/PVP)Traditional electrospinningRoom
temperature
50N/aN/a−54 Hz/ppm d[44]
a R = Ra/Rg. b R = Rg/Ra. c R = |exp [(E′g − Eg)/2kT] − 1|. d Δf = 2f02 Δm/A √ρqμq.

4.1. MOS Material

4.1.1. MOS

Teleki et al. used flame spray pyrolysis (FSP) to prepare nanostructured anatase TiO2 for improving the performance of MEMS CO sensors [33]. Under the heat treatment of 900 °C, the conversion of anatase to rutile means the transformation of N-type to P-type sensing behavior. The original anatase sensor was unable to detect CO, while the rutile sensor could [33]. The operating temperature of the sensor is 500 °C, the detection range is 1–75 ppm, and the response time and recovery time are 30–60 s and 550 s, respectively [33]. The left axis of Figure 5a shows the signal change in the sensor exposed to 15–35 ppm CO after heat treatment at 900 °C [33]. It can be observed from the figure that in this range, the sensor signal decreases as the CO concentration increases [33].

4.1.2. Metal/MOS

Choi et al. prepared nanopowders of Al-doped TiO2 ceramics by the citrate–nitrate automatic combustion method and used the powders to make MEMS CO sensors [35]. Figure 5b–d illustrates that upon the introduction of aluminum doping at concentrations of 0 wt%, 5 wt%, and 7.5 wt%, the sensor response is modulated by variations in CO concentration subsequent to the samples being subjected to calcination at temperatures of 700 °C, 800 °C, and 900 °C, respectively [35]. The resistance of TiO2 samples doped with Al is lower than that of pure TiO2, and the sensor response value reaches the maximum at 600 °C [35]. In addition, the sensor can also be used to detect O2 concentration [35]. Madler has fabricated a high-performance MEMS CO sensor with PT-doped SnO2 nanoparticles [36]. The gas-sensitive material is prepared by flame spray pyrolysis (FSP) and has a response value of 8 to 50 ppm CO in dry air at an operating temperature of 350 °C [36]. The sensor has high sensitivity, a low detection limit, and good reproducibility [36]. Changes in the thickness of the film affect the resistance of the sensor and can be used to improve the performance of the sensor [36]. Hjiri et al. developed an MEMS sensor using Al-doped ZnO (AZO) as a sensitive material using an improved sol–gel technique [37]. Since Al3+ has a smaller ion size compared to Zn2+, Al doping results in an increase in conductivity and a decrease in resistivity, significantly increasing the sensitivity of the sensor to CO gas [37]. According to the difference in Al doping content, the optimal operating temperature changes, and A3ZO obtains the best response at 300 °C [37]. In the detection range of low CO concentration (5–50 ppm), the response value shows an increasing trend with the increase in concentration (Figure 5e) [37]. When the temperature is 300 °C and the concentration is 50 ppm, the response reaches 80%, the sensitivity is 1.6, and the response time/recovery time is 7/30 s [37]. Anukunprasert et al. introduced a Ca/ZnO thin-film MEMS CO sensor coated with LGS [38]. The operating temperature is 400 °C, the operating frequency is 7.89 MHz, the detection range is wide (1000 ppm), and the tres and trec are 87 s and 132 s, respectively [38].

4.2. Carbon-Based Material

Tian et al. developed an MEMS CO sensor based on a two-dimensional (2D) plane of graphene with a large amount of gas adsorption at the active site [39]. In this paper, they studied different forms of graphene, including intrinsic graphene, S-W-defect graphene, and multi-vacancy-defect graphene, and used first principles based on density functional theory to judge the ability of this graphene to adsorb and detect CO [39]. The results show that the introduction of defects improves the sensitivity of graphene to CO and CO2, S-W defects make the sensitivity of the CO sensor reach 35.25%, and multi-vacancy defects make its sensitivity reach 4.14% [39].

4.3. Polymer Material

4.3.1. Polymer

Bayram synthesized ferrocene branched-chain chitosan coating by the drip coating method and applied it to an MEMS CO sensor [40]. Because chitosan can enhance the sensing performance of CO at room temperature, a ferrocene-shell polycarbon alloy is a suitable material for making CO sensors [40]. The sensor works well at room temperature with a response time and recovery time of 38 s and 64 s, respectively [41]. The response of this CO sensor is linearly related to the gas concentration, and the detection range is 0–2000 ppm, and the response value increases as the carbon monoxide concentration increases [40]. The sensor is a quartz crystal microbalance (QCM) sensor, and its sensitivity is calculated using the Solbrey formula, Δf = 2f02 Δm/A √ρquq. Δf and Δm represent the shift in quartz resonance frequency (Hz) and mass change (gm) related to surface adsorption on QCM. f0 is the fundamental frequency (Hz) of quartz; A denotes the active area (cm2) of the QCM rigid film on the electrode; ρq and uq stand for the density (g/cm3) and shear modulus (Pa) of the piezoelectric quartz crystal, respectively. As depicted in Figure 5f, an electrochemical quartz crystal microbalance (EQCM) was employed to ascertain the shift in the resonant frequency of the QCM electrode [40]. By calculation, the sensitivity of the sensor Δf = 108.85 Hz/ppm [40]. Ping et al. reported a QCM gas sensor prepared by electrospray deposition of cryptonuclide A, which can be used to detect CH4, CO, and other gases [41]. When used as a CO sensor, it operates at 80 °C and has a sensitivity of 0.004 Hz/ppm [41].

4.3.2. Metal/Polymer

Dixit et al. prepared metal halide-doped polyaniline (PANI) films by vacuum deposition for efficient and rapid detection of CO [42]. The sensor demonstrated its highest sensitivity to CO in different gas environments (Ra/Rg = 800@150 ppm) [42]. Additionally, it showed response and recovery times of 5 s and 10 s, respectively, when exposed to CO [42]. Yang et al. incorporated a thymine group into the side chain of the diketopyrrole pyrrole (DPP)-based conjugated polymer PDPP4T-T to prepare MEMS gas sensors [43]. PDPP4T-T films have better crystallinity and higher charge mobility than similar polymers with pure alkyl chains [43]. Pd (II) or Hg (II) ions were incorporated into PDPP4T-T films by coordinating thymine with the ions at the air–water interface [43]. As shown in Figure 5g, the PDPP4T-T film and the Pd (II) or Hg (II) ions are transferred to the substrate by dropping a chloroform PDPP4T-T (0.1 mg/mL) solution onto the surface of an aqueous solution containing K2PdCl4 or Hg (ClO4)2, respectively [43]. PDPP4T-T thin-film field-effect transistors (FETs) containing Pd (II) ions are used to detect CO with a detection limit of 10 ppb, which has high sensitivity and selectivity for CO [43]. Moreover, PDPP4T thin-film FETs containing Hg (II) ions can be used to detect H2S with a detection limit as low as 1 ppb [43].

4.3.3. Polymer/Polymer

Zhang et al. successfully synthesized poly (3,4-vinyldioxythiophene)/poly (styrene sulfonate)/polyethylpyrrolidone (PEDOT/PSS/PVP) composite nanofibers by electrospinning and used them to develop QCM CO sensors [44]. At room temperature, the resistivity of PEDOT/PSS/PVP nanofibers is 10~5 Ω·m. At low CO concentrations (5–50 ppm), the response of PEDOT/PSS nanofibers is linear with CO concentration [44]. When the CO concentration exceeds 50 ppm, the sensor’s adsorption capacity reaches saturation, and the resonant frequency range does not change [44].

5. Nitrogen Dioxide Sensors

Nitrogen dioxide, as a pollutant gas, often reacts violently with many organic compounds. It is one of the causes of ozone formation and plays an important role in the formation of acid rain. Therefore, the research and development of NO2 gas sensors is of great significance and can be applied to environmental protection, the manufacturing industry, scientific research, teaching, and other fields. As shown in Table 3, the MEMS NO2 sensors developed in recent years often have the advantages of excellent stability, fast response speed, high sensitivity, and good convenience. The following examples will be introduced.

5.1. MOS Material

5.1.1. MOS

Zhang et al. prepared ZnO-450, ZnO-600, and ZnO-750 layered porous zinc oxide materials with ZIF-90 that can promote gas transport and diffusion by the simple solution precipitation method using a zeolite imidazolate framework (ZIF-90) and the pyrolysis method at different temperatures [45]. They are applied as gas-sensitive materials for high-sensitivity MEMS NO2 sensors [45]. ZnO-450 is a typical n-type metal oxide semiconductor material. The schematic diagram of its gas-sensing mechanism is shown in Figure 6a [45]. When the ZnO-450 gas sensor is exposed to air, oxygen can be adsorbed on the surface of ZnO-450 to generate chemisorbent oxygen by trapping electrons in the ZnO-450 conduction band [45]. The electrons are reduced, and an electron depletion layer (EDL) is formed on the surface of ZnO-450 [45]. Then, when the ZnO-450 MEMS sensor is exposed to NO2, continues to capture electrons in the conduction of ZnO-450 and reacts with chemisorbent oxygen, resulting in a further increase in the thickness of the EDL on the ZnO-450 surface [45]. This sensor’s response to NO2 is higher than that of other gases, and the ZnO-450-based MEMS sensor shows better gas sensitivity at lower operating temperatures (190 °C) compared to ZnO-600 and ZnO-750 [45]. The response/recovery time is 9/26 s, and the resistance amplitude of the sensor gradually increases with the increase in NO2 concentration, showing its good reversibility [45]. As the concentration of NO2 increases in the range of 0.05–40 ppm, the response of the sensor shows an upward trend [45]. According to the experimental data, a high response value (242.18%@10 ppm) is obtained [45]. The LOD of NO2 ZnO-450 is 35 ppb based on the definition of LOD and the experimental results [45]. Geng et al. developed ZnO1-x material with a hexagonal phase and a porous layered structure for the preparation of MEMS NO2 sensors [47]. They deposited a two-dimensional sheet of zinc oxide coating composed of nanoparticles in three steps on a flexible polypropylene paper equipped with gold electrodes [47]. The two-dimensional sheet coating composed of vinamil particles having a large surface–volume ratio and high porosity, which is conducive to the adsorption, desorption, and diffusion of gases (Figure 6b) [47]. The coating responds well to NO2 at room temperature (∆R/R0 = 2.61@1 ppm) [47]. Rana et al. reported a surface acoustic wave (SAW) resonator operating at 99.4 MHz using piezoelectric ST-cut quartz. The resonator device has been successfully integrated with the lead zirconate titanate (PZT) sensing layer and is highly selective to NO2 [48]. As indicated in Figure 6c, the process of depositing amorphous PZT films was executed utilizing a pulsed laser deposition (PLD) system [48]. The sensor device has a wide detection range (80–250 ppm) and high sensitivity (9.6 kHz/ppm) [48]. Hsueh et al. prepared CuO-NWs using RF sputtering technology for the preparation of MEMS NO2 gas sensors [49]. CuO-NWs grow vertically on the MEMS structure, in a process known as the competitive growth model [49]. As displayed in Figure 6d, the growth of CuO nanostructures on the substrate can be divided into three stages: initial growth, merging, and competitive growth [49]. In an environment with a NO2 concentration of 500 ppb and an operating temperature of 119 °C, the sensor has an average response of about 50.1% [49]. Studies have shown that CuO-NW/MEMS sensors have stability and repeatability for NO2 gas [49].

5.1.2. Metal/MOS

Hsueh et al. applied ultrasonic grinding technology to prepare Co3O4-NP materials with surface adsorption of Au-NPs with a diameter of about 1 nm to achieve a high sensitivity of MEMS NO2 sensors [50]. As shown in Figure 6e, this sensor makes it easier for NO2 molecules to react with Co3O4 surface electrons to improve sensitivity through temperature effect, nano-size effect, and Au-NP effect [50]. The response values of the sensor were measured at 100 ppb NO2 concentration and 114 °C, 136 °C, 157 °C, and 178 °C [50]. The sensor achieves a maximum response of 33% at an operating temperature of 136 °C [50]. The Tres/Trec of the sensor under this condition is 84 s/68 s, and the response is better than that of other MEMS gas sensors with Co3O4 as the sensitive material [50]. The sensing response of the Au/Co3O4-NP/MEMS sensor has a linear functional relationship with NO2 concentration, and the relationship between the two is exponential in the range of 10–10,000 ppb [50]. When the NO2 gas concentration is 10 ppm, the sensor response value is 225% [50]. Drmosh et al. reported a SnO2 film with highly uniformly dispersed Au nanoparticles on its surface [50]. The film was prepared by two-step annealing after sputtering and has high selectivity for NO2 [51]. It can be seen from Figure 7a that Au nanoparticles can be used as a catalyst leading to oxygen dissociation, which in turn promotes the adsorption of oxygen ions through the overflow process and improves the NO2 response characteristics of the sensor [51]. The response rate of the NO2 sensor at room temperature is 90%@50 ppm, which is 3.2 times and 5.5 times of the original SnO2 sensor and the SnO2 sensor with the Au layer, respectively, because the addition of highly dispersed Au nanoparticles significantly improves the light transmittance and crystallinity of the SnO2 film [51]. Wang et al. describe a heterogeneous Au/SnO2/NiO film for the preparation of MEMS NO2 sensing [52]. Figure 7b reveals the sensing mechanism of the sensor. When the Au/SnO2/NiO composite film is exposed to a NO2 environment, NO2 molecules react with surface chemisorbed oxygen to form adsorbed NO2(ads) [52]. Because NO2 is much more electronegative than oxygen, NO2 can also extract electrons in both N-type SnO2 and P-type NiO [52]. The rapid decline in electron concentration in SnO2 leads to the expansion of the depletion region, which enables the detection of NO2 concentration. A method of preparing sensing materials by sputtering SnO2 was reported: sputtering SnO2 targets on self-assembled Au-NP arrays and then annealing them with H2 [52]. This method facilitates electron transfer in sensing materials, allowing MEMS-compatible heterogeneous fabrication methods to have wider applications in wafer-scale gases [52]. The sensor has a high response (185@5 ppm), high selectivity, and high stability with an LOD of 50 ppb due to the catalytic action of Au-NPs and the effective Schottky barrier and p-n junction formation [52].

5.1.3. MOS/MOS

Jian et al. used an RF magnetron co-sputtering system to prepare Ta/In2O3 films with different Ta contents for the research and development of MEMS NO2 sensors [53]. The test results show that when the content of Ta is 2.33 at.%, the response and recovery time are relatively short, and the sensitivity is high [53]. At the same time, they also used a laser interferometric lithography system to model the performance of Ta/In2O3 rod arrays of different sizes [54]. The research results show that the number of rods and the surface-to-volume ratio increase with the decrease in the rod array period [54]. Using a sensor film with a 0.6 mm cycle Ta/In2O3 rod array, the NO2 gas sensor has a maximum response rate of 76.1, an optimal operating temperature of 110 °C, an LOD of 0.7 ppm, and a minimum response time and recovery time of 48 s and 329 s, respectively [54]. Wei et al. created a flexible MEMS NO2 sensor using amorphous rhodium oxide-decorated black indium oxide (RhOx/B-In2O3) as the sensing material [55]. The sensor exhibits excellent gas selectivity for NO2 at 5 ppm, high sensitivity (∆R/R0 = 42@5 ppm), and very short response/recovery time (8 s/17 s) at room temperature due to the presence of defects in the B-In2O3 material and the chemical sensitization and electron-rich properties of the amorphous RhOx [55]. The research results show that RhOx/B-In2O3 composites have broad application potential in advanced gas sensing [55]. Yempati et al. developed MEMS NO2 sensors with TeO2-doped ZnO nanostructures by co-sputtering technology [56]. The sensor mechanism, which is mainly based on the formation of an electron depletion layer and the formation of a heterojunction between ZnO and TeO2, is shown in Figure 7c [56]. EDS mapping analysis confirmed that doped TeO2 may increase the base resistance value of the sensor. Compared with 2%ZnO-TeO2 and 4%ZnO-TeO2, the 8%ZnO-TeO2 material has the most uniform deposition on the entire surface, and the added TeO2 produces more absorption sites and more oxygen ions on the surface of the material. Therefore, the response to NO2 is more sensitive [56]. Testing revealed that the selectivity of the 4%TeO2-ZnO sensor for NO2 is much higher than that for other gases, and the Tres/Trec is 13/38 s [56]. At the optimum operating temperature of 100 °C, the response to NO2 at 200 ppb and to NO2 at 1 ppm reaches 42%, and the response degree increases with the increase in concentration in this interval [56]. The performance of TeO2-doped ZnO sensors is better than that of pure ZnO sensors [56].

5.2. Carbon-Based Material

Chung et al. fabricated a flexible MEMS NO2 sensor based on WO3 NTS-MWCNT-RGO mixed on a polyimide/polyethylene terephthalate (PI/PET) substrate [61]. The manufacturing procedure, as depicted in Figure 8a, entails the employment of photolithography and RF magnetron sputtering techniques to deposit a pair of aurous electrode fingers onto the PI/Si substrate [61]. Subsequently, the amalgamated droplets are meticulously positioned between these electrodes and subjected to drying on a heated plate [61]. Following this, the sample undergoes annealing treatment [61]. Ultimately, the PI tape is transferred onto the PET substrate [61]. The prepared sensor has good nitrogen dioxide sensing performance with a low LOD of 1 ppm and a sensitivity of 17%@5 ppm [61]. Even after a curvature angle of 90° and a bending/relaxation process are employed, the sensor showed no significant performance degradation. The research results show that the sensor has the advantages of high sensitivity, high mechanical flexibility, light weight, and economy [61]. Hua et al. investigated the response of SWNT-Fe2O3 composite films obtained by a simple annealing process to NO2, H2S, and other gases [62]. The uniform distribution of Fe2O3 nanoparticles in the porous film helps to improve the gas-sensitive properties of the material, while enabling it to sense more gases and eliminating conventional steps such as chemical functionalization or doping [62]. As a result, MEMS gas sensors prepared from this material can produce a stable response to H2S and exhibit an enhanced sensitivity to NO2 at room temperature (∆R/R0 = 18.3%@100 ppm) [62]. The film can be applied to sensors in configurations that directly fabricate large-area, flexible, or wearable films or fabrics [62]. Li et al. used a simple and cost-effective two-step hydrothermal and lyophilization strategy to prepare 3D SnO2/rGO composites with extremely large surface areas and stable nanostructures [63]. As shown in Figure 8b, two different tin salt precursors, Sn2+ and Sn4+, were utilized to form SnO2/rGO nanocomposites by hydrothermal reaction [63]. The gas sensor made of the composite material has high selectivity and good gas sensitivity to NO2 [63]. Detectable NO2 concentrations range from 2 ppm to 110 ppm with sensitivity up to Ra/Rg = 11.8@110 ppm [63]. Sangeetha et al. reported an MEMS NO2 sensor with molybdenum disulfide (MoS2)/graphene material as the sensitive layer [64]. The combination of MoS2 spherical nanoparticles and two-dimensional graphene (2DG) sheets creates a large active surface area, enhancing the absorption properties of gas molecules in the presence of evanescent wave light [63]. As shown in Figure 8c, in this fiber optic sensor, 2D graphene sheets are interlinked with MoS2 nanoparticles, increasing the surface area and allowing gas molecules to interact through the presence of an evanescent field, thereby increasing the sensor output [64]. This sensor offers cost-effective production, high sensitivity (61%), and rapid response/recovery time (22 s/35 s), making it suitable for a wide range of applications [64].
Recently, some memristor-based ultra-sensitive gas sensors (gasistors) have been reported for the detection of NO2, NO, NH3, and other gases. HfO2 is often chosen as the material of memristors, and conductive filaments (CFs) are used to improve the gas sensitivity and accuracy of memristors [66]. Materials such as CNTs and MOS are used as sensitive materials for gasistors. The formation of CFs causes a transition of resistance states between high resistance (HRS) and low resistance (LRS), resulting in a local rearrangement of oxygen vacancies in high electric fields [66]. This gives the gasistor a gas-triggered switch and memory function [66]. The experimental results show that the addition of CFs can effectively solve the limitations of the CNT gas sensor in terms of sensitivity, recovery process, and humidity effect [58]. For instance, Chae et al. proposed a filament-based memristor heater (MH)-embedded transparent CNT gas sensor for the detection of NO2 gas at room temperature [58]. Nanoscale conductive filaments (CFs) are used to fabricate an MH based on the insulating material hafnium oxide (HfO2) [58]. The MH uses nanoscale CFs to apply heat directly below the sensing layer, allowing for lower power consumption and higher efficiency compared to conventional gas sensors [58]. The Joule heating in the MH solves the problem of CNTs having a large surface area and easily absorbing water and improving the humidity resistance of the material [58]. In addition, the microstructure changes (oxygen vacancies and gaps) caused by grain boundary formation during annealing reduce the operating voltage and power consumption of the MH [58]. As a result, the sensor has high transmittance and low power consumption (6.15 μW), little impact on humidity changes (<7.5%), and fast response/recovery (<1 ms/30.8 μW) [58]. Ahmad et al. report a conductive-filament-based heater (CFH)-embedded CNT NO2 sensor [59]. The pulse recovery mechanism is used to control the sensor recovery process within 1 ms [59]. The virtual heating of the application is increased, resulting in less degradation of the gas-sensing response characteristics at high relative humidity (RH) levels [59]. When vacuum heating at 0.5 V is applied, CFH-embedded CNT sensors degrade only 33% in the range of 30–90% RH levels. The sensor responds to ∆R/R0 = 52.20%@50 ppm [59].

5.3. Polymer Material

Navale et al. used PPy film as a sensing material and rotary coating technology to prepare NO2 sensors with excellent performance at room temperature [57]. The sensor has a detection range of 10–100 ppm, a response time of 126 s, and a sensitivity of (Ra/Rg = 1.12@100 ppm) [57]. Gaikwad et al. deposited conductive polythiophene-modified SWNTs on Si/SiO2 substrate by a charge-controlled potentiostatic deposition method to prepare a high-performance MEMS NO2 sensor [60]. The performance of the sensor was tested by means of chemical resistance, and the test results showed that the sensitivity of the SWNT device to NO2 was enhanced by polythiophene modification. In addition, the sensor exhibits a very wide linear response range (0.01–10 ppm) [60].

5.4. MOF Material

Zhan et al. developed a polyhedral ZIF-8 nanostructured MOF material for the fabrication of MEMS NO2 sensors [65]. The gas-sensitive mechanism is shown in Figure 8d. The interaction between gas molecules and ZIF nanomaterials causes resistance changes [64]. The excellent gas-sensitive performance of the sensor is attributed to the free carrier density and high surface–volume ratio generated by the porous materials [64]. The sensor offers excellent gas-sensitive performance, including a wide detection range (10–100 ppm), high sensitivity (Ra/Rg = 118.5@100 ppm), and fast response and recovery times (113.5 s and 111.5 s) [65]. The research results of this sensor material reflect the potential of MOF materials in the preparation of MEMS NO2 sensors and further expand the application range of MOF materials with high porosity in the gas sensor industrial environment [65].

6. Hydrogen Sulfide Sensors

Hydrogen sulfide is a toxic, acidic gas commonly produced in the refining of oil, the refining of natural gas, and catabolism in nature. Inhaling H2S at certain concentrations can cause damage to human cells. With the continuous contributions of researchers, a MEMS H2S sensor with high sensitivity, requiring minimal power, having a compact size, and possessing robust anti-interference capabilities and durability has been developed so that the H2S sensor can further play an important role in protecting safety. The performance of various MEMS H2S sensors is shown in Table 4.

6.1. MOS Material

6.1.1. MOS

Patricia et al. described the preparation and characterization of a flexible H2S sensor based on TiO2-NTs, using conventional microfabrication techniques to obtain an array of interleaving gold electrodes on one side and a universal heater on the back [67]. The sensor responds well at low temperatures with a sensitivity of ∆R/R0 = 144@38 ppm, enabling monitoring of H2S concentrations ranging from 6 ppm to 38 ppm at room temperature [67]. The research results show that the sensor has a promising application prospect in portable field detection based on low-cost nanomaterials [67]. Li et al. prepared an MEMS H2S sensor based on BiFeO3 (BFO) by a simple sol–gel method and treated ferroelectric semiconductor BFO nanoparticles by corona polarization, making the sensor still show excellent sensing performance for H2S at the ppb concentration level [68]. As can be viewed in Figure 9a, the excellent sensing performance of the BFO-P4 sensor on ppb H2S is mainly due to the adjustment of the electric polarization field by corona polarization, which leads to the difference in the hole accumulation layer thickness on the surface of BFO-P4 [68]. At the optimum operating temperature of 220 °C, the response of the BFO-P0, BFO-P2, and BFO-P4 sensors to 1.2 ppm H2S gas is very stable, and the fluctuation range is controlled within 2.6% [68]. Among them, the BFO-P4 sensor has good repeatability and has the smallest amplitude fluctuation (2.1%) in response to H2S [68]. At the same time, the BFO-P4 sensor can also effectively detect H2S of only 10 ppb, with a response value of Rg/Ra = 1.03 [68]. Additionally, the sensor has been tested to have the ability to respond quickly (about 3 s) and recover (about 7 s) [68].

6.1.2. Metal/MOS

Yempati et al. prepared zinc aluminum oxide (AZO) nanomaterials through hydrothermal synthesis and RF sputtering technology for the development of MEMS H2S sensors [69]. Figure 9b shows the chemisorption process of ZnO as an N-type semiconductor, which leads to the change in resistance [69]. When the best operating temperature of the sensor is 250 °C, the sensitivity ∆R/Rg = 14%@1000 ppb. Compared with CO, SO2, NO2, and other gases, the sensor has higher selectivity for H2S gas [69]. Dong et al. manufactured an MEMS H2S sensor based on Ni-doped CeO2 [72]. As shown in Figure 9c, doping with Ni creates more oxygen vacancies, and when hydrogen sulfide reacts with oxygen ions or is oxidized, electrons return to the cerium oxide surface, resulting in a thinner EDL and lower resistance of the sensor [72]. The change in Ce0.97Ni0.03O1.97 resistance results in a significant voltage change [72]. The response time and recovery time of the sensor are only 8 s and 13 s, respectively, and the LOD can be as low as 10 ppb [72]. The sensitivity can reach 3.1@500 ppb at the optimum operating temperature of 200 °C (in 20–90 RH% for 40 days) [72]. In addition, the sensor showed a lower LOD, down to 10 ppb H2S. The increase in oxygen vacancy and the catalytic action of hydrogen sulfide oxidation make the sensor more sensitive, selective, stable, moisture-proof, and safe [72].

6.1.3. MOS/MOS

As shown in Table 4, many H2S sensors use heterojunctions between MOS/MOS materials to improve sensor performance. Cho et al. developed MEMS H2S sensors with high sensitivity and low power consumption (<6 mW) using SnO2-ZnO hybrid nanostructures as materials [74]. Porous SnO2 films can be formed on the surface of ZnO-NWs by liquid phase deposition (LPD), and ZnO nuclei can be etched at the same time [74]. Finally, porous SnO2-NTS can be formed. Compared with ZnO-NWs, the sensing performance of the synthesized SnO2-NTs is significantly improved [74]. For example, at a H2S concentration of 20 ppm, the response Ra/Rg of the ZnO-NWs sensor is 3.60, while the response Ra/R of the SnO2-NTs sensor is 21.07 [74]. In addition, due to the heterojunction effect, the sensitivity (Ra/Rg = 35.31@20 ppm) and response time (36–55 s) are further improved by forming SnO2-ZnO hybrid nanostructures [74]. Zhang et al. developed c/h-In2O3 with both rhombic corundum and cubic biphasic heterostructures [75]. A single-layer particle film MEMS gas sensor was obtained by the preparation process shown in Figure 9d [75]. First, the sensor is continuously aged in the air at 300 °C for 72 h; the target gas is dried and injected into the chamber through a syringe [75]. The gas concentration is determined according to the ratio of the injected gas volume to the chamber volume, and then the appropriate amount of liquid is extracted with a microsyringe [75]. It is then slowly injected into the hot plate of the gas chamber to form volatile gas and finally tested and calculated [75]. This metal oxide (MOS) material has abundant oxygen vacancies (61.1%) and more adsorbed oxygen (35.9%). Then, a MEMS gas sensor with a single layer of In2O3 particle film was prepared by self-assembly, and the abundant oxygen vacancies, two-phase heterojunction, and single-layer particle film are the reasons for the excellent H2S sensing performance of this c/h-In2O3 sensor [75]. As a metastable phase, h-In2O3 is easily converted into more stable c-In2O3, so the Ra value of the sensor always follows the trend of h-In2O3 < c-In2O3 < c/h-In2O3 [75]. The response value of the three sensors varies with the temperature, and the response rises continuously from 80 °C to 160 °C, reaching the best operating temperature [75]. The sensor has high sensitivity (54.4@50 ppm) and low LOD (20 ppb) [75]. H2S is diffused in the In2O3 particle layer in two different forms: interparticle diffusion and intraparticle pore diffusion [75]. The response time is positively correlated with the number of sensitive layers. When the number of layers changes from four to one, the response rate increases by 25 times, achieving an extremely fast response time (3.3 s) [75]. The excellent selectivity, remarkable repeatability, and long-term stability of the MEMS H2S gas sensor are attributed to the formation of a heterojunction between Fe2O3 and SnO2, the formation of a mat leading to an increase in chemisorbed oxygen, and the precise control of the tin oxide shell thickness and composition achieved by the ALD process [76]. The synthesis process principle and gas-sensing measurement system are shown in Figure 10a [76]. The α-Fe2O3/SnO2 core–shell NWs are ultrasonically dispersed into the deionized water of the foam iron sheet. The uniform dispersion is dripped onto the MEMS device and completely dried in the air [74]. An insulating layer is then inserted into the pair of gold sensing electrodes and a pair of platinum heating electrodes to prevent electrical crosstalk [76]. Finally, the MEMS gas sensor is prepared by connecting the MEMS device with the external circuit through a wire connection [76]. The JF02F gas-sensing measurement system was used to test the sensing characteristics of the sensor [76]. This study is based on the principle of the dynamic volumetric method, in which a dry standard target gas is dynamically mixed with high-purity air to obtain the desired concentration of the test gas [76]. Under the condition of a H2S concentration of 10 ppm and temperature of 250 °C, the sensor response reaches the peak [76]. α-Fe2O3/SnO2 core–shell nanowires with a shell thickness of 18 nm (F/S18) show the highest response of 4.3%, and Tres/Trec is 13.8/104.5 s [76].

6.2. Polymer Material

Su et al. utilized in situ photopolymerization to attach PPy and WO3 nanoparticles (PPy/WO3) to an Al2O3 substrate to prepare an MEMS H2S gas sensor at room temperature [83]. According to the equivalent profile diagram of the hydrogen sulfide gas sensor of the WO3 thin film and PPy/WO3 nanocomposite thin film (Figure 10b), it can be seen that the N-type semiconductor WO3 forms a p-n junction with the P-type semiconductor polypyrrole, whose depletion region is wider than that of WO3. Therefore, nanocomposite films are more sensitive to H2S than WO3 films [83]. The sensor based on the PPy/WO3 nanocomposite film responds ∆R/R0 to 81%@1 ppm at room temperature [83]. Microstructure observation showed that PPy was distributed in the PPy/WO3 nanocomposite membrane [83]. Therefore, a model of the barrier electron conduction of the composite material was used to determine the high response with a loss layer at the stretched PPy interface of WO3 thin films when detecting hydrogen sulfide gas adsorption at room temperature [83]. The sensor reacts well to very low concentrations of H2S gas at room temperature and is easy to manufacture [83]. Geng et al. prepared PPy and WO3 by chemical oxidation polymerization and emulsion methods, respectively, and prepared PPy/WO3 hybrid materials with different PPy mass percentages by mechanical mixing [84]. The proton doping process and N-type semiconductor effect make sensors based on the PPy/WO3 composites far more sensitive to H2S than those based on PPy or WO3 alone [84]. The recovery time and response time of the PPy/WO3 sensor to H2S are 70 s and 34 s, respectively, and its sensitivity is linear with H2S concentration, and the response to 1000 ppm H2S reaches ∆R/R0 = 60% [84].

6.3. Carbon-Based Material

Shboul et al. prepared a flexible H2S sensor using a nanocomposite mixture of In2O3, graphite (Gt) sheets, and polystyrene (PS) [85]. Nanocomposites composed of In2O3/10% Gt/17% PS are considered promising nanocomposites for the preparation of MEMS H2S sensors with excellent properties [85]. Both additives, Gt sheets and the PS modifier, help to increase the surface–volume ratio of the sensing film, improve the sensing performance of the In2O3 sensor and enhance its resistance to humidity changes [85]. Moreover, copper acetate (CuAc) is added to react with H2S gas to further improve the sensitivity and selectivity of the sensor to H2S [85]. It can be seen from Figure 10c that the improved In2O3 NP-based sensor MS2-50 keeps the gas chamber locked after each injection of H2S gas during the sensor test, showing good gas-sensitive performance [85]. The detection limit of this sensor is less than 100 ppb H2S, and its sensing performance is much better than that of some In2O3 H2S sensors, and its moisture resistance has been further improved (≈80% relative humidity) [86]. The device can be used to monitor the degradation of packaged foods, and the preparation method is a cost-effective disposable smart sensing technology [86]. Shewale et al. synthesized a low-temperature MEMS H2S sensor with Cu-doped ZnO (CZO)-decorated RGO nanostructures by the hydrothermal method [87]. As shown in Figure 10d, the formation of p-n heterojunctions between metal–semiconductor Schottky junctions and ZnO/RGO provides a large number of adsorption centers for hydrogen sulfide molecules, increasing the resistance of the sensor [87]. Due to the synergistic effect of Cu dopants and RGO, the performance of CZO/RGO sensors is enhanced compared to ZnO/RGO sensors [87]. The nanocomposite sensor has an inspection range of 136 ppb–250 ppm, a sensitivity of 0.87%@100 ppm at 24 °C, and a response time/recovery time of 14 s/32 s [87]. In addition to H2S, the sensor also shows high selectivity for H2 [87].

7. Ammonia Sensors

Ammonia is a pungent gas that can be produced by combining nitrogen and hydrogen, and it has the potential to cause irritation to the skin, eyes, and respiratory mucous membranes. Excessive inhalation will cause damage to the lungs, sensor, and eyes. MEMS technologies, including SAW technology, QCM technology, and so on, have been widely used in the research and development of NH3 sensors. QCM technology refers to the detection of gas by measuring the resonance frequency shift of quartz to detect the mass change in the sensing layer [4]. SAW technology refers to the excitation of gas monitoring by applying sine waves to a digital intersensor (IDT) deposited on a piezoelectric material [4]. The performance of devices using different technologies is shown in Table 5. The development of sensing films with different materials has further improved the stability and sensitivity of MEMS NH3 sensors.

7.1. MOS Material

7.1.1. MOS

Biskupski et al. prepared a TiO2 thin-film MEMS NH3 sensor using a new sol–gel-based hydrothermal process [88]. After annealing, the sensors were placed on a gas test bench with NH3 concentrations of 56 ppm, 103 ppm, and 156 ppm and a temperature of 350 °C for testing [88]. Anatase films have high selectivity for and sensitivity to NH3 [88]. Therefore, the sensor sensitivity Ra/Rg reaches 1@1000 ppm [86]. Qiu et al. introduce a NH3 sensor based on TiO2-NWs gasistors [89]. The MOS-based sensor has integrated sensing and memory functions that traditional gas sensors do not have. At the same time, it has good sensing performance, Tres/Trec < 1 s, and sensitivity Rg/Ra up to 164.2@1 ppm [89]. Wen et al. prepared WO3-NPs using ultrasonic grinding technology for the development of MEMS NH3 sensors [90]. The test results show that the sensitivity of the WO3-NPs gas sensor to NH3 is obviously better than that of CO, CO2, SO2, and other gases, the detection stability is good, and the sensing performance of NH3 is much better than that of the WO3 thin film and original WO3 powder sensor. When the operating temperature is 142 °C, the response Ra/Rg of the sensor is 16% at 1.3 ppm [90]. Hsieh et al. used the anisotropic wet etching MEMS technique to fabricate a WO3 MEMS NH3 sensor with a suspension structure that can reduce the heat conduction of silicon [91]. As shown in Figure 11a, working at 200 °C, the sensor layer of WO3 is more able to attract NH3 molecules, and gas molecules can also desorb, resulting in the highest sensor response [91]. The chip size of the tiny ammonia sensor is only 5 mm × 5 mm. The sensitivity ∆R/R0 is up to 252%@5 ppm, with a response time of 30 s and an LOD of 40 ppb [91]. Quy et al. reported a wet chemical synthesis of an MEMS NH3 sensor based on ZnO nanorod-coated QCM [96]. Figure 11b shows that the SAW sensor was placed in a test chamber with a volume of 2000 mL. The dynamic volumetric method was used to inject gas into the laboratory with a syringe [93]. When the response reached an equilibrium state, the lid of the small chamber was removed, and the chamber was exposed to the atmosphere in the fume hood [96]. After the chamber was cleaned and the frequency of the sensor was stabilized, ammonia gas was injected into the test chamber again to complete the ammonia sensing characteristic test [96]. The selectivity of the sensor for NH3 is significantly higher than that of CO, CO2, NO2, liquefied petroleum gas (LPG), and other gases, the frequency change for 50 ppm NH3 at room temperature is about 9.1 Hz, and the response time and recovery time vary with the concentration of NH3 [96].

7.1.2. MOS/MOS

Wang et al. fabricated ZnO/SiO2 (ZS) composite films using the sol–gel method and deposited them on the SAW resonator as the sensitive material for MEMS NH3 sensors [98]. The performance of ZnO:/SiO2 sensors in NH3 at different molar ratios (1:1, 1:2, and 1:3) was tested, and the results showed that the sensor response was best when the molar ratio was 1:2 [98]. The reaction between ammonia and water molecules is shown in Figure 11c [98]. Ammonia molecules can react with water molecules on the surface of SiO2 to initiate the proton conduction process through NH4+ [95]. This phenomenon leads to an increase in ionic conductivity, which is the reason for the good selectivity of the sensor [98]. This is attributed to the addition of SiO2, which enhances the absorption of NH3 by the film and the unique surface reaction on the composite film [98].

7.2. Carbon-Based Material

Li et al. created a new dual-layer film QCM NH3 sensor as a sensing layer [100]. The graphene oxide (GO) film acts as the isolating layer between the QCM electrode and the PANI sensing film [100]. Due to the high elastic modulus of the GO isolation layer, the quality factor of QCM is improved, the surface energy loss is reduced, and the prepared sensor has high stability and sensitivity (214 Hz/ppm) [100]. Zhu et al. reported a SAW NH3 gas sensor based on nitrogen-doped diamond-like (N-DLC) films prepared using microwave electron cyclotron resonance plasma chemical vapor deposition (ECR-PECVD) [102]. As shown in Figure 11d, the ammonia sensing mechanism of N-doped DLC films is proposed [102]. The pores in the film can trap ammonia molecules, resulting in a decrease in the porosity of the film [102]. By filling the pores in the sensitive film, the adsorbed ammonia gas molecules can lead to an increase in the area density and modulus of the N-DLC film, ultimately leading to the response caused by the elastic modulus [102]. The N-DLC films fabricated by ECR-PECVD are highly porous and physically and chemically selective in absorbing polar NH3 gas molecules [102]. The NH3 molecules adsorbed on the polar groups of N-DLC films lead to a decrease in their porosity, which leads to an increase in the elastic modulus of the films to detect NH3 [102]. Therefore, the concentration range of NH3 that can be monitored in situ by this sensor at room temperature is 10 ppb–100 ppm, and the response time/recovery time is only 5 s/29 s [102].

7.3. Polymer Material

Aditee et al. fabricated a PPy thin-film MEMS NH3 sensor through in situ chemical polymerization and employed Fourier transform infrared spectroscopy, scanning electron microscopy, and X-ray photoelectron spectroscopy for the analysis and characterization of the thin-film properties [103]. The test results show that the film is highly selective to NH3 at room temperature, the response value increases linearly with the increase in concentration in the detection range of 4–80 ppm, and the sensitivity ∆R/R0 can reach 16%@25 ppm [103]. Lin et al. fabricated a SAW gas sensor coated with Au/PPy films for monitoring low concentrations of ammonia [106]. Through the miniaturization of SAW chips and the use of polymers as sensing materials, the sensitivity of SAW gas sensors can be improved [106]. The SAW gas sensor is designed with a dual-device setup to minimize biases in detection, such as those caused by temperature and humidity [106]. The frequency offset of the SAW sensor is utilized for monitoring various concentrations of ammonia [106]. This developed SAW gas sensor demonstrates excellent repeatability and sensitivity, even when detecting low levels of NH3 [106]. The sensor has a sensitivity of 898 Hz/ppm and a detection range of 2–10 ppm [106].

7.4. Other Material

Shen et al. produced an MEMS sensor with good sensitivity, selectivity, repeatability, and reversibility for NH3 by depositing L-glutamate hydrochloride on a 128°YX LiNbO3 SAW delay line [107]. The sensor has a low LOD of 0.56 ppm at room temperature and a sensitivity of 74 Hz/ppm [107]. L-glutamate hydrochloride was degraded at a rate of only 0.01 ppm per day, indicating that the long-term stability of the sensing period could be guaranteed [107]. Since the cross-sensitivity of humidity interference is about 0.011, the effect of humidity on the device is negligible [107]. Long et al. synthesized ZnS mucosal nanostructures on ST-cut quartz SAW devices by chemical bath deposition to prepare MEMS NH3 sensors [108]. The preparation of the ZnS sensing layer on the SAW device is shown in Figure 11e [108]. The SAW unit is attached to a slide whose IDT and reflector are completely protected with polyimide tape [108]. It is then heated in a water bath in ZnS growth solution [108]. A thin film is then deposited in the sensing area, and the quartz SAW device is removed and soaked in ethanol [108]. Finally, it is ultrasonically cleaned with deionized water to remove any residue from the surface of the device [108]. Due to the larger specific surface area and more active sites of the ZnS mucosal nanostructure, the device has high sensitivity (−1.094 Hz/ppm) and high selectivity for NH3, with a response time/recovery time of 151 s/568 s [108]. Subramanian et al. utilized a surfactant-assisted solution combustion method to produce ultrafine Zn3(VO4)2 nanopowder, aiming for the advancement of fiber optic gas sensors [109]. The sensor operates at room temperature with a wide detection range (20–500 ppm), high sensitivity (0.019 μV ppm−1), and a response/recovery time of 46.8 min/59.0 min [109].

8. Conclusions and Prospect

MEMS gas sensors can detect H2, CO, NO2, H2S, NH3, acetone, formaldehyde, and other gases with their fast response, high sensitivity, small size, and easy integration, showing a good prospect of wide application and great commercial value. Materials such as MOS, MOF, CBM, SBM, and polymer materials, among others, form the sensors. These materials exhibit good sensing properties and can undergo doping, heterogeneous structure construction, and noble metal modification to enhance their performance. The previous article reveals that the exploration of MOF materials and SBMs is relatively less extensive than that of MOS materials. Each sensor type has a most suitable sensing material:
H2 Sensors: Pd/Pt-doped MOS materials showed superior performance. Hydrogen sensors made of these materials have excellent responses even to low concentrations of hydrogen. The addition of metal material changes the surface morphology and greatly improves the difficulty of hydrogen adsorption.
CO Sensors: Polymer and MOS materials demonstrated exceptional performance characteristics. So, researchers often choose them as gas-sensitive materials. There is less research on other kinds of materials, and more attention can be paid to CO sensors composed of metal materials and carbon materials in the future.
NO2 Sensors: MOS materials combined with CNTs or graphite showed remarkable performance. These materials have excellent performance with a fast response and low detection limit and can be widely used in industrial, chemical, and other fields.
H2S Sensors: Various MOS/MOS materials achieved better performance through the formation of p-n junctions. The change in electrical properties caused by heterojunction further improves the response sensitivity of H2S detection. However, these materials have poor moisture resistance. In recent years, some perovskite nanoparticle materials and hydrophobic molecular sieve materials have been used to prepare H2S sensors to improve their moisture resistance.
NH3 Sensors: MOS and MOFs were more commonly used because of their excellent gas sensitivity. MOS materials have been the mainstream choice for commercial NH3 sensors. While there have been significant advances in MOF-based gas sensor research, the development of MOFs for NH3 sensing is still in its infancy, and developing fast response sensors remains a challenge.
Although remarkable achievements have been made in scientific research and commercial use has been achieved, MEMS gas sensors still need to be improved. Some of the problems faced and their solutions, as well as development recommendations, are pointed out as follows:
(1)
The existing gas sensors often have a high response temperature, poor response, and low sensitivity at low temperatures, which limits their application. Therefore, advanced nanomaterials and 2D materials need to be explored to ensure increased sensitivity and selectivity while reducing response temperature.
(2)
Some metal oxide semiconductor materials and the noble metal materials used for modification are expensive, and their large-scale production is difficult to achieve. Appropriate MOS-based materials could be used for combination and preparation to reduce costs. In addition, it is necessary to pay attention to the development of low-cost manufacturing technology to further promote the industrialization of MEMS gas sensors.
(3)
If the size of an MEMS device can be further reduced while its performance is improved, a wider range of applications in the chip can be achieved, while costs can be reduced and production can be expanded, which also requires the development of new devices and substrate materials.
(4)
MEMS sensors can be integrated with IoT systems and intelligent technologies to achieve a wider range of applications. Some flexible MEMS can also be used to prepare wearable devices to achieve improved convenience.
(5)
In order to achieve the purpose of environmental protection, energy-saving sensors using self-powered systems can be focused on in the future. In addition, more environmentally sustainable and biocompatible materials can be considered for the development of MEMS sensors.
In the future, more new materials will be discovered, and more technologies will be developed to integrate MEMS devices with portable devices, energy units, and smart homes with tiny sizes, high sensitivity, and multiple functions.

Author Contributions

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

Funding

This work was co-sponsored by the National Natural Science Foundation of China (Nos: 12174092 and U21A20500), Program for Key Research and Development of Science and Technology in Hubei Province (grant No. 2023BEB002), and Overseas Expertise Introduction Center for Discipline Innovation (D18025).

Conflicts of Interest

Author Ming Lei was employed by the company Wuhan Micro & Nano Sensor Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wang, J.; Yang, J.; Chen, D.; Jin, L.; Zhang, Y.; Xu, L.; Guo, Y.; Lin, F.; Wu, F. Gas detection microsystem with MEMS gas sensor and integrated circuit. IEEE Sens. J. 2018, 18, 6765–6773. [Google Scholar] [CrossRef]
  2. Hashwan, S.S.B.; Khir, M.H.M.; Nawi, I.M.; Ahmad, M.R.; Hanif, M.; Zahoor, F.; Al-Douri, Y.; Algamili, A.S.; Bature, U.I.; Alabsi, S.S.; et al. A review of piezoelectric MEMS sensors and actuators for gas detection application. Discov. Nano 2023, 18, 25–67. [Google Scholar] [CrossRef] [PubMed]
  3. Algamili, A.S.; Khir, M.H.M.; Dennis, J.O.; Ahmed, A.Y.; Alabsi, S.S.; Hashwan, S.S.B.; Junaid, M.M. A review of actuation and sensing mechanisms in MEMS-based sensor devices. Nanoscale Res. Lett. 2021, 16, 16. [Google Scholar] [CrossRef] [PubMed]
  4. Asri, M.I.A.; Hasan, M.N.; Fuaad, M.R.A.; Yunos, Y.M.; Ali, M.S.M. MEMS gas sensors: A review. IEEE Sens. J. 2021, 99, 18381–18397. [Google Scholar] [CrossRef]
  5. Shwetha, H.R.; Sharath, S.M.; Guruprasad, B.; Rudraswamy, S.B. MEMS based metal oxide semiconductor carbon dioxide gas sensor. Micro Nano Eng. 2022, 16, 100156. [Google Scholar] [CrossRef]
  6. Tang, S.; Chen, W.; Jin, L.; Zhang, H.; Li, Y.; Wen, Z. SWCNTs-based MEMS gas sensor array and its pattern recognition based on deep belief networks of gases detection in oil-immersed transformers. Sens. Actuators B 2020, 312, 127998. [Google Scholar] [CrossRef]
  7. Cho, M.; Yun, J.; Kwon, D.; Kim, K.; Park, I. High-sensitivity and low-power flexible Schottky hydrogen sensor based on silicon nanomembrane. ACS Appl. Mater. Interfaces 2018, 10, 12870. [Google Scholar] [CrossRef]
  8. Demir, H.; Daglar, H.; Gulbalkan, H.C.; Aksu, G.O.; Keskin, S. Recent advances in computational modeling of MOFs: From molecular simulations to machine learning. Coord. Chem. Rev. 2023, 484, 215112. [Google Scholar] [CrossRef]
  9. Sennik, E.; Urdem, S.; Erkovan, M.; Kilinc, N. Sputtered platinum thin films for resistive hydrogen sensor application. Mater. Lett. 2016, 177, 104–107. [Google Scholar] [CrossRef]
  10. Walewyns, T.; Spirito, D.; Francis, L.A. A tunable palladium-based capacitive MEMS hydrogen sensor performing high dynamics, high selectivity and ultra-low power sensing. Transducers 2015, 87, 268–271. [Google Scholar] [CrossRef]
  11. Gong, J.; Wang, Z.; Tang, Y.; Sun, J.; Wei, X.; Zhang, Q.; Tian, G.; Wang, H. MEMS-based resistive hydrogen sensor with high performance using a palladium-gold alloy thin film. J. Alloys Compd. 2023, 930, 167398. [Google Scholar] [CrossRef]
  12. Sanger, A.; Kumar, A.; Chauhan, S.; Gautam, Y.K.; Chandra, R. Fast and reversible hydrogen sensing properties of Pd/Mg thin film modified by hydrophobic porous silicon substrate. Sens. Actuators B 2015, 213, 252–260. [Google Scholar] [CrossRef]
  13. Kondalkar, V.V.; Park, J.; Lee, K. MEMS hydrogen gas sensor for in-situ monitoring of hydrogen gas in transformer oil. Sens. Actuators B 2021, 326, 128989. [Google Scholar] [CrossRef]
  14. Yamazaki, H.; Hayashi, Y.; Masunishi, K.; Ono, D.; Ikehashi, T. A review of capacitive MEMS hydrogen sensor using Pd based metallic glass with fast response and low power consumption. Electr. Commun. Jpn. 2019, 138, 312–318. [Google Scholar] [CrossRef]
  15. Hayashi, Y.; Yamazaki, H.; Ono, D.; Masunishi, K.; Ikehashi, T. Investigation of PdCuSi metallic glass film for hysteresis-free and fast response capacitive MEMS hydrogen sensors. Int. J. Hydrogen Energy 2018, 43, 9438–9445. [Google Scholar] [CrossRef]
  16. Sun, Y.; Wang, S.; Hau, H. Electrodeposition of Pd nanoparticles on Single-walled Carbon Nanotubes for Flexible Hydrogen Sensors. Appl. Phys. Lett. 2007, 90, 213107. [Google Scholar] [CrossRef]
  17. Shin, D.H.; Lee, J.S.; Jun, J.; An, J.H.; Kim, S.G.; Cho, K.H.; Jang, J. Flower-like Palladium Nanoclusters Decorated Graphene Electrodes for Ultrasensitive and Flexible Hydrogen Gas Sensing. Sci. Rep. 2015, 5, 12294. [Google Scholar] [CrossRef] [PubMed]
  18. Gao, M.; Cho, M.; Han, H.; Jung, Y.; Park, I. Palladium-Decorated Silicon Nanomesh Fabricated by Nanosphere Lithography for High Performance, Room Temperature Hydrogen Sensing. Small 2018, 14, 170391. [Google Scholar]
  19. Michaud, J.; Portail, M.; Alquier, D.; Certon, D.; Dufour, I. Silicon-carbide-based MEMS for gas detection applications. Mater. Sci. Semicond. Process. 2024, 171, 107986. [Google Scholar] [CrossRef]
  20. Luo, N.; Wang, C.; Zhang, D.; Guo, M.; Wang, X.; Cheng, Z.; Xu, J. Ultralow detection limit MEMS hydrogen sensor based on SnO2 with oxygen vacancies. Sens. Actuators B 2022, 354, 130982. [Google Scholar] [CrossRef]
  21. Mozalev, A.; Calavia, R.; Vazquez, R.M.; Grdcia, I.; Cane, C.; Correig, X.; Vilanova, X.; Gispert-Guirado, F.; Hubalek, J.; Llobet, E. MEMS-microhotplate-based hydrogen gas sensor utilizing the nanostructured porous-anodic-alumina-supported WO3 active layer. Int. J. Hydrogen Energy 2013, 39, 8011–8021. [Google Scholar] [CrossRef]
  22. Ong, W.L.; Zhang, C.; Ho, G.W. Ammonia plasma modification towards a rapid and low temperature approach for tuning electrical conductivity of ZnO nanowires on flexible substrates. Nanoscale 2011, 3, 4206–4214. [Google Scholar] [CrossRef] [PubMed]
  23. Choi, B.; Ahn, J.H.; Lee, J.; Yoon, J.; Choi, S.J. A bottom-gate silicon nanowire field-effect transistor with functionalized palladium nanoparticles for hydrogen gas sensors. Solid State Electron. 2015, 114, 76–79. [Google Scholar] [CrossRef]
  24. Rashid, T.R.; Phan, D.T.; Chung, G.S. A flexible hydrogen sensor based on Pd nanoparticles decorated ZnO nanorods grown on polyimide tape. Sens. Actuators B 2013, 185, 777–784. [Google Scholar] [CrossRef]
  25. Lim, M.A.; Kim, D.H.; Park, C.O.; Lee, Y.W.; Han, S.W.; Li, Z.; Williams, R.S.; Park, I. A New Route toward Ultrasensitive, Flexible Chemical Sensors: Metal Nanotubes by Wet-Chemical Synthesis along Sacrificial Nanowire Templates. ACS Nano 2012, 6, 598–608. [Google Scholar] [CrossRef]
  26. Zhang, M.; Yuan, Z.; Song, J.; Cheng, Z. Improvement and mechanism for the fast response of a Pt/TiO2 gas sensor. Sens. Actuators B 2010, 148, 87–92. [Google Scholar] [CrossRef]
  27. Zhang, M.; He, Z.; Cheng, W.; Li, X.; Zan, X.; Bao, Y.; Gu, H.; Homewood, K.; Gao, Y.; Zhang, S.; et al. A room-temperature MEMS hydrogen sensor for lithium ion battery gas detecting based on Pt-modified Nb doped TiO2 nanosheets. Int. J. Hydrogen Energy 2024, 74, 307–315. [Google Scholar] [CrossRef]
  28. Kadhim, I.H.; Hassan, H.A.; Abdullah, Q.N. Hydrogen Gas Sensor Based on Nanocrystalline SnO2 Thin Film Grown on Bare Si Substrates. Nano Micro Lett. 2016, 8, 20–28. [Google Scholar] [CrossRef]
  29. Guo, M.; Luo, N.; Bai, Y.L.; Xue, Z.; Hu, Q.; Xu, J. MEMS sensor based on MOF-derived WO3-C/In2O3 heterostructures for hydrogen detection. Sens. Actuators B 2024, 398, 134151. [Google Scholar] [CrossRef]
  30. Weber, M.J.; Kim, J.H.; Lee, J.H.; Kim, J.Y.; Xiao, C. High-performance nanowire hydrogen sensors by exploiting the synergistic effect of Pd nanoparticles and metal-organic framework membranes. ACS Appl. Mater. Interfaces 2018, 10, 34765–34773. [Google Scholar] [CrossRef] [PubMed]
  31. Daly, J.T.; Johnson, E.A.; Moelders, N.; Mcneal, M.P.; Pralle, M.U.; Greenwald, A.C.; Ho, W.; Puscasu, I.; George, T.; Choi, D.S. MEMS-based sensor system for environmental monitoring. SPIE Opt. East 2002, 4576, 49–55. [Google Scholar]
  32. Wei, W.; Yang, J.; Tao, T.; Xia, X.; Bao, Y.; Lourenco, M.; Homewood, K.; Huang, Z.; Gao, Y. A CuO/TiO2 heterojunction based CO sensor with high response and selectivity. Energy Environ. Mater. 2023, 6, e12570. [Google Scholar] [CrossRef]
  33. Teleki, A.; Pratsinis, S.E.; Kalyanasundaram, K.; Gouma, P.I. Sensing of organic vapors by flame-made TiO2 nanoparticles. Sens. Actuators B 2006, 119, 683–690. [Google Scholar] [CrossRef]
  34. Bagolini, A.; Gaiardo, A.; Crivellari, M.; Demenev, E.; Bartali, R.; Picciotto, A.; Valt, M.; Ficorella, F.; Guidi, V.; Bellutti, P. Development of MEMS MOS gas sensors with CMOS compatible PECVD inter-metal passivation. Sens. Actuators B 2019, 292, 225–232. [Google Scholar] [CrossRef]
  35. Choi, Y.J.; Seeley, Z.; Bandyopadhyay, A.; Bose, S.; Akbaret, S.A. Aluminum-doped TiO2 nano-powders for gas sensors. Sens. Actuators B 2007, 124, 111–117. [Google Scholar] [CrossRef]
  36. Madler, L.; Roessler, A.; Pratsinis, S.E.; Sahm, T.; Gurlo, A.; Barsan, N.; Weimar, U. Direct formation of highly porous gas-sensing films by in situ thermophoretic deposition of flame-made Pt/SnO2 nanoparticles. Sens. Actuators B 2006, 114, 283–295. [Google Scholar] [CrossRef]
  37. Hjiri, M.; Mir, L.E.; Leonardi, S.G.; Pistone, A.; Mavilia, L.; Neri, G. Al-doped ZnO for highly sensitive CO gas sensors. Sens. Actuators B 2014, 196, 413–420. [Google Scholar] [CrossRef]
  38. Anukunprasert, T.; Saiwan, C.; Traversa, E. The development of gas sensor for carbon monoxide monitoring using nanostructure of Nb-TiO2. Sci. Technol. Adv. Mater. 2005, 6, 359–363. [Google Scholar] [CrossRef]
  39. Tian, W.; Niu, J.; Li, W.; Liu, X. Study on adsorption of CO and CO2 by graphene gas sensor. Mod. Phys. Lett. B 2021, 35, 2150154. [Google Scholar] [CrossRef]
  40. Bayram, A.; Ozbek, C.; Senel, M.; Okur, S. CO gas sorption properties of ferrocene branched chitosan derivatives. Sens. Actuators B 2017, 241, 308–313. [Google Scholar] [CrossRef]
  41. Ping, S.; Jiang, Y.; Xie, G.; Du, X.; H Jia, H. A room temperature supramolecular-based quartz crystal microbalance (QCM) methane gas sensor. Sens. Actuators B 2009, 141, 104–108. [Google Scholar]
  42. Dixit, V.; Misra, S.; Sharma, B. Carbon monoxide sensitivity of vacuum deposited polyaniline semiconducting thin films. Sens. Actuators B 2005, 104, 90–93. [Google Scholar] [CrossRef]
  43. Yang, Y.; Liu, Z.; Chen, L.; Yao, J.; Lin, G.; Zhang, X.; Zhang, G.; Zhang, D. Conjugated semiconducting polymer with thymine groups in the side chains: Charge mobility enhancement and application for selective field-effect transistor sensors toward CO and H2S. Chem. Mater. 2019, 31, 1800–1807. [Google Scholar] [CrossRef]
  44. Zhang, H.D.; Yan, X.; Zhang, Z.H.; Yu, G.F.; Han, W.P.; Zhang, J.C.; Long, Y.Z. Electrospun PEDOT: PSS/PVP nanofibers for CO gas sensing with quartz crystal microbalance technique. Int. J. Polym. Sci. 2016, 2016, 3021353. [Google Scholar] [CrossRef]
  45. Zhang, D.; Luo, N.; Xue, Z.; Bai, Y.L.; Xu, J. Hierarchically porous ZnO derived from zeolitic imidazolate frameworks for high-sensitive MEMS NO2 sensor. Talanta 2024, 274, 125995. [Google Scholar] [CrossRef]
  46. Rana, L.; Gupta, R.; Tomar, M.; Gupta, V. ZnO/ST-Quartz SAW resonator: An efficient NO2 gas sensor. Sens. Actuators B 2017, 252, 840–845. [Google Scholar] [CrossRef]
  47. Geng, X.; Zhang, C.; Luo, Y.; Debliquy, M. Flexible NO2 gas sensors based on sheet-like hierarchical ZnO1−x coatingas sensor deposited on polypropylene papers by suspension flame spraying. J. Taiwan Inst. Chem. E. 2017, 75, 280–286. [Google Scholar] [CrossRef]
  48. Rana, L.; Gupta, R.; Kshetrimayum, R.; Tomar, M.; Gupta, V. Fabrication of surface acoustic wave based wireless NO2 gas sensor. Surf. Coat. Technol. 2018, 343, 89–92. [Google Scholar] [CrossRef]
  49. Hsueh, T.J.; Li, P.S.; Fang, S.Y.; Hsu, C.L. A vertical CuO-NWS/MEMS NO2 gas sensor that is produced by sputtering. Sens. Actuators B 2022, 355, 131260. [Google Scholar] [CrossRef]
  50. Hsueh, T.J.; Wu, S.S. Highly sensitive Co3O4 nanoparticles/MEMS NO2 gas sensor with the adsorption of the Au nanoparticles. Sens. Actuators B 2020, 329, 129201. [Google Scholar] [CrossRef]
  51. Drmosh, Q.A.; Yamani, Z.H.; Mohamedkhair, A.K.; Hendi, A.H.Y.; Hossain, M.K.; Ibrahim, A. Gold nanoparticles incorporated SnO2 thin film: Highly responsive and selective detection of NO2 at room temperature. Mater. Lett. 2018, 214, 283–286. [Google Scholar] [CrossRef]
  52. Wang, Y.; Liu, C.; Wang, Z.; Song, Z.; Zhou, X.; Han, N.; Chen, Y. Sputtered SnO2:NiO thin films on self-assembled Au nanoparticle arrays for MEMS compatible NO2 gas sensors. Sens. Actuators B 2019, 278, 28–38. [Google Scholar] [CrossRef]
  53. Jian, L.Y.; Lee, H.Y.; Lee, C.T. Enhanced nitrogen dioxide gas-sensing performance using tantalum pentoxide-alloyed indium oxide sensing membrane. IEEE Sens. J. 2019, 19, 7829–7834. [Google Scholar] [CrossRef]
  54. Jian, L.Y.; Lee, C.T.; Lee, H.Y. Performance Improvement of NO2 Gas Sensor Using Rod-Patterned Tantalum Pentoxide-Alloyed Indium Oxide Sensing Membranes. IEEE Sens. J. 2020, 99, 2134–2139. [Google Scholar] [CrossRef]
  55. Wei, W.; Luo, N.; Wang, X.; Xue, Z.; Shah, L.A.; Hu, Q.; Xu, J. Amorphous RhOx decorated black indium oxide for rapid and flexible NO2 detection at room temperature. Sens. Actuators B 2024, 414, 135944. [Google Scholar] [CrossRef]
  56. Nagarjuna, Y.; Hsiao, Y.J. TeO2 doped ZnO nanostructure for the enhanced NO2 gas sensing on MEMS sensor device. Sens. Actuators B 2024, 401, 134891. [Google Scholar] [CrossRef]
  57. Navale, S.T.; Mane, A.T.; Chougule, M.A.; Sakhare, R.D.; Nalage, S.R.; Patil, V.B. Highly selective and sensitive room temperature NO2 gas sensor based on polypyrrole thin films. Synth. Met. 2014, 189, 94–99. [Google Scholar] [CrossRef]
  58. Chae, M.; Lee, D.; Kim, H.D. Dynamic response and swift recovery of filament heater-Integrated low-power transparent CNT gas sensor. Adv. Funct. Mater. 2024, 34, 2405260. [Google Scholar] [CrossRef]
  59. Ahmad, I.; Lee, D.; Chae, M.; Kim, H.D. Advanced recovery and enhanced humidity tolerance of CNTs gas sensor using a filament heater. Chem. Eng. J. 2024, 496, 154014. [Google Scholar] [CrossRef]
  60. Gaikwad, S.; Bodkhe, G.; Deshmukh, M.; Patil, H.; Rushi, A.; Shirsat, M.D.; Koinkar, P.; Kim, Y.H.; Mulchandani, A. Chemiresistive sensor based on polythiophene modified single-walled carbon nanotubes for detection of NO2. Mod. Phys. Lett. B. 2015, 29, 1540046. [Google Scholar] [CrossRef]
  61. Chung, G.S.; Yaqoob, U.; Uddin, A.S.M. A high-performance flexible NO2 sensor based on WO3 NPs decorated on MWCNTs and RGO hybrids on PI/PET substrates. Sens. Actuators B 2016, 224, 738–746. [Google Scholar]
  62. Hua, C.; Shang, Y.; Wang, Y.; Xu, J.; Zhang, Y.; Li, X.; Cao, A. A flexible gas sensor based on single-walled carbon nanotube-Fe2O3 composite film. Appl. Surf. Sci. 2017, 405, 405–411. [Google Scholar] [CrossRef]
  63. Li, L.; He, S.; Liu, M.; Zhang, C.; Chen, W. Three-dimensional mesoporous graphene aerogel-supported SnO2 nanocrystals for high-performance NO2 gas sensing at low temperature. Anal. Chem. 2015, 87, 1638–1645. [Google Scholar] [CrossRef]
  64. Sangeetha, M.; Madhan, D. Ultra sensitive molybdenum disulfide (MoS2)/graphene based hybrid sensor for the detection of NO2 and formaldehyde gases by fiber optic clad modified method. Opt. Laser Technol. 2020, 127, 106193. [Google Scholar] [CrossRef]
  65. Zhan, M.; Hussain, S.; Algarni, T.S.; Shah, S.; Liu, G. Facet controlled polyhedral ZIF-8 MOF nanostructures for excellent NO2 gas-sensing applications. Mater. Res. Bull. 2020, 136, 111133. [Google Scholar] [CrossRef]
  66. Kim, T.; Lee, D.; Chae, M.; Kim, K.H.; Kim, H.D. Enhancing the resistive switching properties of transparent HfO2-based memristor devices for reliable gasistor applications. Sensors 2024, 24, 6382. [Google Scholar] [CrossRef]
  67. Patricia, P.; Daniel, R. TiO2 nanotubes membrane flexible sensor for low-temperature H2S detection. Chemosensors 2016, 11, 3377. [Google Scholar] [CrossRef]
  68. Li, X.; Zhang, L.; Luo, N.; Chen, J.; Cheng, J.; Ren, W.; Xu, J. Enhanced H2S sensing performance of BiFeO3 based MEMS gas sensor with corona poling. Sens. Actuators B 2022, 358, 131477. [Google Scholar] [CrossRef]
  69. Yempati, N.; Lin, J.C.; Wang, S.C.; Yu, J.H. AZO-Based ZnO nanosheet MEMS sensor with different Al concentrations for enhanced H2S gas sensing. J. Nanomater. 2021, 13, 3377. [Google Scholar]
  70. Chou, C.H.; Nagarjuna, Y.; Yang, Z.C.; Hsiao, Y.; Wang, S.C. Catalytic effect of Ag embedded with ZnO prepared by Co-sputtering on H2S gas sensing MEMS device. Vacuum 2022, 202, 111210. [Google Scholar] [CrossRef]
  71. Chen, Y.; Yuan, T.; Li, Y. Ppb-level H2S sensor with super selectivity based on Fe-NiOx nanotube assembled by AAO template. Ceram. Int. 2024, 50, 22243–22251. [Google Scholar] [CrossRef]
  72. Dong, Z.; Hu, Q.; Liu, H.; Wu, Y.; Ma, Z.; Fan, Y.; Li, R.; Xu, J.; Wang, X. 3D flower-like Ni doped CeO2 based gas sensor for H2S detection and its sensitive mechanism. Sens. Actuators B 2022, 357, 131227. [Google Scholar] [CrossRef]
  73. Zhang, C.; Liu, Y.; Wang, C.; Xu, J.; Bai, Y.L. In2O3 decorated Co32 cluster for high performance H2S MEMS sensors. Inorg. Chem. Commun. 2023, 157, 111256. [Google Scholar] [CrossRef]
  74. Cho, I.; Kang, K.; Yang, D.; Yun, J.; Park, J. Localized Liquid-Phase Synthesis of Porous SnO2 Nanotubes on MEMS Platform for Low-Power, High Performance Gas Sensors. ACS Appl. Mater. Interfaces 2017, 9, 27111–27119. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, Y.; Zhang, Z.; Lv, G.; Zhang, Y.; Chen, J.; Luo, Y.; Duan, g. Ultrafast-response H2S MEMS gas sensor based on double phase In2O3 monolayer particle film. Sens. Actuators B 2024, 412, 135787. [Google Scholar] [CrossRef]
  76. Miao, X.Y.; Zhu, L.Y.; Wu, X.Y.; Mao, L.M.; Jin, X.H.; Lu, H.L. Precise preparation of α-Fe2O3/SnO2 core-shell nanowires via atomic layer deposition for selective MEMS-based H2S gas sensor. Sens. Actuators B 2023, 378, 133111. [Google Scholar] [CrossRef]
  77. Mao, L.M.; Xu, L.; Jin, X.H.; Zhu, L.Y.; Tao, W.T.; Lu, H.L. Excellent long-term stable H2S gas sensor based on Nb2O5/SnO2 core-shell heterostructure nanorods. Appl. Surf. Sci. 2022, 602, 154339. [Google Scholar] [CrossRef]
  78. He, L.; Jia, Y.; Meng, F.; Li, M.; Liu, J. Development of sensors based on CuO-doped SnO2 hollow spheres for ppb level H2S gas sensing. J. Mater. Sci. 2009, 44, 4326–4333. [Google Scholar] [CrossRef]
  79. He, H.; Zhao, C.; Xu, J.; Qu, K.; Jiang, Z.; Gao, Z.; Song, Y.Y. Exploiting Free-Standing p-CuO/n-TiO2 Nanochannels as a Flexible Gas Sensor with High Sensitivity for H2S at Room Temperature. ACS Sens. 2021, 6, 3387–3397. [Google Scholar] [CrossRef]
  80. Meng, F.N.; Di, X.P.; Dong, H.W.; Zhan, Y.; Zhu, C.L.; Li, C.; Chen, Y.J. Ppb H2S gas sensing characteristics of Cu2O/CuO sub-microspheres at low-temperature. Sens. Actuators B 2013, 182, 197–204. [Google Scholar] [CrossRef]
  81. Ramgir, N.S.; Goyal, C.P.; Sharma, P.K.; Goutam, U.K.; Bhattacharya, S.; Datta, N.; Kaur, M.; Debnath, A.K.; Aswal, D.K.; Gupta, S.K. Selective H2S sensing characteristics of CuO modified WO3 thin films. Sens. Actuators B 2013, 188, 525–532. [Google Scholar] [CrossRef]
  82. Zhang, C.; Wu, K.; Liao, H.; Debliquy, M. Room temperature WO3-Bi2WO6 sensors based on hierarchical microflowers for ppb-level H2S detection. Chem. Eng. J. 2022, 430, 132813. [Google Scholar] [CrossRef]
  83. Su, P.G.; Peng, Y.T. Fabrication of a room-temperature H2S gas sensor based on PPy/WO3 nanocomposite films by in-situ photopolymerization. Sens. Actuators B 2014, 193, 637–643. [Google Scholar] [CrossRef]
  84. Geng, L. Gas sensitivity study of polypyrrole/WO3 hybrid materials to H2S. Synth. Met. 2010, 160, 1708–1711. [Google Scholar] [CrossRef]
  85. Shboul, A.A.; Shih, A.; Izquierdo, R. A Flexible Indium Oxide Sensor with Anti-Humidity Property For Room Temperature Detection of Hydrogen Sulfide. IEEE Sens. J. 2024, 99, 9667–9674. [Google Scholar] [CrossRef]
  86. Shboul, A.A.A.; Izquierdo, R. Printed chemiresistive In2O3 nanoparticle-based sensors with ppb detection of H2S gas for food packaging. ACS Appl. Nano Mater. 2021, 4, 9508–9517. [Google Scholar] [CrossRef]
  87. Shewale, P.S.; Yun, K.S. Synthesis and characterization of Cu-doped ZnO/RGO nanocomposites for room-temperature H2S gas sensor. J. Alloys Compd. 2020, 837, 155527. [Google Scholar] [CrossRef]
  88. Biskupski, D.; Herbig, B.; Schottner, G.; Moos, R. Nanosized titania derived from a novel sol–gel process for ammonia gas sensor applications. Sens. Actuators B 2011, 153, 329–334. [Google Scholar] [CrossRef]
  89. Qiu, P.; Qin, Y.; Xia, Q. Ultrasensitive memristor-based gas sensor (gasistor) with gas-triggered switch and memory function for dilute NH3 detection. Sens. Actuators B 2022, 373, 132730. [Google Scholar] [CrossRef]
  90. Wen, P.H.; Zheng, H.Y.; Hsueh, T.J. A WO3-NPs/MEMS NH3 Gas Sensor. J. Electrochem. Soc. 2023, 170, 037506. [Google Scholar] [CrossRef]
  91. Heish, C.H.; Lin, T.W.; Juang, F.R.; Huang, I.Y. Micro-gas sensors with a suspended micro-heater for ammonia gas detection. J. Micromechanics Microeng. 2023, 13, 015004. [Google Scholar]
  92. Maricar, M.S.; Sastikumar, D.; Vanga, P.R.; Ashok, M. Fiber optic gas sensor response of hydrothermally synthesized nanocrystalline bismuth tungas sensortate to methanol. Mater. Lett. 2021, 288, 129337. [Google Scholar] [CrossRef]
  93. Wang, X.; Zhang, J.; Zhu, Z. Ammonia sensing characteristics of ZnO nanowires studied by quartz crystal microbalance. Appl. Surf. Sci. 2006, 252, 2404–2411. [Google Scholar] [CrossRef]
  94. Minh, V.A.; Le, A.T.; Huy, T.Q.; Hung, V.N.; Quy, N.V. Enhanced NH3 gas sensing properties of a QCM sensor by increasing the length of vertically orientated ZnO nanorods. Appl. Surf. Sci. 2013, 265, 458–464. [Google Scholar] [CrossRef]
  95. Quy, N.V.; Minh, V.A.; Luan, N.V.; Hung, V.N.; Hieu, N.V. Gas sensing properties at room temperature of a quartz crystal microbalance coated with ZnO nanorods. Sens. Actuators B 2011, 153, 188–193. [Google Scholar]
  96. Li, W.; Guo, Y.; Tang, Y.; Zu, X.; Ma, J.; Wang, L.; Fu, Y.Q. Room-temperature ammonia sensor based on ZnO nanorods deposited on ST-cut quartz surface acoustic wave devices. Sensors 2017, 17, 1142. [Google Scholar] [CrossRef]
  97. Zhu, Y.; Fu, H.; Ding, J.; Li, H.; Zhang, M.; Zhang, J.; Liu, Y. Fabrication of three-dimensional zinc oxide nanoflowers for high-sensitivity fiber-optic ammonia gas sensors. Appl. Optics 2018, 57, 7924. [Google Scholar] [CrossRef]
  98. Wang, S.Y.; Ma, J.Y.; Li, Z.J.; Su, H.Q.; Alkurd, N.R.; Zhou, W.L.; Wang, L.; Du, B.; Tang, Y.L.; Ao, D.Y.; et al. Surface acoustic wave ammonia sensor based on ZnO/SiO2 composite film. Microelectron. Eng. 2015, 285, 368–374. [Google Scholar] [CrossRef] [PubMed]
  99. Wu, Z.; Chen, X.; Zhu, S.; Zhou, Z.; Yao, Y.; Quan, W.; Liu, B. Enhanced sensitivity of ammonia sensor using graphene/polyaniline nanocomposite. Sens. Actuators B 2013, 178, 485–493. [Google Scholar] [CrossRef]
  100. Li, X.; Chen, X.; Yao, Y.; Li, N.; Chen, X. High-stability quartz crystal microbalance ammonia sensor utilizing graphene oxide isolation layer. Sens. Actuators B 2014, 196, 183–188. [Google Scholar] [CrossRef]
  101. Jia, Y.; Yu, H.; Zhang, Y.; Dong, F.; Li, Z. Cellulose acetate nanofibers coated layer-by-layer with polyethylenimine and graphene oxide on a quartz crystal microbalance for use as a highly sensitive ammonia sensor. Colloid. Surf. B. 2016, 148, 263–269. [Google Scholar] [CrossRef]
  102. Zhu, H.; Xie, D.; Lin, S.; Zhang, W.; Yang, Y.; Zhang, R.; Shi, X.; Wang, H.; Zhang, Z.; Zu, X.; et al. Elastic loading enhanced NH3 sensing for surface acoustic wave sensor with highly porous nitrogen doped diamond like carbon film. Sens. Actuators B 2021, 344, 130175. [Google Scholar] [CrossRef]
  103. Aditee, J.; Gangal, S.A.; Gupta, S.K. Ammonia sensing properties of polypyrrole thin films at room temperature. Sens. Actuators B 2011, 156, 938–942. [Google Scholar]
  104. Cho, J.H.; Yu, J.B.; Kim, J.S.; Sohn, S.O.; Lee, D.D.; Huh, J.S. Sensing behaviors of polypyrrole sensor under humidity condition. Sens. Actuators B 2005, 108, 389–392. [Google Scholar] [CrossRef]
  105. Malkeshi, H.; Moghaddam, H.M. Ammonia gas-sensing based on polythiophene film prepared through electrophoretic deposition method. J. Polym. Res. 2016, 23, 108. [Google Scholar] [CrossRef]
  106. Lin, T.H.; Li, Y.T.; Hao, H.C.; Fang, I.C.; Yao, D.J. Surface acoustic wave gas sensor for monitoring low concentration ammonia. In Proceedings of the 2011 16th International Solid-State Sensors, Actuators and Microsystems Conference, Beijing, China, 5–9 June 2011; Volume 1140–1143. [Google Scholar]
  107. Shen, C.Y.; Huang, C.P.; Huang, W.T. Gas-detecting properties of surface acoustic wave ammonia sensors. Sens. Actuators B 2004, 101, 1–7. [Google Scholar] [CrossRef]
  108. Long, G.; Guo, Y.; Li, W.; Tang, Q.; Fu, Y. Surface acoustic wave ammonia sensor based on ZnS mucosal-like nanostructures. Microelectron. Eng. 2019, 222, 111201. [Google Scholar] [CrossRef]
  109. Subramanian, M.; Dhayabaran, V.V.; Sastikumar, D.; Shanmugavadivel, M. Development of room temperature fiber optic gas sensor using clad modified Zn3(VO4)2. J. Alloys Compd. 2018, 750, 153–163. [Google Scholar] [CrossRef]
Figure 1. Summary of sensing materials and application areas for five MEMS gas sensors.
Figure 1. Summary of sensing materials and application areas for five MEMS gas sensors.
Sensors 24 08125 g001
Figure 2. (a) Schematic diagram for H2 ad/absorption on the surface of Pt film [9]. Copyright 2016, Elsevier. (b) Schematic diagram for hydrogen-sensitive mechanism on Pd alloy thin film [11]. Copyright 2023, Elsevier. (c) Schematic diagram of hydrogenation and dehydrogenation of Pd/Mg thin film [12]. Copyright 2015, Elsevier. (d) Fabrication procedure of the hydrogen sensor array combined with temperature sensor and planar microheater [13]. Copyright 2021, Elsevier. (e) Schematic cross-sectional diagrams of MEMS capacitive H2 sensor and sensing mechanism and sample structure for analysis [14]. Copyright 2018, IEEJ. (f) Schematic of the SnO2-D MEMS sensor formation process and production principle [21]. Copyright 2022, Elsevier.
Figure 2. (a) Schematic diagram for H2 ad/absorption on the surface of Pt film [9]. Copyright 2016, Elsevier. (b) Schematic diagram for hydrogen-sensitive mechanism on Pd alloy thin film [11]. Copyright 2023, Elsevier. (c) Schematic diagram of hydrogenation and dehydrogenation of Pd/Mg thin film [12]. Copyright 2015, Elsevier. (d) Fabrication procedure of the hydrogen sensor array combined with temperature sensor and planar microheater [13]. Copyright 2021, Elsevier. (e) Schematic cross-sectional diagrams of MEMS capacitive H2 sensor and sensing mechanism and sample structure for analysis [14]. Copyright 2018, IEEJ. (f) Schematic of the SnO2-D MEMS sensor formation process and production principle [21]. Copyright 2022, Elsevier.
Sensors 24 08125 g002
Figure 3. (a) Fabrication schematic diagram of porous anodized alumina-loaded WO3 sensing layer based on micro-hot plate [22]. Copyright 2013, Elsevier. (b) Sliding transfer method, rolling transfer method, and heat transfer method schematic diagram of ZnO nanowire MEMS H2 sensor [23]. Copyright 2011, Royal Society of Chemistry. (c) The response of Pd/ZnO hydrogen sensors with different bending angles varying with H2 concentrations [24]. Copyright 2013, Elsevier. (d) Schematic atomic structure of adsorption of H2 by pure TiO2 and Pt/TiO2 and separation of H on pure TiO2 and Pt/TiO2 [27]. Copyright 2024, Elsevier. (e) Pd grid contact diagram deposited on SnO2 nanocrystalline film [28]. Copyright 2016, Springer Nature. (f) Schematic view of H2 sensing mechanism for WO3-C/In2O3 sensor [29]. Copyright 2024, Elsevier.
Figure 3. (a) Fabrication schematic diagram of porous anodized alumina-loaded WO3 sensing layer based on micro-hot plate [22]. Copyright 2013, Elsevier. (b) Sliding transfer method, rolling transfer method, and heat transfer method schematic diagram of ZnO nanowire MEMS H2 sensor [23]. Copyright 2011, Royal Society of Chemistry. (c) The response of Pd/ZnO hydrogen sensors with different bending angles varying with H2 concentrations [24]. Copyright 2013, Elsevier. (d) Schematic atomic structure of adsorption of H2 by pure TiO2 and Pt/TiO2 and separation of H on pure TiO2 and Pt/TiO2 [27]. Copyright 2024, Elsevier. (e) Pd grid contact diagram deposited on SnO2 nanocrystalline film [28]. Copyright 2016, Springer Nature. (f) Schematic view of H2 sensing mechanism for WO3-C/In2O3 sensor [29]. Copyright 2024, Elsevier.
Sensors 24 08125 g003
Figure 4. (a) Schematic diagram of the 5 lithographic steps of the process flow diagram of the 3C-SiC cantilever beam: (a1) patterning alignment crosses, (a2) defining cantilevers geometries, (a3) introduction of an isolation layer to separate 3C–SiC and metal contacts, (a4) metal deposition for electromagnetic actuation and inductive detection and (a5) cantilever releasing from the rear side, by etching the Si substrate with KOH [20]. Copyright 2024, Materials Science in Semiconductor Processing. (b) Energy bending diagram and sensing mechanism diagram of the flexible Pd/SiNM H2 sensor [7]. Copyright 2018, American Chemical Society. (c) Schematic of a SiNW FET with a bottom-gate structure for H2 sensing [19]. Copyright 2015, Solid-State Electronics.
Figure 4. (a) Schematic diagram of the 5 lithographic steps of the process flow diagram of the 3C-SiC cantilever beam: (a1) patterning alignment crosses, (a2) defining cantilevers geometries, (a3) introduction of an isolation layer to separate 3C–SiC and metal contacts, (a4) metal deposition for electromagnetic actuation and inductive detection and (a5) cantilever releasing from the rear side, by etching the Si substrate with KOH [20]. Copyright 2024, Materials Science in Semiconductor Processing. (b) Energy bending diagram and sensing mechanism diagram of the flexible Pd/SiNM H2 sensor [7]. Copyright 2018, American Chemical Society. (c) Schematic of a SiNW FET with a bottom-gate structure for H2 sensing [19]. Copyright 2015, Solid-State Electronics.
Sensors 24 08125 g004
Figure 5. (a) Signal diagram of S2 sensor after heat treatment at 900 °C in a dry O2/N2 atmosphere at 500 °C, exposed to CO and ethanol [33]. Copyright 2006, Elsevier. (b) Diagram of the change in response value of 0 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (c) Diagram of the change in response value of 5 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (d) Diagram of the change in response value of 7.5 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (e) A response diagram of the AZO nanoparticles as a function of CO concentration at the temperature of 300 °C [37]. Copyright 2014, Elsevier. (f) Double trimerization synthesis of cryptophane-A [40]. Copyright 2009, Elsevier. (g) Schematic diagram showing the process of creating FETs using thin films of PDPP4T-T-Pd(II) and PDPP4T-T-Hg(II) for the purpose of testing CO and H2S, respectively [43]. Copyright 2019, American Chemical Society.
Figure 5. (a) Signal diagram of S2 sensor after heat treatment at 900 °C in a dry O2/N2 atmosphere at 500 °C, exposed to CO and ethanol [33]. Copyright 2006, Elsevier. (b) Diagram of the change in response value of 0 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (c) Diagram of the change in response value of 5 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (d) Diagram of the change in response value of 7.5 wt% Al-doped TiO2 sensor with CO concentration at 700 °C, 800 °C, and 900 °C [35]. Copyright 2007, Elsevier. (e) A response diagram of the AZO nanoparticles as a function of CO concentration at the temperature of 300 °C [37]. Copyright 2014, Elsevier. (f) Double trimerization synthesis of cryptophane-A [40]. Copyright 2009, Elsevier. (g) Schematic diagram showing the process of creating FETs using thin films of PDPP4T-T-Pd(II) and PDPP4T-T-Hg(II) for the purpose of testing CO and H2S, respectively [43]. Copyright 2019, American Chemical Society.
Sensors 24 08125 g005
Figure 6. (a) Schematic diagram depicting the sensing mechanism of n-type ZnO-450 for NO2 detection [45]. Copyright 2024, Elsevier. (b) Schematic of the gas-sensing process of the MEMS NO2 sensor based on ZnO1−x coating [47]. Copyright 2017, Journal of the Taiwan Institute of Chemical Engineers. (c) Schematic of pulsed laser deposition system used for deposition of PZT thin film [48]. Copyright 2018, Elsevier. (d) The initial growth of CuO-NWs is depicted in the schematic diagram, showing the occurrence of coalescence and competitive growth [49]. Copyright 2022, Elsevier. (e) The mechanism of Co3O4 NO2 gas sensing, taking into account the impact of temperature, nanoscale properties, and the presence of Au-NPs [50]. Copyright 2021, Elsevier.
Figure 6. (a) Schematic diagram depicting the sensing mechanism of n-type ZnO-450 for NO2 detection [45]. Copyright 2024, Elsevier. (b) Schematic of the gas-sensing process of the MEMS NO2 sensor based on ZnO1−x coating [47]. Copyright 2017, Journal of the Taiwan Institute of Chemical Engineers. (c) Schematic of pulsed laser deposition system used for deposition of PZT thin film [48]. Copyright 2018, Elsevier. (d) The initial growth of CuO-NWs is depicted in the schematic diagram, showing the occurrence of coalescence and competitive growth [49]. Copyright 2022, Elsevier. (e) The mechanism of Co3O4 NO2 gas sensing, taking into account the impact of temperature, nanoscale properties, and the presence of Au-NPs [50]. Copyright 2021, Elsevier.
Sensors 24 08125 g006
Figure 7. (a) Au/SnO2 sensor mechanism schematic diagram [51]. Copyright 2018, Elsevier. (b) Schematic model for the Au/SnO2/NiO sensor exposed in air and NO2 [52]. Copyright 2019, Elsevier. (c) Gas-sensing mechanism of TeO2/ZnO NO2 sensor [56]. Copyright 2024, Elsevier.
Figure 7. (a) Au/SnO2 sensor mechanism schematic diagram [51]. Copyright 2018, Elsevier. (b) Schematic model for the Au/SnO2/NiO sensor exposed in air and NO2 [52]. Copyright 2019, Elsevier. (c) Gas-sensing mechanism of TeO2/ZnO NO2 sensor [56]. Copyright 2024, Elsevier.
Sensors 24 08125 g007
Figure 8. (a) The schematic illustration of the MEMS NO2 sensor based on WO3 NTS-MWCNT-RGO [61]. Copyright 2016, Sensors and Actuators B: Chemical. (b) Schematic representation of the synthesis process for SnO2/rGO nanocomposites using various tin salt precursors [63]. Copyright 2015, ACS Analytical Chemistry. (c) Schematic depiction of the NO2 sensing mechanism on the active surface area of MoS2/graphene [64]. Copyright 2020, Optics and Laser Technology. (d) Gas-sensing mechanism of ZIF-8 nanostructures [65]. Copyright 2021, Materials Research Bulletin.
Figure 8. (a) The schematic illustration of the MEMS NO2 sensor based on WO3 NTS-MWCNT-RGO [61]. Copyright 2016, Sensors and Actuators B: Chemical. (b) Schematic representation of the synthesis process for SnO2/rGO nanocomposites using various tin salt precursors [63]. Copyright 2015, ACS Analytical Chemistry. (c) Schematic depiction of the NO2 sensing mechanism on the active surface area of MoS2/graphene [64]. Copyright 2020, Optics and Laser Technology. (d) Gas-sensing mechanism of ZIF-8 nanostructures [65]. Copyright 2021, Materials Research Bulletin.
Sensors 24 08125 g008
Figure 9. (a) Schematic diagram illustrating the sensing mechanism and the energy band structures of BFO-P0 and BFO-P4, (a1a4) Schematic diagram illustrating of the sensing mechanism and (a5,a6) the energy band structures of BFO-P0 and BFO-P4. [68]. Copyright 2022, Elsevier. (b) Gas-sensing mechanism of ZnO sensor with H2S gas [69]. Copyright 2022, Elsevier. (c) Reaction model for Ce0.97Ni0.03O1.97 and schematic illustration for gas-sensing mechanism of CeO2 and Ce0.97Ni0.03O1.97 [72]. Copyright 2022, Elsevier. (d) The schematic diagram for the fabrication of a c/h-In2O3 MEMS gas sensor with monolayer particle film [75]. Copyright 2024, Elsevier.
Figure 9. (a) Schematic diagram illustrating the sensing mechanism and the energy band structures of BFO-P0 and BFO-P4, (a1a4) Schematic diagram illustrating of the sensing mechanism and (a5,a6) the energy band structures of BFO-P0 and BFO-P4. [68]. Copyright 2022, Elsevier. (b) Gas-sensing mechanism of ZnO sensor with H2S gas [69]. Copyright 2022, Elsevier. (c) Reaction model for Ce0.97Ni0.03O1.97 and schematic illustration for gas-sensing mechanism of CeO2 and Ce0.97Ni0.03O1.97 [72]. Copyright 2022, Elsevier. (d) The schematic diagram for the fabrication of a c/h-In2O3 MEMS gas sensor with monolayer particle film [75]. Copyright 2024, Elsevier.
Sensors 24 08125 g009
Figure 10. (a) The diagram illustrating the synthesis process of α-Fe2O3/SnO2 core–shell NWs on MEMS and the diagram depicting the gas-sensing measurement system [76]. Copyright 2024, Elsevier. (b) Comparable cross-sectional perspectives of H2S gas sensor utilizing WO3 film and PPy/WO3 nanocomposite film with a heterostructure (Au/p-PPy/n-WO3/p-PPy/Au) [83]. Copyright 2024, Elsevier. (c) Preparation procedure for the modified In2O3-NPs H2S sensors (MS) [86]. Copyright 2021, American Chemical Society. (d) Schematic representation of the gas-sensing mechanism and band diagram illustration of the Cu-doped ZnO/RGO nanocomposite sensor before and after exposure to ambient air and H2S gas, showing the heterojunction [87]. Copyright 2020, Elsevier.
Figure 10. (a) The diagram illustrating the synthesis process of α-Fe2O3/SnO2 core–shell NWs on MEMS and the diagram depicting the gas-sensing measurement system [76]. Copyright 2024, Elsevier. (b) Comparable cross-sectional perspectives of H2S gas sensor utilizing WO3 film and PPy/WO3 nanocomposite film with a heterostructure (Au/p-PPy/n-WO3/p-PPy/Au) [83]. Copyright 2024, Elsevier. (c) Preparation procedure for the modified In2O3-NPs H2S sensors (MS) [86]. Copyright 2021, American Chemical Society. (d) Schematic representation of the gas-sensing mechanism and band diagram illustration of the Cu-doped ZnO/RGO nanocomposite sensor before and after exposure to ambient air and H2S gas, showing the heterojunction [87]. Copyright 2020, Elsevier.
Sensors 24 08125 g010
Figure 11. (a) Schematic operation of the WO3 thin-film MEMS NH3 sensor at temperatures exceeding 200 °C [91]. Copyright 2023, Elsevier. (b) Schematic diagram of an ammonia measurement system featuring a ZnO-nanorod MEMS sensor [96]. Copyright 2017, MDPI AG. (c) Schematic diagram of the reaction between NH3 and H2O molecules [98]. Copyright 2015, Elsevier. (d) Proposed mechanism for detecting NH3 using N-doped DLC film [102]. Copyright 2021, Elsevier. (e) Process diagram of preparing ZnS sensing layer on SAW device [108]. Copyright 2020, Elsevier.
Figure 11. (a) Schematic operation of the WO3 thin-film MEMS NH3 sensor at temperatures exceeding 200 °C [91]. Copyright 2023, Elsevier. (b) Schematic diagram of an ammonia measurement system featuring a ZnO-nanorod MEMS sensor [96]. Copyright 2017, MDPI AG. (c) Schematic diagram of the reaction between NH3 and H2O molecules [98]. Copyright 2015, Elsevier. (d) Proposed mechanism for detecting NH3 using N-doped DLC film [102]. Copyright 2021, Elsevier. (e) Process diagram of preparing ZnS sensing layer on SAW device [108]. Copyright 2020, Elsevier.
Sensors 24 08125 g011
Table 3. Performance comparison of nitrogen dioxide MEMS gas sensors and sensitive materials.
Table 3. Performance comparison of nitrogen dioxide MEMS gas sensors and sensitive materials.
Sensitive Material CategorySensitive MaterialFabrication TechniqueWorking Temperature (°C)Detection Range (ppm)Response Time (s)Recovery Time (s)SensitivityRef.
MOSZnO-450Pyrolysis of ZIF-901900.035–409262.42@10 ppm c[45]
MOSZnORF magnetron sputteringRoom temperature0.4–16N/aN/a6.0 kHz/ppm e[46]
MOSZnO1-xSuspension flame spray (SFS)Room temperature0.25–11803002.61@1 ppm c[47]
MOSPZTPulsed laser depositionRoom temperature80–250N/aN/a9.6 Hz/ppm e[48]
MOSCuO-NWsRF sputtering119100–500136.3272.350.1%@0.5 ppb c[49]
Metal/MOSAu/Co3O4-NPsUltrasonic wave grinding technology1360.01–10846851%@0.2 ppm c[50]
Metal/MOSAu-NP/SnO2 thin filmDC/RF sputteringRoom temperature0.6–5070N/a90%@50 ppm d[51]
Metal/MOSAu-NP/SnO2/NiO thin filmsE-beam evaporation2000.05–5N/aN/a180%@5 ppm a[52]
Metal/MOSTa/In2O3RF magnetron co-sputtering1201–1005933964.5@100 ppm c[53]
Metal/MOSTa/In2O3RF magnetron sputtering1100.7–1004832976.1@100 ppm c[54]
MOS/MOSRhOx/B-In2O3Dip coating25–1251–2081342@5 ppm c[55]
MOS/MOSTeO2/ZnOCo-sputtering technique1000.2–1133880%@1 ppm d[56]
PolymerPPy thin filmSpin coating techniqueRoom temperature10–10012621701.12@100 ppm a[57]
CBMCNTRF sputteringRoom temperature100.0010.0011.77@10 ppm a[58]
CBMCNTRF sputteringRoom temperature10–50N/aN/a52.20%@50 ppm c[59]
Polymer/CBMPolythiophene-SWNTsPotent ion static
deposition
Room temperature0.01–1020N/a28@10 ppm b[60]
MOS/CBMWO3-NP/MWCNT-RGOPhotolithography and radio frequency magnetron sputteringRoom temperature1–2542080017%@5 ppm c[61]
MOS/CBMSWNT-Fe2O3Floating catalytic chemical vapor deposition methodRoom temperature1–100N/aN/a18.3%@100 ppm c[62]
MOS/CBMSnO2/rGOHummers method5514–1108038011.8@110 ppm a[63]
Inorganic substance/CBMMoS2/GrapheneChemical methodRoom temperature0–500223561%@500 ppm c[64]
MOFPolyhedral ZIF-8 nanostructuresSolvothermal method35010–100113.5111.5118.5@100 ppm a[65]
a R = Ra/Rg. b R = Rg/Ra. c R = |Rg − Ra|/Ra × 100%. d R = |Rg − Ra|/Rg × 100%. e Δf = 2f02 Δm/A √ρquq.
Table 4. Performance comparison of hydrogen sulfide MEMS gas sensors and sensitive materials.
Table 4. Performance comparison of hydrogen sulfide MEMS gas sensors and sensitive materials.
Sensitive Material CategorySensitive MaterialFabrication TechniqueWorking Temperature (°C)Detection Range (ppm)Response Time (s)Recovery Time (s)SensitivityRef.
MOSTiO2Sputtering
evaporation
706–38N/aN/a144@38 ppm c[67]
MOSBiFeO3Facile sol–gel method2200.01–1.2374.8@1.2 ppm a[68]
Metal/MOSAZORF sputtering
technique
2500.2–1.0N/aN/a14%@1000 ppb d[69]
Metal/MOSAg/ZnOCo-sputtering technique2500.2–1.03N/a16%@0.001 ppm c[70]
Metal/MOSFe-NiOx nanotubesDrop coating2700.05–0.83.28.15.24@0.8 ppm c[71]
Metal/MOSNi/CeODrop coating2000.01–5.68133.06@0.5 ppm a[72]
Metal/MOSIn2O3/Co32Drop coating1850.05–2.53076N/a[73]
MOS/MOSZnO/SnO2LPD25–3001–2036–55N/a35.31@20 ppm a[74]
MOS/MOSc/h-In2O3Self-assembly method1600.02–503.3N/a54.4@50 ppm a[75]
MOS/MOSα-Fe2O3/SnO2Atomic layer deposition (ALD)2501–1013.8104.54.3@10 ppm a[76]
MOS/MOSNb2O5/SnO2ALD2751–2020974.0@20 ppm a[77]
MOS/MOSCuO/SnO2Dipping method350.01–1090N/a56,000@10 ppm a[78]
MOS/MOSCuO/TiO2Simple electrochemical anodizationRoom temperature3–400419246.81%@100 ppm c[79]
MOS/MOSCu2O/CuOOne-step reduction approach950.05–1N/a762.1@0.05 ppm b[80]
MOS/MOSCuO/WO3RF sputtering
technique
3000–15517 min534@10 ppm a[81]
MOS/MOSWO3-Bi2WO6Facile hydrothermal
technique
Room temperature0.002–0.050521194.4@0.050 ppm a[82]
Polymer/MOSPPy/WO3In situ photopolymerizationRoom temperature0.1–136012,60081%@1 ppm c[83]
Polymer/MOSPPy/WO3Chemical oxidation polymerization and mechanical mixing900.2–1703461%@1000 ppm c[84]
Polymer/MOSPDPP4T-T-Hg(II)Air–water interface coordination reactions of thymine groups with ionsRoom temperature0.001–1000N/aN/aN/a[44]
CBM/polymer/MOSGt/Ps/In2O3Doctor blade methodRoom temperature0.1–1N/aN/a70@1 ppm a[85]
CBM/polymer/MOSGt/Ps/CuAc/In2O3Doctor blade methodRoom temperature0.1–360N/a18.1@0.1 ppm a[86]
Metal/MOS/CBMCu/ZnO/RGORF magnetron sputtering240.136–25014320.87%@100 ppm c[87]
a R = Ra/Rg. b R = Rg/Ra. c R = |Rg − Ra|/Ra × 100%. d R = |Rg − Ra|/Rg × 100%.
Table 5. Performance comparison of NH3 MEMS gas sensors and sensitive materials.
Table 5. Performance comparison of NH3 MEMS gas sensors and sensitive materials.
Sensitive Material CategorySensitive MaterialFabrication TechniqueWorking Temperature (°C)Detection Range (ppm)Response Time (s)Recovery Time (s)SensitivityRef.
MOSTiO2Hydrothermal process35056, 103, 156N/aN/a1@1000 ppm a[88]
MOSTiO2Hydrothermal methodRoom temperature1<1 s<1 s164.2@1 ppm b[89]
MOSWO3Ultrasonic wave grinding2000.04–5301352.525@5 ppm a[90]
MOSWO3-NPsPlasma-enhanced chemical vapor deposition (PECVD)1421.3594716%@1.3 ppm c[91]
MOSBismuth tungstate (Bi2WO6) nanomaterialsHydrothermal techniqueRoom temperature0–500N/aN/a5 counts/kpa[92]
MOSZnO-NWsDrop-coating techniqueRoom temperature10004–5N/a−956 Hz/ppm d[93]
MOSZnO nanorodsWet chemical routeRoom temperature150N/aN/a0.62 Hz/ppm d[90]
MOSZnO-NRsHydrothermal methodRoom temperature800720140011.33@100 ppm a[94]
MOSZnO-NRsHydrothermal methodRoom temperature100151568−1.094 Hz/ppm d[95]
MOSZnO nanofilmSol–gel and spin coatingRoom temperature100143426−0.307 Hz/ppm d[96]
MOSThree-dimensional ZnO nanoflowersDrop coatingRoom temperature0–728126364.32 counts ppm−1[97]
MOS/MOSZnO/SiO2(ZcS) composite filmSol–gel method
and spin coating
Room temperature50N/aN/a0.02264 kHz/ppm d[98]
CBMSWNT filmCVDRoom temperature10–2002503004%@200 ppm c[62]
CBMPolyaniline/grapheneDrop-coating methodRoom temperature20–100503585 Hz/ppm d[99]
CBMPolyaniline/GO layerDrop-coating
method
Room temperature800793214 Hz/ppm d[100]
CBMCA/PEI/GO nanofiberSol–gel methodRoom temperature80<10N/a11.3 Hz/ppm d[101]
CBMN-DLCECR-PECVDRoom temperature0.1–1005293.3 kHz/ppm d[102]
PolymerPPy thin filmPolymerization processRoom temperature4–802080016%@25 ppm c[103]
PolymerPPy nanofiberReactive template approachRoom temperature20–15015N/a1.53@20 ppm b[104]
PolymerIodine-doped polythiophene (PTh) film (IPTF)Electrophoretic
deposition
technique
Room temperature460–18507834694.64%@1850ppm c[105]
Metal/polymerAu/PPySpin coatingRoom temperature2–105972898 Hz/ppm d[106]
Inorganic substanceL-glutamic acid hydrochlorideAir-brushed coatingRoom temperature0.56–4.0<300<18074 Hz/ppm d[107]
Inorganic substanceZnS nanostructuresChemical bath deposition methodRoom temperature204514862.5 Hz/ppm d[108]
Inorganic substanceZinc vanadate (Zn3(VO4)2) nanopowderDip coatingRoom temperature20–500280835400.019 µV/ppm[109]
a R = Ra/Rg. b R = Rg/Ra. c R = |Rg − Ra|/Ra × 100%. d Δf = 2f02 Δm/A √ρqμq.
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

Wu, Y.; Lei, M.; Xia, X. Research Progress of MEMS Gas Sensors: A Comprehensive Review of Sensing Materials. Sensors 2024, 24, 8125. https://doi.org/10.3390/s24248125

AMA Style

Wu Y, Lei M, Xia X. Research Progress of MEMS Gas Sensors: A Comprehensive Review of Sensing Materials. Sensors. 2024; 24(24):8125. https://doi.org/10.3390/s24248125

Chicago/Turabian Style

Wu, Yingjun, Ming Lei, and Xiaohong Xia. 2024. "Research Progress of MEMS Gas Sensors: A Comprehensive Review of Sensing Materials" Sensors 24, no. 24: 8125. https://doi.org/10.3390/s24248125

APA Style

Wu, Y., Lei, M., & Xia, X. (2024). Research Progress of MEMS Gas Sensors: A Comprehensive Review of Sensing Materials. Sensors, 24(24), 8125. https://doi.org/10.3390/s24248125

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