Next Article in Journal
A Low-Power Opamp-Less Second-Order Delta-Sigma Modulator for Bioelectrical Signals in 0.18 µm CMOS
Next Article in Special Issue
Flame-Made La2O3-Based Nanocomposite CO2 Sensors as Perspective Part of GHG Monitoring System
Previous Article in Journal
Comprehensive mPoint: A Method for 3D Point Cloud Generation of Human Bodies Utilizing FMCW MIMO mm-Wave Radar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of Ag Addition on the Gas-Sensing Properties of Nanostructured Resistive-Based Gas Sensors: An Overview

1
Division of Materials Science and Engineering, Hanyang University, Seoul 04763, Korea
2
The Research Institute of Industrial Science, Hanyang University, Seoul 04763, Korea
3
Department of Materials Science and Engineering, Inha University, Incheon 22212, Korea
4
Department of Materials Science and Engineering, Faculty of Engineering, Urmia University, Urmia 5756-151818, Iran
5
Department of Materials Science and Engineering, Shiraz University of Technology, Shiraz 71557-13876, Iran
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sensors 2021, 21(19), 6454; https://doi.org/10.3390/s21196454
Submission received: 3 September 2021 / Revised: 22 September 2021 / Accepted: 23 September 2021 / Published: 27 September 2021
(This article belongs to the Special Issue Gas Sensors Based on Semiconductor Materials)

Abstract

:
Nanostructured semiconducting metal oxides (SMOs) are among the most popular sensing materials for integration into resistive-type gas sensors owing to their low costs and high sensing performances. SMOs can be decorated or doped with noble metals to further enhance their gas sensing properties. Ag is one of the cheapest noble metals, and it is extensively used in the decoration or doping of SMOs to boost the overall gas-sensing performances of SMOs. In this review, we discussed the impact of Ag addition on the gas-sensing properties of nanostructured resistive-based gas sensors. Ag-decorated or -doped SMOs often exhibit better responsivities/selectivities at low sensing temperatures and shorter response times than those of their pristine counterparts. Herein, the focus was on the detection mechanism of SMO-based gas sensors in the presence of Ag. This review can provide insights for research on SMO-based gas sensors.

1. Semiconducting Metal Oxide (SMO)-Based Gas Sensors

Toxic gases pose a major threat to modern life. Therefore, the early detection of toxic gases using reliable electronic gas sensors is extremely important. In this respect, chemiresistive gas sensors are potentially attractive due to their easy manufacturing, simple operating principle, and low cost [1,2,3,4,5,6,7,8,9,10]. In the area of resistive-based gas sensors, different semiconducting materials can be used for the realization of gas sensors. For example, the organic semiconducting materials such as organic π-conjugated materials have advantages of tailorable chemical structures, solution processability, and mechanical flexibility; thus, they are potential candidates for applications in low-cost, low-temperature, and portable gas sensors. Nonetheless, they have some shortages such as relatively poor sensitivity, slow response, and low recovery [1,2,3]. The functionalization of these materials with noble metals such as Ag through chemical methods can increase the response and selectivity to a particular gas; however, as far as we know, there is no or few studies in this regard. On the other hand, semiconducting metal oxides (SMO)-based gas sensors have advantages such as high response, high stability, relatively fast dynamics, simple fabrication, and low costs. So, they are very popular for the detection of various gases [4,5,6,7,8,9,10,11,12,13,14]. However, they work at high temperatures and show poor selectivity.
Typically, in the laboratory, gas-sensing measurements are dynamically conducted in an enclosed test chamber of a defined volume equipped with an inlet and outlet for gas flow [15]. Generally, ideal gas sensors must be inexpensive, operate at low temperatures or room temperature, detect gases at low levels, and be highly stable, sensitive, selective, and fast [16]. Nano-based gas-sensing materials possess high surface areas and unique electrical properties and consequently are preferred over their micron-sized counterparts for the development of gas sensors [17]. The detection mechanism of SMOs significantly depends on the modulation of electrical resistance due to the interaction of the sensing layer with the target gases [18]. The general sensing mechanism of resistive-based gas sensors is schematically shown in Figure 1 for n-type and p-type SMOs, in the presence of oxidizing and reducing gases. When an SMO is exposed to air, depending on the sensing temperature, the adsorption of oxygen on the sensor surface causes the ionization of oxygen molecules in the form of molecular or atomic ions. As a result of oxygen adsorption, a so-called electron depletion layer with a low concentration of electrons forms on the surfaces of n-type SMOs, which has a higher resistance than the core region of the metal oxide. Alternatively, a so-called hole accumulation layer forms on the surfaces of p-type SMOs, which has a lower resistance than the core regions of the metal oxides due to an increase in the number of holes as major carriers (Figure 1a). Upon exposure of the gas sensor to reducing gases, the gases will be adsorbed on the surface of the sensing layer and react with the already adsorbed oxygen ions, liberating the electrons back to the surface of the gas sensor. Thus, the width of the electron depletion layer in n-type metal oxides decreases, leading to a decrease in the resistance of the sensor. For p-type metal oxides, the width of the hole accumulation layer decreases because of the combination of the released electrons with holes, resulting in an increase in sensor resistance, which contributes to the sensor signal (Figure 1b). For oxidizing gases due to the further abstraction of electrons, the widths of both the electron depletion layer and hole accumulation layer increase, leading to an increase in resistance in n-type metal oxides and decrease in the resistance in p-type metal oxides (Figure 1c) [19,20].
Since the adsorption of gas molecules on the surface of the gas sensor depends on the surface area, different morphologies have been investigated for gas-sensing applications. In fact, the larger surface area results in more availability of the adsorption sites on the surface of the gas sensor, resulting in a higher response of the gas sensor [21]. In addition to the increase in the surface area, several strategies, such as doping [22], decoration of noble metals [23], and formation of heterojunctions [24] have been reported to enhance the gas-sensing performances of SMOs. Selectivity is a challenging issue of resistive-based gas sensors and is not completely solved for gas sensors. However, there are some strategies to increase the selectivity of the gas sensor to a particular gas. For example, the control of sensing temperature, use of membranes, use of catalytic materials, and noble metal decoration have been suggested to increase the selectivity of resistive-based gas sensors [25].

2. Noble Metal Decoration

In the area of SMOs-based gas sensors, there are many review papers [26,27,28,29,30,31,32,33,34,35]. However, less attention has been paid to review the effect of noble metals on the gas-sensing performance of SMO-based gas sensors. Mirzaei et al. [36], Korotcenkov et al. [37], and Luo et al. [38] have discussed the effect of Pd, Au, and metal doping, respectively, on the gas response of SMO-based gas sensors. Mirzaei et al. [23] briefly discussed the effect noble metals on the gas response of SMO-based gas sensors. So, as far as we know, there is no review paper devoted solely to the effect of the addition of Ag in the form of decoration and doping on the gas response characteristics of SMO-based gas sensors. Thus, in this review paper, we have tried to cover this area of SMO-based gas sensors. It should be noted that in this manuscript, both decoration and loading are interchangeable words, and both of them mean the dispersion of Ag NPs on the surface of SMO.
Among the various strategies, the decoration of SMOs with noble metals is perhaps the most attractive strategy, and enhanced gas responses of the resulting sensors are mainly attributed to the chemical and electronic sensitization effects of noble metals [23,39]. Typically, noble metals are recognized for their catalytic activities, and they can decrease the adsorption activation energies of gases on the surfaces of SMOs and enhance the sensing activities of SMO-based sensors. The surface doping of noble metals may generate active surface sites that facilitate the preferred adsorption of target gases and increase their concentration. Moreover, noble metals can offer reaction paths that can reduce the activation energy and then improve the reaction rate and selectivity, sensitivity or response, and reliability of the sensor [22,23]. Generally, the chemical sensitization effect originates from a spillover effect, where the noble metal may act as a potential site for the adsorption and dissociation of O2 molecules [40]. Subsequent spillover onto neighboring SMOs results in enhancement of the adsorption rate of O2. Additionally, noble metals may catalyze the dissociation of target gases and boost the rates of reactions between the target gases and the previously chemisorbed O2 species. Basically, the chemical sensitization effect particularly depends on the operating temperature at which the catalytic pathway is preferred and is definite to the case where chemical affinities exist among the species concerned. However, the electronic sensitization effect arises from the direct electronic interaction between the promoter and the semiconductor surface. Typically, the electronic sensitization effect depends on the presence of a junction potential barrier between the semiconductor and the noble metal. Differences between the Fermi levels of the SMO and the catalyst cause the depletion or accumulation of charge carriers at the semiconductor, adjacent to the particles, because of the pinning effect [41,42,43]. When the oxidation state of a noble metal, including Ag, changes with respect to the surrounding atmosphere, the electronic state of the semiconductor accordingly varies because Ag can form Ag2O or AgO in air, whereas Ag2O or AgO is simply reduced to metallic Ag when exposed to a reducing gas. In fact, Ag2O or AgO create an electron-depletion layers within the semiconductor, whereas the electronic interaction is disrupted when Ag2O or AgO is reduced to Ag [41,42,43]. As an example, a schematic of the chemical and electronic sensitization effect of noble metals and SMOs (such as TiO2) with H2 gas as the target gas is shown in Figure 2 [22,42].
In this case, the H2 molecules may dissociate into H atoms on noble metal clusters and then spillover onto the TiO2 surface, which accelerates the reaction [22,39,42]. Herein, the noble metal clusters simply reduce the activation energy of the reaction or increase the reaction rate. The electronic sensitization effect can be attributed to electronic interactions between TiO2 and noble metal clusters. Accordingly, because of the difference between the work functions and electron affinities of TiO2 and noble metal, a depleted space charge region is generated close to the noble metal/TiO2 interface, which causes band bending [23,39]. The electronic sensitization effect is primarily controlled by the spillover effect of H2 on the TiO2 surface and reduces the TiO2 resistance via the transfer of electrons at the noble metal/TiO2 interface. In another case of doping, the dopant atoms enter the TiO2 framework and subsequently distort the TiO2 lattice, thereby producing many oxygen defects that provide higher local electric fields and facilitate the dissociation of H2 [22]. Consequently, the rate of the reaction between H and chemisorbed oxygen ions is enhanced, thereby reducing the response time. Additionally, more oxygen species may adsorb around the doped atoms than those in the case of pristine TiO2, thereby increasing the resistance and sensitivity or response of the sensor [23,39,41,42].
Generally, noble metals are very expensive. If even small amounts of noble metals are used for decorating or doping SMOs, the overall cost of the sensor inevitably increases. Nevertheless, among various noble metals such as Pd, Pt, and Au, Ag has the lowest price. Furthermore, Ag has several interesting properties, for example, high efficiency and high reactivity toward O2 adsorption. Additionally, the Ag ion has been reported as a high mobility cation, which has been utilized in resistive-switching devices [43,44]. Moreover, Ag catalysts are efficient for not only improving the sensitivity of the sensor but also for reducing the working temperature owing to the unique catalytic and electronic characteristics of Ag. As an effective catalyst, Ag can attract O2 molecules in the air, transfer them to the surfaces of SMOs, and subsequently promote the capture of electrons from the conduction band of SMOs by these O2 molecules [41,42,43,44]. Therefore, Ag can be an appropriate sensitizer for introduction into the surfaces of SMO-based resistive gas sensors. Although the addition of Ag may boost the performance of the sensing material, a reasonable doping level/concentration of Ag is needed to realize the maximum sensing response. Typically, the decoration of SMO with Ag at an optimal concentration can result in the highest sensing performance of SMO. In fact, as the catalytic and electronic effects of Ag are unsatisfactory, the best sensing properties cannot be achieved when the loading amount of Ag is extremely low. When the loading or decoration concentration of Ag is substantially high, the electrons are conducted along metallic Ag, thereby decreasing the response of the gas sensor [45]. Generally, the best outcomes are obtained using ultrafine Ag nanoparticles (NPs) at low concentrations (0.1–10 wt.%) and high surface dispersion, promoting the catalytic activity of the sensor without compromising the function of the sensing layer [46]. In the following section, we discuss some of the most recent and important studies on the application of Ag-decorated/loaded semiconductors for the detection of toxic and hazardous gases.

3. Ag-Decorated/Loaded Gas Sensors

3.1. Ag-Decorated/Loaded Acetone (CH3COCH3) Gas Sensors

CH3COCH3 belongs to the family of volatile organic compounds (VOCs) that are employed as solvents in various industries. CH3COCH3 at higher levels may have a negative impact on the central nervous system and may be harmful to the eyes and nose. Therefore, the occupational threshold limit value for CH3COCH3 was set at 250 ppm considering a time-weighted average of 8 h [47]. Accordingly, the development of sensitive CH3COCH3-sensing devices is required from the perspective of safety. In this regard, some attempts have been made to establish selective CH3COCH3 sensors by adding Ag into SMOs. For instance, Xu et al. [48] described the synthesis of Ag-decorated SnO2 hollow nanofibers (NFs) with large surface areas and their CH3COCH3 sensing properties. Typically, noble metals have higher electrical conductivities that enable the rapid transfer of electrons and catalysis of the oxidation of reducing gas molecules such as VOCs [45,46,47,48]. Thus, noble metal-incorporated sensing materials can be suitable candidates for detecting VOCs. The developed Ag-decorated SnO2 sensor showed an excellent response and higher selectivity for CH3COCH3 at 160 °C. A schematic of the sensing interactions between the CH3COCH3 gas molecules and pure and Ag-decorated SnO2 is shown in Figure 3. Owing to the presence of p-type Ag2O crystals, p-n heterojunction interfaces formed between p-type Ag2O crystals and n-type SnO2, generating a broader depletion layer on the SnO2 side, thereby increasing the initial resistance of the sensing material. Upon exposure to CH3COCH3, significant modulation of the heterojunctions occurred and contributed to the sensor response. The Ag NPs produced Ag2O in air, which has facilitated the development of highly electron-depleted layers while removing electrons from SnO2. This process has increased the total resistance and the area adjacent to the Ag–SnO2 interfaces, thus increasing the sensitivity of SnO2 to CH3COCH3. Furthermore, one-dimensional (1D) SnO2 hollow nanostructures fabricated via electrospinning considerably facilitated the diffusion and transport of electrons in CH3COCH3 molecules, thereby rapidly and substantially changing the sensor resistance. Additionally, SnO2 hollow NFs with both outer and inner surfaces exhibited high surface areas and eventually contributed to high CH3COCH3 responses. In summary, the significantly high CH3COCH3 sensing performances of Ag/SnO2 composites are attributed to the Ag dopant and 1D NF structures of SnO2. Ag with high electrical conductivity enabled the rapid transfer of electrons and catalyzed the oxidation of the reducing CH3COCH3 gas molecules. Moreover, the SnO2 NFs with higher surface areas led to an effective dispersion of catalyst particles and possessed a porous tube-like structure that promoted rapid gas flow and a superior ability to store and release oxygen ions.
In another study, Kilic et al. [49] reported the CH3COCH3 sensing characteristics of Ag-loaded TiO2 nanorods (NRs). They used a seed-mediated hydrothermal approach to grow TiO2 NRs and subsequently loaded Ag onto TiO2 NRs by thermally evaporating metallic Ag at different times of 30, 45, and 90 s. The authors demonstrated that the enhancement in the CH3COCH3 detection performance of the Ag-loaded TiO2 (45 s) sensor was due to the catalytic activity of Ag. Actually, Ag loading increased the number of adsorption sites on the surface of TiO2 and accelerated the rate of electron exchange between the TiO2 surface and CH3COCH3 molecules. Under a CH3COCH3 atmosphere, the bonds between CH3COCH3 molecules were easily dissociated by Ag, allowing these molecules to quickly interact with the chemisorbed oxygen species. Moreover, the decoration of Ag NPs onto TiO2 NRs pinned the Fermi level of TiO2 because of the transfer of electrons from TiO2 to Ag. This led to surface band bending and induced a more pronounced electron–hole separation effect, thereby enhancing the sensitivity of the sensor to CH3COCH3. The lower CH3COCH3 response of the Ag-loaded TiO2 (90 s) sensor was associated with the accumulation of Ag clusters on the TiO2 surface, which impeded O2 diffusion within TiO2 and reduced the catalytic activity of Ag.

3.2. Ag-Decorated/Loaded Chlorine (Cl2) Gas Sensors

Cl2 gas is very hazardous to the human respiratory mucous membrane at concentrations in the range 0.2–3.5 ppm. It causes psychological disorders, skin infections, and even liver damage. Hence, the early detection and successive monitoring of hazardous Cl2 by reliable gas sensors are highly important. Few studies have been reported on the Cl2 detection properties of Ag-loaded sensors. Li et al. [50] illustrated the effect of Ag loading on the Cl2 response of bismuth ferrite (BiFeO3, BFO) nanospheres (BFO NSs); they synthesized BFO NSs using a sol–gel route and successively loaded Ag NP onto BFO NSs via a photodeposition technique. Typically, BFO has a distinct surface reactivity, O2 adsorption ability, and a narrow band gap, which are beneficial for improving the gas-sensing characteristics. Therefore, BFO NSs are important candidates for the detection of Cl2. The observed response of the optimum 4 mg AgNO3-modified BFO sensor to 10 ppm Cl2 was 72.62 at a working temperature of 240 °C. This response was 2.5 times that of the pure BFO sensor. A schematic of the energy band structure and hole transfer after the loading of Ag NPs is shown in Figure 4. Upon the loading of Ag NPs onto BFO, the positive charges on BFO reduced, and downward band bending occurred because the holes on the BFO NSs transferred to the Ag NPs. Along with the separation of electrons and holes, the electrons probably combined with O2 and Cl2 on the surface of the gas sensor. No barriers were formed because of the reduction in the number of conduction band electrons, thus supporting the progress of the reaction. Moreover, the Fermi levels of Ag and BFO correspond to one another at the same level, following contact because BFO (Φm) has a lower work function than that of Ag (Φs). Here, Ag NPs might also be used as distinct adsorption sites for O2 and Cl2. Furthermore, at the Ag/BFO interface, a Schottky junction was created that decreased the intergranular barriers and enhanced the interfacial effect [50].
In contrast, the number of electrons at the reaction sites increased, and the reaction quickly saturated because of the rapid transfer of charge carriers from BFO to Ag. Consequently, the gas response significantly improved, and the response time shortened. However, the amount of adsorbed O ions decreased when high amounts of Ag were loaded onto the BFO NSs, which reduced the reactions of O2 and Cl2, thereby decreasing the sensor response. Thus, this study demonstrates that the addition of noble metals presents a new way of enhancing the sensitivity of p-type semiconductors toward Cl2 gas.

3.3. Ag-Decorated/Loaded Acetylene (C2H2) Gas Sensors

C2H2 is a flammable and colorless gas with a peculiar odor and is commonly utilized as a fuel in oxyacetylene welding and metal cutting and as a raw material in various industries. C2H2 poses serious threats due to its inherent instability, primarily in the cases of liquefaction, pressurization, heating, or mixing with air. This means that C2H2 can cause massive explosions if it leaks. Accordingly, from environmental and safety perspectives, the development of highly efficient C2H2 sensors is becoming increasingly important [31,51,52]. In this regard, Uddin et al. [51] described the C2H2 sensing properties of chemically synthesized 0–5 wt.% Ag-loaded ZnO-reduced graphene oxide (ZG-Ag) ternary hybrids. Generally, herein, the use of rGO was preferred over that of graphene because of difficulties in the large-scale production of graphene and lack of functional groups and band gaps in graphene [53]. Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images (Figure 5) support the Ag/ZnO/Gr configuration of these heterostructures.
The Ag-loaded sensor based on ternary hybrids showed a higher response than that of the sensor without Ag. The sensor without Ag could detect 16–100 ppm C2H2 at 250 °C. The response of the 3 wt.% Ag-loaded ZnO–Gr (ZG-Ag3) sensor was 22 at 150 °C, indicating that 3 wt.% was the optimum loading amount of Ag. The loading of 3 wt.% Ag onto the ZnO–Gr sensor not only enhanced the sensor response but also reduced the optimum sensing temperature. Ag has a superior capacity to dissociate O2 than ZnO, and it catalyzes the dissociation of molecular O2 on the sensor surface, which generates higher C2H2-sensing active sites and thereby a higher sensor response. The development of depletion layers around the ZnO NPs due to the presence of Ag was mostly correlated with the modulation of nano-Schottky barriers, thereby improving the low-temperature surface reactivity. Upon the loading of Ag onto the ZnO–Gr sensor, further dynamic sites generated at the sensing layer/Ag interfaces because of the spillover effect, and a larger number of C2H2 molecules adsorbed on the sensor surface. Therefore, a high sensor response was obtained when the loading amount of Ag was optimal. Nevertheless, when an excess amount of Ag was loaded, the O2 molecules dissociated on the surface and surpassed the percolation threshold, which caused an overlapping spillover zone, ultimately affecting the efficient transfer of O2 and decreasing the chances of C2H2 adsorption; this reduced the sensor response.
In another study associated with the detection of C2H2, Uddin et al. [54] designed a C2H2 sensor using Ag-decorated ZnO NRs supported by a flexible polyimide (PI)/polytetrafluoroethylene (PTFE) substrate (Figure 6).
Flexible gas sensors should have the following characteristics: (i) the ability to detect gases at low concentrations and (ii) low- or room-temperature operation [55]. In this study, radio frequency magnetron sputtering (125 W, 7 mTorr) was applied to load Ag NPs onto the sensing layer of ZnO NRs in an Ar environment at different loading times of 6, 8, and 10 s. Gas-sensing analysis demonstrated that the flexible sensor developed using 8 s Ag-loaded ZnO NRs showed a higher response of 13.8 toward 100 ppm C2H2 at 200 °C than those of the pristine (3.21), 6 s (9.33), and 10 s (9.57) Ag-loaded sensors. Furthermore, upon repeated bending and relaxation for up to 5 × 104 cycles, the sensor response only slightly decreased, which demonstrated the outstanding sustainability and mechanical strength of the designed flexible sensor. This performance improvement was attributed to the integration of the Ag–ZnO structure with the flexible PI substrate, which improved the physical binding and adherence of the Ag–ZnO structure to the supporting substrate. ZnO NRs with small diameters ensured an improved surface area and a larger aspect ratio of the resulting sensor. This caused inadequate surface atomic coordination and higher surface energy, leading to better adsorption of O ions and hence improved sensor response. Furthermore, because of catalytic interactions, the Ag NP/ZnO NR interface might produce further charge carriers or O vacancies on the sensor surface. In the case of Ag-loaded ZnO NRs, the ZnO surface transformed from an electron-depletion state to an almost flat band state along with a redox change at Ag. Additionally, the improved C2H2 sensing activity observed at an optimum temperature of 200 °C was ascribed to the height of the Schottky barrier between the ZnO grains and Ag in the Ag-loaded ZnO NRs. The improvement in the performance of the Ag-loaded ZnO NRs may also be controlled by changing the Ag concentration for the sensitive variation of the Ag oxidation state during exposure to C2H2. When the Ag loading amount was high, a thick Ag layer developed on the ZnO NR surface, which decreased the efficacy of the open surface porosity of ZnO, thereby reducing the gas sensor response. Herein, the observed higher selectivity of the Ag/ZnO NR sensor toward C2H2 was attributed to the chemical sensitization effect. Since Ag has excellent catalytic activity, it acts as a certain adsorption site for the dissociation of O2 and separation of H2 molecules owing to the spillover effect.

3.4. Ag-Decorated/Loaded Triethylamine (TEA) Gas Sensors

TEA, a clear and flammable liquid with a strong odor of ammonia, is widely used in chemical industries [56]. However, TEA can cause many severe health issues, such as pulmonary edema, gastroenteritis, headaches, and even death, because of its toxicity [57]. Additionally, dead fish and other decaying marine products may release TEA, and the concentration of TEA is expected to progressively increase over time [58]. Consequently, TEA may serve as a chemical indicator to assess the quality of marine food [59]. Therefore, designing an advanced sensor for the detection and monitoring of TEA is required. In contrast, recently, three-dimensional (3D) structural nanomaterials have received considerable attention in gas-sensing applications. In fact, 3D structures may offer highly efficient specific surface areas that improve the response of the corresponding gas sensor. Accordingly, Shen et al. [60] reported the TEA sensing characteristics of Ag-loaded 3D porous ZnO microspheres. Their morphological studies indicated that the ZnO microspheres (3–5 μm in size) were assembled by numerous thin and porous nanosheets with sizes of ≈20 nm. The developed sensor exhibited better cross-selectivity toward TEA, which was significantly attributed to the differences between the reactivity of the target gases caused by their different bond energies and chemical molecular structures. The reported C-N, C-O, C-H, C=C, C=O, and O-H bond energies are 307, 326, 414, 610.3, 798.9, and 458.8 kJ/mol, respectively. As a result of the lower C-N bond energy, the TEA molecules were more easily reduced by Ag–ZnO.
In this case, the baseline resistance of the Ag-loaded sensor (2737.8 MΩ) was higher than that of the pristine sensor (28.5 MΩ). Furthermore, O species may more readily adsorb on the Ag NPs via a known spillover effect. Thus, highly active negative O ions spilled on the ZnO surface, which extracted the electrons from the conduction band of ZnO, thereby increasing the thickness of the electron-depletion layer and consequently the sensor resistance. Furthermore, the active O species accelerated the reaction of the sensor with TEA. Overall, the widely dispersed Ag NPs were highly favorable to the spillover effect and might catalyze the sensing reaction to increase the response of the sensor to TEA (Figure 7a,b). Finally, the special hierarchical 3D porous nanostructure comprised hundreds of porous nanosheets was an additional meaningful component for the excellent TEA sensing properties of the Ag-loaded gas sensor. This unique structure with a high surface area and large pores (Figure 7c,d) increased the regions of the sensing reaction and promoted the TEA sensing performance of the sensor.

3.5. Ag-Decorated/Loaded Formaldehyde (HCHO) Gas Sensors

HCHO is extensively utilized in several industrial applications. Typically, HCHO is a colorless gas with a pungent odor, and it easily evaporates from the products into the indoor air. Many health-related issues, including headaches, nausea, cancer, and mucosal and respiratory irritation, are associated with exposure to HCHO [61,62,63,64]. Therefore, the development of highly sensitive HCHO gas sensors is urgently required to ensure the safety of people. In this regard, Xing et al. [65] prepared highly porous 1–5 at % Ag-loaded ZnO for application in HCHO sensors. In this study, the sensor fabricated using 1 at % Ag-loaded ZnO exhibited the maximum response to HCHO (170.42) at an optimal temperature of 240 °C. Unique hierarchically structured porous Ag-loaded ZnO provided an adequate surface area for interaction between the HCHO molecules and the sensing material. The high response of the optimal gas sensor to HCHO was ascribed to the production of heterojunctions between Ag and ZnO in addition to the catalytic activity of Ag. Schematics of the energy bands of Ag/ZnO in the presence of air and HCHO gas are shown in Figure 8a,b, respectively. Herein, the electrons transferred from Ag to ZnO because the work function of Au (Wm = 4.4 eV) is smaller than that of ZnO (Ws = 5.2 eV), generating an electron-depletion layer because of the increase in the concentration of electrons at the Ag/ZnO interface. When the Ag/ZnO sensor was exposed to air, the electrons released from ZnO were captured by the adsorbed O ions to form Ox, which showed strong oxidation activity. This process reduced the sensor resistance. Upon exposure to HCHO, the previously adsorbed Ox strongly reacted with the HCHO molecules and oxidized HCHO into CO2 and H2O, thereby decreasing the sensor resistance. This study reveals that the HCHO sensing performance of the sensor is directly influenced by the Ag/ZnO heterojunction and the developed Ox.
Furthermore, Wang et al. [66] reported the HCHO sensing activity of a two-step solution-processed Ag-loaded sunflower-like hierarchical In2O3 nanoarchitecture. The observed HCHO sensing performance was attributed to the highly porous and hierarchically designed sunflower-like nanostructure and the catalytic activity of Ag. The exclusive hierarchical 3D sunflower-like nanostructures possess numerous radial nanobranches with irregular surfaces, which offer a larger exposed surface area and additional paths for the exchange of electrons during the entire gas diffusion and surface reaction processes. An appropriate Ag loading amount plays an important role in chemical and electronic sensitization during the detection of HCHO by the Ag-loaded In2O3 sensor. Chemical sensitization favored gas reactions by dissociating HCHO via a spillover effect, and electronic sensitization substantially enhanced the direct exchange of electrons between In2O3 and Ag. However, higher Ag loading amount resulted in the conduction of electrons along the metallic Ag NPs irrespective of chemoresistive variations, thereby deteriorating the sensor response.

3.6. Ag-Decorated/Loaded Carbon Monoxide (CO) Gas Sensors

CO is among the most hazardous gases as it is invisible, tasteless, odorless, and colorless and thus cannot be detected by human sensory organs [67]. Consequently, numerous efforts have been made to detect CO gas by various materials and strategies. Molybdenum disulfide (MoS2) is a distinct graphene-like two-dimensional (2D) layered transition metal dichalcogenide. In addition to a conventional band gap of 1.8 eV, it has a large surface area-to-volume ratio with excellent physical/mechanical properties [68]. Nevertheless, the gradual degradation of MoS2 nanosheets occurs under ambient conditions owing to atmospheric oxidation and surface contamination, which may eventually decrease their sensing performances. Thus, to achieve an appropriate CO sensing performance of pure MoS2 nanosheets, sensing should be performed under an inert N2 atmosphere, significantly restricting their commercial applications. To overcome this limitation, Zhang et al. [69] fabricated a novel ternary Ag-loaded ZnO/MoS2 nanocomposite via layer-by-layer self-assembly for the detection of CO gas.
The presence of ZnO NRs, Ag NPs, and MoS2 nanosheets was verified by morphological investigations (the scanning electron microscopy (SEM) image is shown in Figure 9a). The developed nanocomposite sensor demonstrated excellent room temperature CO sensing performance. This confirmed that the existence of MoS2 might reduce the operating temperature and energy consumption probably because of the higher specific surface area and better conductivity of MoS2 along with the synergistic effect between ZnO and MoS2. Ag effectively improved the catalytic activity of the sensor and mobilities of the carriers for the interaction of gas with the surface of MoS2 or ZnO by modifying the energy band structure and surface morphology of the sensing material. Furthermore, substantial dynamic sites for CO adsorption might be generated upon Ag loading. After Ag modification, the interactions of CO molecules with O ions dramatically reduced the resistance of the Ag–ZnO/MoS2 nanocomposites as compared to that of the ZnO/MoS2 film, resulting in a high CO response of the sensor. A schematic of the interactions between the CO molecules and the Ag–ZnO/MoS2 sensor is depicted in Figure 9b. Herein, Ag and Pt were employed to improve the CO sensing performance of the ZnO/MoS2 composite sensor. The Ag-loaded sensor exhibited the maximum response as compared to that of the Pt-loaded sensor under similar experimental conditions. The higher response of the Ag-loaded sensor was explained with respect to the work function. Upon the introduction of a noble metal into ZnO/MoS2, the electrical characteristics of ZnO/MoS2 may be influenced by the existence of a Schottky barrier. Therefore, noble metals with lower work functions are preferred to reduce the Schottky barrier and thus improve the response of the corresponding ZnO/MoS2 sensor. As Ag has a lower work function (4.26 eV) than that of Pt (5.65 eV), the Ag–ZnO/MoS2-based sensor demonstrated higher CO sensing performance.
In another study, Ag-modified flower-like ZnO microspheres were prepared using a combination of solvothermal strategy and an impregnation approach [70]. Dual selective sensing of methane (CH4) and CO using 1.5 at % Ag-loaded ZnO by monitoring the working temperature was reported in this study. Ag exhibits high catalytic activity for low-temperature oxidation of CO. However, the low-temperature oxidation of CH4 is very difficult to achieve, as CH4 is a thermally stable sp3-hybridized non-polar molecule. Accordingly, no considerable low-temperature response of the sensor to CH4 was observed. In contrast, at higher temperatures, CH4 was sufficiently activated by Ag because there was adequate thermal energy to overcome the barrier, thereby increasing the sensor response. Consequently, the sensor fabricated using 1.5 at % Ag-loaded ZnO exhibited temperature-modulated dual selectivity toward CO and CH4 at 130 and 200 °C, respectively.

3.7. Ag-Decorated/Loaded Ethanol (C2H5OH) Gas Sensors

C2H5OH is one of the most commonly used VOCs. Exposure to C2H5OH vapors may cause headache, drowsiness, eye irritation, liver damage, and breathing difficulties. Furthermore, the consumption of C2H5OH is the leading cause of motor vehicle accidents worldwide, and alcohol-impaired driving fatalities account for 31% of the total number of traffic-related deaths in the United States [71]. Thus, the design of low-temperature operating and selective sensors for the detection and monitoring of C2H5OH is required from the viewpoint of safety. Graphitic carbon nitride (gC3N4, g-CN) is a non-metallic polymer semiconductor with a 2D layered structure. As a result of its unique physicochemical characteristics, suitable band gap, and outstanding thermal and chemical stabilities, g-CN can be utilized as an active sensing element in gas-sensing applications. In this regard, Tomer et al. [72] fabricated well-ordered mesoporous Ag-decorated meso-CNs using a hard template by nanocasting to detect C2H5OH at trace levels. The response of cubic meso-CN to C2H5OH gas was approximately two times that of conventional g-CN. This improvement was caused by the higher meso-CN surface area along with uniform pore channels through which the target gas penetrated the deep portions of the sensor. The response of the 3 wt.% Ag-decorated meso-CNs to C2H5OH gas was two times that of meso-CN. A schematic of the C2H5OH sensing mechanism of the Ag/g-CN sensor is shown in Figure 10. In this case, the electrons transferred from g-CN to Ag due to the lower work function of g-CN (4.3 eV) than that of Ag (4.7 eV), thereby creating a potential barrier. This barrier prevented the transfer of electrons through the mesoporous sensor, and the electrons existing on the sensor surface developed dynamic sites for the adsorption of O2, thereby increasing the sensor response. Moreover, the loading of Ag NPs strongly accelerated catalytic oxidation and chemical sensitization, which ultimately enhanced the number of dynamic O species on the surface of meso-CN. In fact, the chemical sensitization effect (namely, spillover effect) of Ag NPs improved the rate of adsorption–desorption of molecular O2 on the sensor surface and transformed molecular O2 into O ions. Furthermore, the excellent catalytic activity of Ag NPs accelerated the decomposition of C2H5OH gaseous molecules into active radicals, thus enhancing the interactions between the chemisorbed O ions and C2H5OH molecules. Owing to the transfer of electrons from meso-CN to Ag NPs, a negatively charged layer was produced around the Ag/meso-CN interface. Under a C2H5OH atmosphere, the electrons returned to the sensing layer via the Ag2O/Ag redox reaction, and thus, the areas along the Ag/meso-CN interface became very sensitive to C2H5OH.

3.8. Ag-Decorated/Loaded Nitrogen Dioxide (NO2) Gas Sensors

NO2 is a highly poisonous and polluting gas that produces acid rain and photochemical smog [73]. As a result of its highly toxic nature, detection and monitoring systems for NO2 are needed. Accordingly, Wang et al. conducted NO2 gas-sensing studies employing Ag-loaded mesoporous WO3 designed using 3D cubic KIT-6 as a hard template [74]. Herein, Ag at different molar ratios of 0.2, 0.5, and 1.0% was loaded onto mesoporous WO3. The sensor developed using 0.5% Ag exhibited a higher response of 44 to 1 ppm NO2 at 75 °C because of its larger surface area and diffusion paths. NO2 is a polar molecule having a positive charge localized on the N atom and a negative charge over the O atom. The interaction of electrons with Ag repels the O atom and attracts the positively charged N atom. Ag was used as a dynamic catalyst to create more active sites for the detection of NO2. Although the addition of Ag may enhance the sensing properties of the material, the loading of Ag at excessively high concentrations can decrease the catalytic performance of the material. When the Ag concentration exceeded 0.5%, the total surface area reduced, and interconnected Ag NPs changed the electron path to be from Ag instead of WO3.
The advantages of UV-activated sensors are their high stabilities, abilities to detect flammable gases, and low energy consumptions. In a typical UV-based gas-sensing device, the UV wavelength and power intensity determine the energy of the photons arriving at the exposed surface. This energy influences the steady-state surface reactions and plays a significant role in the response curve. Espid et al. [75] explained the NO2 gas detection characteristics of Ag-loaded ZnO nanoellipsoids, which were synthesized by a simple coprecipitation method, under UV illumination. When the surface of the gas sensor was irradiated with the photons released from a UV light-emitting diode, the electrons of the sensor hopped from the valence band to the conduction band, leaving holes in the valence band. A schematic of electron mobilization and the related reactions of the Ag-loaded ZnO sensor with NO2 gas under UV illumination is depicted in Figure 11. Photogenerated electron/hole pairs supported the direct adsorption of O2 molecules. Under an air atmosphere, O2 interacted with the excited electrons and adsorbed in the ionic form on the sensor surface. This slightly increased the sensor resistance. Under a NO2 atmosphere, the gas molecules reacted with the chemisorbed oxygen ions or directly adsorbed on the sensor surface by accepting electrons. A change in the electron flow stimulated by these reactions modulated the electrical resistivity. Under continuous UV irradiation, when the NO2 flow was stopped, the accelerated photons eliminated the NO2 species from the sensor surface. The observed enhancement in the NO2 detection performance of ZnO nanoellipsoids after Ag loading was associated with the synergistic effects of the semiconductor composite and Ag. Additionally, the improvement of the sensing response in the presence of Ag was linked to the increased electron utilization ratio and the decreased recombination rate of photogenerated electrons and holes due to the trapping of electrons in the Ag NPs during excitation. Similarly, the adsorption ability was directly correlated with the number of unoccupied oxygen sites. When Ag was embedded in the ZnO lattice, O-vacancies formed because of the differences between the charges of Ag+ and Zn2+ ions, which indicated that additional adsorption sites were created, resulting in a higher sensor response.
Furthermore, Zhang et al. [76] designed a room-temperature light-assisted NO2 gas sensor using Ag-decorated ZnO NPs. The sensor developed using 3 mol.% Ag exhibited the highest NO2 response. A schematic of the NO2 detection mechanism of pure and Ag-loaded ZnO NPs under various light conditions is shown in Figure 12. Herein, characterization studies revealed the development of a heterojunction between Ag NPs and ZnO NPs. Nevertheless, the improved NO2 detection performance of the sensor after Ag loading was ascribed to prominent interactions between Ag NPs and the O vacancies. In this study, Ag NPs and surface oxygen vacancies acted as electron sinks to boost charge separation and catalytic activities of Ag NPs for promoting the adsorption of gas molecules. The effective separation of photogenerated electron–hole pairs due to the interfacial charge transfer between ZnO and Ag and charge-carrier trapping through the surface oxygen vacancies enabled further accumulation of unpaired electrons on the surfaces of ZnO NPs. Moreover, Ag NPs and oxygen vacancies on the surface enhanced the adsorption of gas molecules, and thus, more oxygen molecules adsorbed on the sensor surface. An excess of Ag NPs reversibly acted as charge-carrier recombination centers, leading to negatively charged electrostatic attraction between Ag and the positively charged holes, thereby reducing the photoquantum efficiency. Additionally, sometimes, it is difficult to completely desorb the molecular or atomic oxygen species chemisorbed on the Ag surface through light irradiation at room temperature. In brief, O2 molecules play an essential role in the absorption/desorption of NO2 molecules on the surfaces of ZnO NPs. Specifically, the exposure degree of O2(ads) (hν) directly stimulates the sensing characteristics, such as response and response/recovery speeds, of light-assisted NO2 sensors.

3.9. Ag-Decorated/Loaded Methyl Mercaptan (CH3SH) Gas Sensors

CH3SH is a gas that is commonly used as an additive to other gases including propane and natural gas. The rotten egg odor of CH3SH facilitates the detection of CH3SH leakage. The recommended airborne exposure limit for CH3SH is 0.5 ppm during an 8 h work shift. Exposure to higher levels of CH3SH results in eye and throat irritation, drowsiness, and even bronchitis [77]. Consequently, designing a sensor that can detect CH3SH at low levels is highly desirable. Additionally, CH3SH gas-sensing properties of SMOs have rarely been reported. In this regard, Garcia et al. [77] constructed a sensor using mesoporous Ag-loaded hematite (α-Fe2O3) to detect CH3SH at room temperature. Herein, the loading of 3 wt.% Ag onto α-Fe2O3 significantly enhanced the sensor response owing to the deeper electron-depletion layer. Ag loaded onto the surface of α-Fe2O3 served as a catalyst to accelerate the rate of conversion of O2 into ionic O. Therefore, more electrons were trapped, which generated a thicker depletion layer. When the Ag-loaded α-Fe2O3 sensor was exposed to CH3SH, the CH3SH molecules favorably chemisorbed due to the strong affinities of thiol groups for metallic Ag. Ag loading accelerated the reaction of CH3SH with oxygen ions via a spillover effect (Figure 13). In this case, the deep electron-depletion layer converted into a flat layer, which resulted in a better sensor response. However, when the concentration of Ag exceeded 3 wt.%, the Ag NPs generated a connected network on the surface of α-Fe2O3, and consequently, the resistance of the sensor reduced. Thus, the adsorption of O2 and the gas–O2 interaction substantially reduced, causing an inferior response of the sensor to the target gas.

3.10. Ag-Decorated/Loaded Xylene (C8H10) Gas Sensors

C8H10 is a VOC that is colorless, odorless, and highly toxic in nature [78,79]. It causes severe damage to the human body, even at very low levels and upon long-term exposure [21]. Prolonged exposure to C8H10, even at low levels, is very harmful to living organisms; therefore, detection and monitoring of C8H10 is necessary. In a study reported by Zhang et al. [80], C8H10 detection properties of 0.2, 0.5, and 1 at % Ag-loaded hedgehog-like TiO2 architectures are described. Herein, Ag-loaded hedgehog-like TiO2 nanostructures composed of hundreds of 1D NRs were synthesized via a simple hydrothermal process followed by isometric impregnation. Gas-sensing studies showed that the 0.5 at % Ag-loaded TiO2 sensor demonstrated the highest C8H10 response as compared to those of the 0.2 and 1 at % Ag-loaded TiO2 sensors. A schematic of the sensing interactions between the C8H10 gas molecules and Ag/TiO2 is depicted in Figure 14. A typical gas-sensing mechanism of n-type SMOs, based on the change in resistance upon the interaction of these SMOs with gases, was used to explain the interactions between C8H10 and Ag/TiO2 [65,66,77]. The C8H10 responses observed for the pristine and 0.5 at % Ag-loaded TiO2 sensors were 3.19 and ≈6.49 at 375 °C, respectively. The sensor response reduced when the Ag loading amount was increased up to 1 at %; this was mainly due to the coverage of the TiO2 surface sites by excessive Ag. Interestingly, all the Ag-loaded sensors exhibited fast responses and recovery dynamics; that is, the acquired response and recovery times were ≈5–7 and ≈1–2 s, respectively. The rapid response and recovery times were attributed to the strong electron mobility of the 1D TiO2 NRs and the superior porosity of the hedgehog-like Ag-loaded TiO2 architecture. The improvement in the C8H10 sensing characteristics of the TiO2 sensor upon Ag loading was ascribed to the catalytic activity of Ag and the development of Ag/TiO2 heterojunctions. The catalytic Ag NPs facilitated more and quicker reactions between the adsorbed oxygen ions and the C8H10 molecules.

3.11. Ag-Decorated/Loaded Ammonia (NH3) Gas Sensors

NH3 is a hazardous gas that is colorless and toxic. It has significant applications in various areas such as compound fertilizers, synthetic fibers, and biofuels. Additionally, exposure to NH3 at high levels may result in serious effects, including irritation of the eyes/skin/throat and the respiratory tract, on the human body [81,82]. Karaduman et al. [83] reported the NH3-sensing characteristics of Ag-decorated rGO. In this case, when NH3 gas molecules adsorbed on the surface of rGO by physisorption, the holes of rGO interacted with the electron-donating NH3 gas. Therefore, the degree of delocalization of the conjugated p-electrons of the detecting surface enhanced through the transfer of charge carriers from the adsorbed NH3 molecules. This process reduced the concentrations of charge carriers, thereby increasing the electrical resistance of the sensing material. The enhanced NH3 sensing response of the Ag-loaded rGO sensor was attributed to the following factors: (i) The functional groups and defects in rGO provided numerous active sites for NH3 adsorption, increasing the gas adsorption capacity. (ii) The existence of Ag NPs resulted in the adsorption of more NH3 molecules on the rGO surface as NH3 molecules strongly bound to the rGO surface. The catalytic characteristics of Ag NPs stimulated the breakdown of NH3 into active radicals and enhanced the reaction between chemisorbed O ions and NH3 molecules. (iii) The development of an n-AgxOy/p-rGO heterointerface led to a depletion region, which concurrently lowered the concentrations of carriers on both sides. When the sensor was exposed to NH3 gas, the electronic conductance of the sensor considerably increased, which improved the sensing response.
Organic/inorganic hybrid nanocomposites possess unique characteristics owing to the combined properties of polymers and SMOs. Additionally, decoration with noble metals, including Ag, may enhance the unique characteristics of the hybrid materials because of the catalytic activity of Ag. Qin et al. [84] reported the NH3-sensing properties of Ag-loaded polypyrrole@Si nanowire (Ppy@SiNW) core–shell structures (Figure 15).
Herein, electrical transport occurred in both the shell and the nearby core at the Si/PPy interface because of the thin layer of the PPy shell. Moreover, the Ag-loaded sensor exhibited baseline resistance that was higher than that of the pristine sensor, indicating strong electronic sensitization effect of the Ag NPs. NH3 sensing measurement was conducted using the Ag-loaded PPy@SiNW sensor under high ambient humidity conditions. In this case, a higher NH3 response and weaker humidity interference were noticed following the loading of Ag NPs onto the PPy@SiNW sensor. Actually, the Ag NPs coordinated with –N groups (tertiary N) in the organic chains. Note that the tertiary groups of N are the general coordination sites of the water molecules adsorbed on the polymer surface. Therefore, throughout the polymerization of PPy, Ag0 preferably coordinated with –N groups in PPy, which inhibited further adsorption of water on similar –N groups. Simultaneously, the loaded Ag NPs were highly hydrophobic. Thus, no significant water adsorption occurred on the surfaces of the Ag NPs, even at 100% relative humidity. Accordingly, the enhanced anti-humidity interference properties of the Ag-PPy@SiNWs originated from the attenuation of water adsorption on the Ag-loaded PPy shell. The Ag NPs loaded on the PPy chain demonstrated high hydrophobicity (Figure 16), which resulted in nearly no water adsorption on the PPy shells. The minimal adsorption of water molecules was also beneficial for the electronic and chemical effects of Ag NPs. These factors also caused higher response amplitudes at high ambient humidity [84].

3.12. Summary of Ag-Decorated Gas Sensors

In the above sections, we have described the effect of Ag decoration on the gas response of resistive-based gas sensors. Table 1 [48,49,50,51,52,54,60,65,66,69,70,72,74,76,77,80,85,86,87,88,89,90,91,92,93,94,95] shows the gas-sensing performances of Ag-loaded SMO-based sensors for various toxic gases. In addition, other researchers [95,96,97,98,99,100,101,102,103,104,105,106,107,108,109] have reported enhanced gas sensing after Ag decoration. Depending on the type of SMO and the synergistic effects between Ag and SMO, the sensing temperature can vary from room temperature to high temperatures. Furthermore, in some cases, higher responses were realized after loading Ag NPs onto the surfaces of SMOs. Overall, in the presence of Ag NPs, not only the sensing temperatures reduced but also higher responses were attained when compared with those of the pure gas sensors, and selectivity improved owing to the catalytic activities of the Ag NPs.

4. Ag-Doped Gas Sensors

Doped metal oxide is the structure where additives are incorporated in the lattice of host material, as shown in Figure 17a [110], and it can affect the gas response of the doped gas sensor (Figure 17b). In gas-sensing studies, very less attention has been devoted to Ag doping than to Ag loading.
Ag doping is often performed by incorporating Ag atoms into the SMO lattice. When Ag atoms are homogeneously dispersed in the semiconductor lattice, mobile charge carriers are generated, thereby changing the electrical properties of the gas sensor [111]. An overview of some Ag-doped gas sensors is presented in this section.
Hydrogen sulfide (H2S) is a highly toxic gas that is typically produced by the oil industry, natural gas plants, and sewage plants. Exposure to H2S at higher levels (namely, 100 ppm) may cause sudden collapse with respiratory loss, and the possibility of fatality is very high [112].
Ovsianytskyi et al. [113] developed a H2S gas sensor using Ag NP-doped graphene. When the Ag-doped graphene sensor was exposed to H2S gas, H2S adsorbed on Ag rather than on C because Ag is less electronegative than C; the adsorption of H2S led to a possible dissociation of H2S, and accordingly, SO2 and H2O were formed by the release of electrons (Figure 18). The released electrons entered graphene and recombined with intrinsic holes. This process decreased the concentrations of charge carriers and increased the resistance of Ag-doped graphene. Furthermore, Kolhe et al. [114] prepared SnO2 thin films doped with 1.5, 3.0, and 4.5 mol.% Ag by chemical spray pyrolysis for application in H2S sensors. Herein, the 3.0 mol.% Ag-doped SnO2 sensor exhibited the maximum H2S response. The enhanced response of SnO2 thin films toward H2S gas upon Ag doping was related to the catalytic activity of Ag or Ag2O and the formation of heterojunctions between Ag/SnO2 and Ag2O/SnO2. When the doping concentration of Ag was excessively high, the sensor response decreased because of the decrease in the number of dynamic sites owing to the accumulation of Ag grains. Additionally, Anand et al. [115] synthesized In2O3 and Ag-doped In2O3 NPs using a simple coprecipitation technique and utilized them for the detection of C2H5OH. Ag doping induced defects and vacancies in In2O3, which improved the number of dynamic sites on the sensor surface and accordingly increased the adsorption of C2H5OH. Consequently, the width of the depletion layer expanded, and thus, the resistance of the Ag-doped In2O3 sensor considerably increased. When the sensor was exposed to C2H5OH gas, the width of the depletion layer substantially reduced, which was a contributing factor to the sensing signal. Moreover, Ag acted as a catalyst and facilitated the sensor response by forming activated species of the chemisorbed oxygen ions, which were subsequently released to the In2O3 surface and caused the C2H5OH molecules to rapidly react with the chemisorbed oxygen ions.
Finally, Ding et al. [116] prepared 0.5, 1.0, and 3.0 mol.% Ag-doped hollow urchin-like In2O3 spheres via a one-step hydrothermal method for application in NO2 sensors. Herein, the sensor fabricated using 1.0 mol.% Ag demonstrated the maximum response to NO2 gas at an operating temperature of 62 °C. An excellent response of 190 was obtained to 1 ppm NO2 gas at 62 °C, which was almost 23 times that of the pristine In2O3 sensor. The improvement in the NO2 sensing response of the In2O3 sensor after Ag doping was ascribed to the chemical sensitization effect of Ag, which acted as a dynamic catalyst and produced more active sites on the sensor surface. The generated active sites caused the direct adsorption of NO2, which finally led to a higher NO2 response. When the Ag doping concentration was low, the catalytic effect was inadequate to make the entire surface of the sensor available for the adsorption of NO2 gas. Nevertheless, when 3.0 mol% Ag was doped, the sensor response decreased. Ag at higher doping concentrations covered the active sites on the sensor surface, which prevented the sensor from responding to the NO2 gas. Furthermore, the surface area of the optimal sensor was 100.6 m2/g, which also contributed to the improved gas response.
Table 2 presents the sensing properties of Ag-doped SMO-based gas-sensing devices [92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107]. Different gases at various temperatures can be detected by doping Ag into the pristine sensing device, which suggests the promising role of Ag as a noble metal dopant in the detection of toxic gases.

5. Conclusions and Perspectives

In this review, we have examined the promising effects of Ag addition on the detection performances of chemiresistive gas sensors. This review demonstrated that several toxic gases and VOCs can be efficiently detected by introducing Ag into SMOs and corresponding composite-based chemiresistive sensors. The gas-sensing performances of chemiresistive sensors, such as SMOs and related composites, can be enhanced by Ag doping owing to the catalytic activity and electronic/chemical sensitization effects of Ag. As an effective catalyst, Ag can attract abundant O2 molecules from the air and transfer them to the surfaces of SMOs, accordingly promoting the capture of electrons from SMOs by O2 molecules. Moreover, the higher electrical conductivity of Ag NPs facilitated rapid electron transfer and thereby improved the sensor response. Therefore, Ag may be an excellent choice as a sensitizer to improve the sensing performances of chemiresistive sensors as it offers additional active adsorption sites and charge transfer pathways to enhance surface reactions. Generally, the introduction of Ag in an optimal amount into the sensing material can lead to the best sensing properties, and a bell-shaped relationship typically exists between the sensor response and the addition amount of Ag [129]. Due to the low cost of Ag than that of other noble metals, the incorporation of Ag into gas sensors is a highly promising strategy to not only reduce the overall price of these sensors but also enhance their sensing properties.
Although different approaches have been introduced for the synthesis of Ag-based chemiresistive sensors, prospects for the development of Ag-based chemiresistive sensors are still considerable. Nevertheless, to date, the effect of Ag particle size on the gas sensor response has not been systematically investigated, and it should be realized in future efforts related to the decoration of Ag on the surfaces of SMOs and related composites. Additionally, the effects of annealing temperature on the final responses of Ag-decorated gas sensors have not been examined to date, and it may constitute another exciting area of research. In fact, less efforts have been dedicated toward studying Ag doping than those toward studying Ag decoration on the surfaces of SMOs for the detection of gases. Particularly, further Ag doping studies should be conducted in the future. Furthermore, selectivity toward a target gas remains a major concern for Ag-based SMOs and related composite-based resistive gas sensors. Additionally, most of the reported Ag-based chemiresistive sensors operate at higher temperatures, which hinders their commercialization. Therefore, the development of novel functional nanomaterials with superior characteristics is required to address this issue. Efforts should be made to develop room-temperature sensors to reduce energy consumption with enhanced sensitivity. For example, UV irradiation [130], bimetallic decoration [131], use of quantum dots [132] or their combinations [133] are among different strategies that can reduce the sensing temperature even to room temperature. In addition, a fluctuation-enhanced sensing (FSA) approach can be used to improve the sensitivity of the sensors [134,135,136,137].
Moreover, understanding the plausible detection mechanism of a specific target gas is a challenge in the field of chemiresistive sensors. Thus, to gain a deeper understanding of the basic sensing principles and respective surface adsorption/desorption kinetics between the target gas and sensing materials, appropriate technology or reasonable models need to be established. More importantly, all reported Ag-based chemiresistive sensors are laboratory-made and probably contain instrumental errors. Thus, further studies are necessary to enhance the robustness of the data, and the respective efforts must be dedicated to the commercialization of these devices in practice.

Author Contributions

Conceptualization, A.M., H.W.K. and S.S.K.; writing—original draft preparation, A.M., M.S. and S.N.; writing—review and editing, A.M., M.S. and S.N.; supervision, H.W.K. and S.S.K.; funding acquisition, H.W.K. and S.S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the PAP program (year: 2021) of the Korea Polar Research Institute and a National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2021R1A2C1009790).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Song, R.; Wang, Z.; Zhou, X.; Huang, L.; Chi, L. Gas-sensing performance and operation mechanism of organic π-conjugated materials. ChemPlusChem 2019, 84, 1222–1234. [Google Scholar] [CrossRef] [PubMed]
  2. Li, H.; Shi, W.; Song, J.; Jang, H.J.; Dailey, J.; Yu, J.; Katz, H.E. Chemical and biomolecule sensing with organic field-effect transistors. Chem. Rev. 2018, 119, 3–35. [Google Scholar] [CrossRef]
  3. Ali, S.; Gupta, A.; Shafiei, M.; Langford, S.J. Recent advances in perylene diimide-based active materials in electrical mode gas sensing. Chemosensors 2021, 9, 30. [Google Scholar] [CrossRef]
  4. Vajhadin, F.; Mazloum-Ardakani, M.; Amini, A. Metal oxide-based gas sensors for the detection of exhaled breath markers. Med. Devices Sens. 2021, 4, e10161. [Google Scholar] [CrossRef]
  5. Wang, Y.; Duan, L.; Deng, Z.; Liao, J. Electrically transduced gas sensors based on semiconducting metal oxide nanowires. Sensors 2020, 20, 6781. [Google Scholar] [CrossRef]
  6. Chowdhury, N.K.; Bhowmik, B. Micro/nanostructured gas sensors: The physics behind the nanostructure growth, sensing and selectivity mechanisms. Nanoscale Adv. 2021, 3, 73–93. [Google Scholar] [CrossRef]
  7. Majhi, S.M.; Mirzaei, A.; Navale, S.; Kim, H.W.; Kim, S.S. Boosting the sensing properties of resistive-based gas sensors by irradiation techniques: A review. Nanoscale 2021, 13, 4728–4757. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal oxide gas sensors: Sensitivity and influencing factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Godse, P.R.; Mane, A.T.; Navale, Y.H.; Navale, S.T.; Mulik, R.N.; Patil, V.B. Hydrothermally grown 1D ZnO nanostructures for rapid detection of NO2 gas. SN Appl. Sci. 2021, 3, 360. [Google Scholar] [CrossRef]
  10. Ji, H.; Zeng, W.; Li, Y. Gas sensing mechanisms of metal oxide semiconductors: A focus review. Nanoscale 2019, 11, 22664–22684. [Google Scholar] [CrossRef]
  11. Navale, S.T.; Yang, Z.B.; Liu, C.; Cao, P.J.; Patil, V.B.; Ramgir, N.S.; Mane, R.S.; Stadler, F.J. Enhanced acetone sensing properties of titanium dioxide nanoparticles with a sub-ppm detection limit. Sens. Actuators B Chem. 2018, 255, 1701–1710. [Google Scholar] [CrossRef]
  12. Lin, T.; Lv, X.; Hu, Z.; Xu, A.; Feng, C. Semiconductor metal oxides as chemoresistive sensors for detecting volatile organic compounds. Sensors 2019, 19, 233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cao, P.; Gui, X.; Navale, S.; Han, S.; Xu, W.; Fang, M.; Liu, X.; Zeng, Y.X.; Liu, W.; Zhu, D.; et al. Design of flower-like V2O5 hierarchical nanostructures by hydrothermal strategy for the selective and sensitive detection of xylene. J. Alloy. Compd. 2020, 815, 152378. [Google Scholar] [CrossRef]
  14. Bhati, V.S.; Kumar, M.; Banerjee, R. Gas sensing performance of 2D nanomaterials/metal oxide nanocomposites: A review. J. Mater. Chem. C 2021, 9, 8776–8808. [Google Scholar] [CrossRef]
  15. Meng, F.; Hou, N.; Jin, Z.; Sun, B.; Guo, Z.; Kong, L.; Xiao, X.; Wu, H.; Li, M.; Liu, J. Ag-decorated ultra-thin porous single-crystalline ZnO nanosheets prepared by sunlight induced solvent reduction and their highly sensitive detection of ethanol. Sens. Actuators B Chem. 2015, 209, 975–982. [Google Scholar] [CrossRef]
  16. Pandey, S.; Goswami, G.K.; Nanda, K.K. Nanocomposite based flexible ultrasensitive resistive gas sensor for chemical reactions studies. Sci. Rep. 2013, 3, 1–6. [Google Scholar] [CrossRef]
  17. Jimenez-Cadena, G.; Riu, J.; Rius, F.X. Gas sensors based on nanostructured materials. Analyst 2007, 132, 1083–1099. [Google Scholar] [CrossRef]
  18. Chen, L.; Tsang, S.C. Ag doped WO3-based powder sensor for the detection of NO gas in air. Sens. Actuators B Chem. 2003, 89, 68–75. [Google Scholar] [CrossRef]
  19. Kim, H.J.; Lee, J.H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B Chem. 2014, 192, 607–627. [Google Scholar] [CrossRef]
  20. Kwon, Y.J.; Kang, S.Y.; Mirzaei, A.; Choi, M.S.; Bang, J.H.; Kim, S.S.; Kim, H.W. Enhancement of gas sensing properties by the functionalization of ZnO-branched SnO2 nanowires with Cr2O3 nanoparticles. Sens. Actuators B Chem. 2017, 249, 656–666. [Google Scholar] [CrossRef]
  21. Kim, H.-J.; Yoon, J.-W.; Choi, K.-I.; Jang, H.W.; Umar, A.; Lee, J.-H. Ultraselective and sensitive detection of xylene and toluene for monitoring indoor air pollution using Cr-doped NiO hierarchical nanostructures. Nanoscale 2013, 5, 7066–7073. [Google Scholar] [CrossRef]
  22. Annanouch, F.E.; Haddi, Z.; Ling, M.; Maggio, F.D.; Vallejos, S.; Vilic, T.; Zhu, Y.; Shujah, T.; Umek, P.; Bittencourt, C.; et al. Aerosol-assisted CVD-grown PdO nanoparticle decorated tungsten oxide nanoneedles extremely sensitive and selective to hydrogen. ACS Appl. Mater. Inter. 2016, 8, 10413–10421. [Google Scholar] [CrossRef]
  23. Mirzaei, A.; Bang, J.H.; Kim, S.S.; Kim, H.W. Effect of noble metals on hydrogen sensing properties of metal oxide-based gas sensors. J. Sens. Sci. Tech. 2020, 29, 365–368. [Google Scholar]
  24. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B Chem. 2014, 204, 250–272. [Google Scholar] [CrossRef]
  25. Korotcenkov, G.; Cho, B.K. Engineering approaches for the improvement of conductometric gas sensor parameters: Part 1. Improvement of sensor sensitivity and selectivity (short survey). Sens. Actuators B Chem. 2013, 188, 709–728. [Google Scholar] [CrossRef]
  26. Rzaij, J.M.; Abass, A.M. Review on: TiO2 thin film as a metal oxide gas sensor. J. Chem. Rev. 2020, 2, 114–121. [Google Scholar] [CrossRef] [Green Version]
  27. Tian, X.; Cui, X.; Lai, T.; Ren, J.; Yang, Z.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Gas sensors based on TiO2 nanostructured materials for the detection of hazardous gases: A review. Nano Mater. Sci. 2021. [Google Scholar] [CrossRef]
  28. Dong, C.; Zhao, R.; Yao, L.; Ran, Y.; Zhang, X.; Wang, Y. A review on WO3 based gas sensors: Morphology control and enhanced sensing properties. J. Alloy. Compd. 2020, 820, 153194. [Google Scholar] [CrossRef]
  29. Hariharan, V.; Gnanavel, B.; Sathiyapriya, R.; Aroulmoji, V. A review on tungsten oxide (WO3) and their derivatives for sensor applications. Int. J. Adv. Sci. Eng. 2019, 5, 1163–1168. [Google Scholar] [CrossRef]
  30. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Kim, T.W. Recent advances in energy-saving chemiresistive gas sensors: A review. Nano Energy 2020, 79, 105369. [Google Scholar] [CrossRef] [PubMed]
  31. Abideen, Z.U.; Kim, J.H.; Lee, J.H.; Kim, J.Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Electrospun metal oxide composite nanofibers gas sensors: A review. J. Korean Ceram. Soc. 2017, 54, 366–379. [Google Scholar] [CrossRef] [Green Version]
  32. Amiri, V.; Roshan, H.; Mirzaei, A.; Sheikhi, M.H. A review of nanostructured resistive-based vanadium oxide gas sensors. Chemosensors 2020, 8, 105. [Google Scholar] [CrossRef]
  33. Al-Hashem, M.; Akbar, S.; Morris, P. Role of oxygen vacancies in nanostructured metal-oxide gas sensors: A review. Sens. Actuators B Chem. 2019, 301, 126845. [Google Scholar] [CrossRef]
  34. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B 2018, 229, 206–217. [Google Scholar] [CrossRef]
  35. Das, S.; Jayaraman, V. SnO2: A comprehensive review on structures and gas sensors. Prog. Mater. Sci. 2014, 66, 112–255. [Google Scholar] [CrossRef]
  36. Mirzaei, A.; Yousefi, H.R.; Falsafi, F.; Bonyani, M.; Lee, J.H.; Kim, J.H.; Kim, H.W.; Kim, S.S. An overview on how Pd on resistive-based nanomaterial gas sensors can enhance response toward hydrogen gas. Int. J. Hydrog. Energy 2019, 44, 20552–20571. [Google Scholar] [CrossRef]
  37. Korotcenkov, G.; Brinzari, V.; Cho, B.K. Conductometric gas sensors based on metal oxides modified with gold nanoparticles: A review. Microchim. Acta 2016, 183, 1033–1054. [Google Scholar] [CrossRef]
  38. Luo, Y.; Zhang, C.; Zheng, B.; Geng, X.; Debliquy, M. Hydrogen sensors based on noble metal doped metal-oxide semiconductor: A review. Int. J. Hydrog. Energy 2017, 42, 20386–20397. [Google Scholar] [CrossRef]
  39. Cattabiani, N.; Baratto, C.; Zappa, D.; Comini, E.; Donarelli, M.; Ferroni, M.; Ponzoni, A.; Faglia, G. Tin oxide nanowires decorated with Ag nanoparticles for visible light-enhanced hydrogen sensing at room temperature: Bridging conductometric gas sensing and plasmon-driven catalysis. J. Phys. Chem. C 2018, 122, 5026–5031. [Google Scholar] [CrossRef]
  40. Conner, W.C., Jr.; Falconer, J.L. Spillover in heterogeneous catalysis. Chem. Rev. 1995, 95, 759–788. [Google Scholar] [CrossRef]
  41. Yamazoe, N. New approaches for improving semiconductor gas sensors. Sens. Actuators B Chem. 1991, 5, 7–19. [Google Scholar] [CrossRef]
  42. Li, Z.; Yao, Z.; Haidry, A.A.; Plecenik, T.; Xie, L.; Sun, L.; Fatima, Q. Resistive-type hydrogen gas sensor based on TiO2: A review. Inter. J. Hydrog. Energy 2018, 43, 21114–21132. [Google Scholar] [CrossRef]
  43. Barbosa, M.S.; Suman, P.H.; Kim, J.J.; Tuller, H.L.; Orlandi, O.L. Investigation of electronic and chemical sensitization effects promoted by Pt and Pd nanoparticles on single crystalline SnO nanobelt-based gas sensors. Sens. Actuators B Chem. 2019, 301, 127055. [Google Scholar] [CrossRef]
  44. Molavi, R.; Sheikhi, M.H. Low temperature carbon monoxide gas sensor based on Ag-Co3O4 thick film nanocomposite. Mater. Lett. 2018, 233, 74–77. [Google Scholar] [CrossRef]
  45. Li, M.; Zhu, H.; Wang, B.; Cheng, J.; Yan, W.; Xia, S.; Tang, Z. Ultrasensitive and highly selective detection of methoxy propanol based on Ag-decorated SnO2 hollow nanospheres. Sens. Actuators B Chem. 2016, 232, 545–556. [Google Scholar] [CrossRef]
  46. Korotcenkov, G. Gas response control through structural and chemical modification of metal oxide films: State of the art and approaches. Sens. Actuators B Chem. 2005, 107, 209–232. [Google Scholar] [CrossRef]
  47. Amiri, V.; Roshan, H.; Mirzaei, A.; Neri, G.; Ayesh, A.I. Nanostructured metal oxide-based acetone gas sensors: A review. Sensors 2020, 20, 3096. [Google Scholar] [CrossRef]
  48. Xu, X.; Chen, Y.; Zhang, G.; Ma, S.; Lu, Y.; Bian, H.; Chen, Q. Highly sensitive VOCs-acetone sensor based on Ag-decorated SnO2 hollow nanofibers. J. Alloy. Compd. 2017, 703, 572–579. [Google Scholar] [CrossRef]
  49. Kılıç, A.; Alev, O.; Özdemir, O.; Arslan, L.Ç.; Büyükköse, S.; Öztürk, Z.Z. The effect of Ag loading on gas sensor properties of TiO2 nanorods. Thin Solid Film. 2021, 726, 138662. [Google Scholar] [CrossRef]
  50. Li, Q.; Zhang, W.; Wang, C.; Ma, J.; Ning, L.; Fan, H. Ag modified bismuth ferrite nanospheres as a chlorine gas sensor. RSC Adv. 2018, 8, 33156–33163. [Google Scholar] [CrossRef] [Green Version]
  51. Iftekhar Uddin, A.S.M.; Phan, D.-T.; Chung, G.-S. Low temperature acetylene gas sensor based on Ag nanoparticles-loaded ZnO-reduced graphene oxide hybrid. Sens. Actuators B Chem. 2015, 207, 362–369. [Google Scholar] [CrossRef]
  52. Lee, K.-W.; Uddin, A.S.M.I.; Phan, D.-T.; Chung, G.-S. Fabrication of low-temperature acetylene gas sensor based on Ag nanoparticles-loaded hierarchical ZnO nanostructures. Electron. Lett. 2014, 51, 572–574. [Google Scholar] [CrossRef]
  53. Gupta Chatterjee, S.; Chatterjee, S.; Ray, A.K.; Chakraborty, A.K. Graphene–metal oxide nanohybrids for toxic gas sensor: A review. Sens. Actuators B Chem. 2015, 221, 1170–1181. [Google Scholar] [CrossRef]
  54. Uddin, A.S.M.I.; Yaqoob, U.; Phan, D.-T.; Chung, G.-S. A novel flexible acetylene gas sensor based on PI/PTFE-supported Ag-loaded vertical ZnO nanorods array. Sens. Actuators B Chem. 2015, 222, 536–543. [Google Scholar] [CrossRef]
  55. Espinosa, E.H.; Ionescu, R.; Bittencourt, C.; Felten, A.; Erni, R.; Van Tendeloo, G.; Pireaux, J.J.; Llobet, E. Metal-decorated multi-wall carbon nanotubes for low temperature gas sensing. Thin Solid Film. 2007, 515, 8322–8327. [Google Scholar] [CrossRef]
  56. Mirzaei, A.; Leonardi, S.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  57. Ju, D.X.; Xu, H.Y.; Qiu, Z.W.; Zhang, Z.C.; Xu, Q.; Zhang, J.; Wang, J.Q.; Cao, B.Q. Near room temperature, fast-response, and highly sensitive triethylamine sensor assembled with Au-loaded ZnO/SnO2 core-shell nanorods on flat alumina substrates. ACS Appl. Mater. Interfaces 2015, 7, 19163–19171. [Google Scholar] [CrossRef]
  58. Mitsubayashi, K.; Kubotera, Y.; Yano, K.; Hashimoto, Y.; Kon, T.; Nakakura, S.; Nishi, Y.; Endo, H. Trimethylamine biosensor with flavin-containing monooxygenase type 3 (FMO3) for fish-freshness analysis. Sens. Actuators B Chem. 2004, 103, 463–467. [Google Scholar] [CrossRef]
  59. Li, W.; Xu, H.; Yu, H.; Zhai, T.; Xu, Q.; Yang, X.; Wang, J.; Cao, B. Different morphologies of ZnO and their triethylamine sensing properties. J. Alloy. Compd. 2017, 706, 461–469. [Google Scholar] [CrossRef]
  60. Shen, Z.; Zhang, X.; Mi, R.; Liu, M.; Chen, Y.; Chen, C.; Ruan, S. On the high response towards TEA of gas sensors based on Ag-loaded 3D porous ZnO microspheres. Sens. Actuators B Chem. 2018, 270, 492–499. [Google Scholar] [CrossRef]
  61. Castro-Hurtado, I.; Mandayo, G.G.; Castaño, E. Conductometric formaldehyde gas sensors. A review: From conventional films to nanostructured materials. Thin Solid Film. 2013, 548, 665–676. [Google Scholar] [CrossRef]
  62. Xiong, J.; Zhang, P.; Huang, S.; Zhang, Y. Comprehensive influence of environmental factors on the emission rate of formaldehyde and VOCs in building materials: Correlation development and exposure assessment. Environ. Res. 2016, 151, 734–741. [Google Scholar] [CrossRef]
  63. Wang, J.; Yunus, R.; Li, J.; Li, P.; Zhang, P.; Kim, J. In situ synthesis of manganese oxides on polyester fiber for formaldehyde decomposition at room temperature. Appl. Surf. Sci. 2015, 357, 787–794. [Google Scholar] [CrossRef]
  64. Dong, C.; Liu, X.; Han, B.; Deng, S.; Xiao, X.; Wang, Y. Nonaqueous synthesis of Ag-functionalized In2O3/ZnO nanocomposites for highly sensitive formaldehyde sensor. Sens. Actuators B Chem. 2016, 224, 193–200. [Google Scholar] [CrossRef]
  65. Xing, X.; Xiao, X.; Wang, L.; Wang, Y. Highly sensitive formaldehyde gas sensor based on hierarchically porous Ag-loaded ZnO heterojunction nanocomposites. Sens. Actuators B Chem. 2017, 247, 797–806. [Google Scholar] [CrossRef]
  66. Wang, S.; Xiao, B.; Yang, T.; Wang, P.; Xiao, C.; Li, Z.; Zhao, R.; Zhang, M. Enhanced HCHO gas sensing properties by Ag-loaded sunflower-like In2O3 hierarchical nanostructures. J. Mater. Chem. A 2014, 2, 6598–6604. [Google Scholar] [CrossRef]
  67. Nakate, U.T.; Patil, P.; Na, S.-I.; Yu, Y.T.; Suh, E.-k.; Hahn, Y.-B. Fabrication and enhanced carbon monoxide gas sensing performance of p-CuO/n-TiO2 heterojunction device. Colloids Surf. A Physicochem. Eng. 2020, 612, 125962. [Google Scholar] [CrossRef]
  68. Niakan, H.; Zhang, C.; Hu, Y.; Szpunar, J.A.; Yang, Q. Thermal stability of diamond-like carbon–MoS2 thin films in different environments. Thin Solid Film 2014, 562, 244–249. [Google Scholar] [CrossRef]
  69. Zhang, D.; Sun, Y.E.; Jiang, C.; Yao, Y.; Wang, D.; Zhang, Y. Room-temperature highly sensitive CO gas sensor based on Ag-loaded zinc oxide/molybdenum disulfide ternary nanocomposite and its sensing properties. Sens. Actuators B Chem. 2017, 253, 1120–1128. [Google Scholar] [CrossRef]
  70. Wang, Y.; Cui, Y.; Meng, X.; Zhang, Z.; Cao, J. A gas sensor based on Ag-modified ZnO flower-like microspheres: Temperature-modulated dual selectivity to CO and CH4. Surf. Interfaces 2021, 24, 101110. [Google Scholar] [CrossRef]
  71. Mirzaei, A.; Janghorban, K.; Hashemi, B.; Bonyani, M.; Leonardi, S.G.; Neri, G. Highly stable and selective ethanol sensor based on α-Fe2O3 nanoparticles prepared by Pechini sol–gel method. Ceram. Inter. 2016, 42, 6136–6144. [Google Scholar] [CrossRef]
  72. Tomer, V.K.; Malik, R.; Kailasam, K. Near-room-temperature ethanol detection using Ag-loaded mesoporous carbon nitrides. ACS Omega 2017, 2, 3658–3668. [Google Scholar] [CrossRef]
  73. Zhang, Q.; Zang, P.; Hu, W.; Li, J.; Liu, Y.; Liu, Y.; Yu, F.; Zhang, C.; Xu, M. Performance degradation mechanism of the light-activated room temperature NO2 gas sensor based on Ag-ZnO nanoparticles. Appl. Surf. Sci. 2021, 541, 148418. [Google Scholar] [CrossRef]
  74. Wang, Y.; Cui, X.; Yang, Q.; Liu, J.; Gao, Y.; Sun, P.; Lu, G. Preparation of Ag-loaded mesoporous WO3 and its enhanced NO2 sensing performance. Sens. Actuators B Chem. 2016, 225, 544–552. [Google Scholar] [CrossRef]
  75. Espid, E.; Taghipour, F. Facile synthesis and UV-activated gas sensing performance of Ag: ZnO nano-ellipsoids. ECS J. Solid State Sci. Technol. 2018, 7, 3089. [Google Scholar] [CrossRef] [Green Version]
  76. Zhang, Q.; Xie, G.; Xu, M.; Yu, S.; Tai, H.; Du, H.; Jiang, Y. Visible light-assisted room temperature gas sensing with ZnO-Ag heterostructure nanoparticles. Sens. Actuators B Chem. 2018, 259, 269–281. [Google Scholar] [CrossRef]
  77. Garcia, D.; Picasso, G.; Hidalgo, P.; Peres, H.E.M.; Sun Kou, R.; Gonçalves, J.M. Sensors based on Ag-loaded hematite (α-Fe2O3) nanoparticles for methyl mercaptan detection at room temperature. Anal. Chem. Res. 2016, 12, 74–81. [Google Scholar] [CrossRef]
  78. Mirzaei, A.; Kim, J.-H.; Kim, H.W.; Kim, S.S. Resistive-based gas sensors for detection of benzene, toluene and xylene (BTX) gases: A review. J. Mater. Chem. C 2018, 6, 4342–4370. [Google Scholar] [CrossRef]
  79. Zheng, Y.; Wang, L.; Tian, H.; Qiao, L.; Zeng, Y.; Liu, C. Bimetal carbonaceous templates for multi-shelled NiCo2O4 hollow sphere with enhanced xylene detection. Sens. Actuators B Chem. 2021, 339, 129862. [Google Scholar] [CrossRef]
  80. Zhang, Y.; Bai, J.; Zhou, L.; Du, L.; Liu, F.; Xiang, L.; Gao, Y.; Liu, F.; Yan, X.; Lu, G. Preparation of silver-loaded titanium dioxide hedgehog-like architecture composed of hundreds of nanorods and its fast response to xylene. J. Colloid Interface Sci. 2019, 536, 215–223. [Google Scholar] [CrossRef]
  81. Kwak, D.; Lei, Y.; Maric, R. Ammonia gas sensors: A comprehensive review. Talanta 2019, 204, 713–730. [Google Scholar] [CrossRef]
  82. Timmer, B.; Olthuis, W.; Aan, V.; Berg, D. Ammonia sensors and their applications-a review. Sens. Actuators B Chem. 2005, 107, 666–677. [Google Scholar] [CrossRef]
  83. Karaduman, I.; Er, E.; Çelikkan, H.; Erk, N.; Acar, S. Room-temperature ammonia gas sensor based on reduced graphene oxide nanocomposites decorated by Ag, Au and Pt nanoparticles. J. Alloy. Compd. 2017, 722, 569–578. [Google Scholar] [CrossRef]
  84. Qin, Y.; Zang, J.; Wen, Z. Synergistic functionalization of aligned silicon nanowires by Ag nanoparticles & PPy wrapping for improving gas-sensing response at high humidity level. Phys. E Low-Dimens. Syst. Nanostructures 2020, 118, 113957. [Google Scholar]
  85. Yoon, J.-W.; Hong, Y.J.; Chan Kang, Y.; Lee, J.-H. High performance chemiresistive H2S sensors using Ag-loaded SnO2 yolk-shell nanostructures. RSC Adv. 2014, 4, 16067–16074. [Google Scholar] [CrossRef]
  86. Uddin, A.S.M.I.; Lee, K.-W.; Chung, G.-S. Acetylene gas sensing properties of an Ag-loaded hierarchical ZnO nanostructure-decorated reduced graphene oxide hybrid. Sens. Actuators B Chem. 2015, 216, 33–40. [Google Scholar] [CrossRef]
  87. Zhou, L.; Bai, J.; Liu, Y.; Liu, F.; Wang, H.; Zhang, Y.; Lu, G. Highly sensitive C2H2 gas sensor based on Ag modified ZnO nanorods. Ceram. Inter. 2020, 46, 15764–15771. [Google Scholar] [CrossRef]
  88. Yousefi, H.R.; Hashemi, B.; Mirzaei, A.; Roshan, H.; Sheikhi, M.H. Effect of Ag on the ZnO nanoparticles properties as an ethanol vapor sensor. Mater. Sci. Semicond. Proc. 2020, 117, 105172. [Google Scholar] [CrossRef]
  89. Nadargi, D.Y.; Bateer, D.R.; Tamboli, M.S.; Mulla, I.S.; Suryavanshi, S.S. A greener approach towards the development of graphene-Ag loaded ZnO nanocomposites for acetone sensing applications. RSC Adv. 2019, 9, 33602–33606. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, H.; Li, Q.; Zheng, X.; Wang, C.; Ma, J.; Yan, B.; Du, Z.; Li, M.; Wang, W.; Fan, H. 3D porous flower-like ZnO microstructures loaded by large-size Ag and their ultrahigh sensitivity to ethanol. J. Alloy. Compd. 2020, 829, 154453. [Google Scholar] [CrossRef]
  91. Xue, Y.-Y.; Wang, J.-L.; Li, S.-N.; Jiang, Y.-C.; Hu, M.-C.; Zhai, Q.-G. Mesoporous Ag/In2O3 composite derived from indium organic framework as high performance formaldehyde sensor. J. Solid State Chem. 2017, 251, 170–175. [Google Scholar] [CrossRef]
  92. Jiang, C.; Zhang, D.; Yin, N.; Yao, Y.; Shaymurat, T.; Zhou, X. Acetylene gas-sensing properties of layer-by-layer self-assembled Ag-decorated tin dioxide/graphene nanocomposite film. Nanomaterials 2017, 7, 278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Chen, M.; Zhang, D.; Hu, J.; Wang, H.; Zhang, Y.; Li, K.; Rong, Q.; Zhou, S.; Zhang, J.; Zhu, Z.; et al. Excellent toluene gas sensing properties of molecular imprinted Ag-LaFeO3 nanostructures synthesized by microwave-assisted process. Mater. Res. Bull. 2019, 111, 320–328. [Google Scholar] [CrossRef]
  94. Liu, M.; Song, P.; Liang, D.; Yang, Z.; Wang, Q. Highly sensitive and selective triethylamine gas sensor based on Ag nanoparticles-decorated MoO3 nanobelts. Mater. Res. Express 2019, 6, 125910. [Google Scholar] [CrossRef]
  95. Liu, H.; Shen, W.; Chen, X. A room temperature operated ammonia gas sensor based on Ag-decorated TiO2 quantum dot clusters. RSC Adv. 2019, 9, 24519–24526. [Google Scholar] [CrossRef] [Green Version]
  96. Kamble, C.; Panse, M.; Nimbalkar, A. Ag decorated WO3 sensor for the detection of sub-ppm level NO2 concentration in air. Mater. Sci. Semicond. Proc. 2019, 103, 104613. [Google Scholar] [CrossRef]
  97. Ghanbari, R.; Safaiee, R.; Sheikhi, M.H.; Golshan, M.M.; Horastani, Z.K. Graphene decorated with silver nanoparticles as a low-temperature methane gas sensor. ACS Appl. Mater. Interfaces 2019, 11, 21795–21806. [Google Scholar] [CrossRef]
  98. Liu, D.; Pan, J.; Tang, J.; Liu, W.; Bai, S.; Luo, R. Ag decorated SnO2 nanoparticles to enhance formaldehyde sensing properties. J. Phys. Chem. Solids 2019, 124, 36–43. [Google Scholar] [CrossRef]
  99. Lin, Z.D.; Young, S.J.; Hsiao, C.H.; Chang, S.J. Adsorption sensitivity of Ag-decorated carbon nanotubes toward gas-phase compounds. Sens. Actuators B Chem. 2013, 188, 1230–1234. [Google Scholar] [CrossRef]
  100. Li, Z.; Zhang, G.; Gao, W.; Zhao, R.; Wang, Y. Ag decorated ZnO nanocrystallines synthesized by a low-temperature solvothermal method and their application for high response H2 gas sensor. J. Mater. Sci. Mater. Electron. 2019, 30, 18959–18969. [Google Scholar]
  101. Shim, Y.S.; Zhang, L.; Kim, Y.H.; Choi, Y.R.; Nahm, S.H.; Kang, C.Y.; Lee, W.; Jang, H.W. Highly sensitive and selective H2 and NO2 gas sensors based on surface-decorated WO3 nanoigloos. Sens. Actuators B Chem. 2014, 198, 294–301. [Google Scholar] [CrossRef]
  102. Navaneethan, M.; Patil, V.L.; Ponnusamy, S.; Muthamizhchelvan, C.; Kawasaki, S.; Patil, P.S.; Hayakawa, Y. Sensitivity enhancement of ammonia gas sensor based on Ag/ZnO flower and nanoellipsoids at low temperature. Sens. Actuators B Chem. 2018, 255, 672–683. [Google Scholar]
  103. Yin, Y.; Li, F.; Zhang, N.; Ruan, S.; Zhang, H.; Chen, Y. Improved gas sensing properties of silver-functionalized ZnSnO3 hollow nanocubes. Inorg. Chem. Front. 2018, 5, 2123–2131. [Google Scholar] [CrossRef]
  104. Choi, M.S.; Kim, M.Y.; Ahn, J.; Choi, S.J.; Lee, K.H. Ag-functionalized SnO2 nanowires-based sensor for NO2 detection at low operating temperature. J. Microelectron. Packag. Soc. 2020, 27, 11–17. [Google Scholar]
  105. Park, S.; An, S.; Ko, H.; Jin, C.; Lee, C. Enhancement of ethanol sensing of TeO2 nanorods by Ag functionalization. Curr. Appl. Phys. 2013, 13, 576–580. [Google Scholar] [CrossRef]
  106. Zhou, S.; Chen, M.; Lu, Q.; Zhang, Y.; Zhang, J.; Li, B.; Wei, H.; Hu, J.; Wang, H.; Liu, Q. Ag nanoparticles sensitized In2O3 nanograin for the ultrasensitive HCHO detection at room temperature. Nanoscale Res. Lett. 2019, 14, 1–11. [Google Scholar] [CrossRef] [Green Version]
  107. Liu, J.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. A high-response formaldehyde sensor based on fibrous Ag-ZnO/In2O3 with multi-level heterojunctions. J. Hazard. Mater. 2021, 413, 125352. [Google Scholar] [CrossRef] [PubMed]
  108. Yu, H.; Li, J.; Li, Z.; Tian, Y.; Yang, Z. Enhanced formaldehyde sensing performance based on Ag@WO3 2D nanocomposite. Powder Tech. 2019, 343, 1–10. [Google Scholar] [CrossRef]
  109. Feng, D.L.; Zhu, Z.Y.; Du, L.L.; Xing, X.; Wang, C.; Chen, J.; Tian, Y.T.; Yang, D.C. Improved sensing performance of WO3 nanoparticles decorated with Ag and Pt nanoparticles. Rare Met. 2021, 40, 1642–1650. [Google Scholar] [CrossRef]
  110. Degler, D.; Weimar, U.; Barsan, N. Current understanding of the fundamental mechanisms of doped and loaded semiconducting metal-oxide-based gas sensing materials. ACS Sens. 2019, 4, 2228–2249. [Google Scholar] [CrossRef]
  111. Ristein, J. Surface transfer doping of semiconductors. Sci. N. Y. Wash. 2006, 313, 1057. [Google Scholar] [CrossRef]
  112. Mirzaei, A.; Kim, S.S.; Kim, H.W. Resistance-based H2S gas sensors using metal oxide nanostructures: A review of recent advances. J. Haz. Mater. 2018, 357, 314–331. [Google Scholar] [CrossRef]
  113. Ovsianytskyi, O.; Nam, Y.-S.; Tsymbalenko, O.; Lan, P.-T.; Moon, M.-W.; Lee, K.-B. Highly sensitive chemiresistive H2S gas sensor based on graphene decorated with Ag nanoparticles and charged impurities. Sens. Actuators B Chem. 2018, 257, 278. [Google Scholar] [CrossRef]
  114. Kolhe, P.S.; Koinkar, P.M.; Maiti, N.; Sonawane, K.M. Synthesis of Ag doped SnO2 thin films for the evaluation of H2S gas sensing properties. Phys. B Condens. Matter 2017, 524, 90–96. [Google Scholar] [CrossRef]
  115. Anand, K.; Kaur, J.; Singh, R.C.; Thangaraj, R. Preparation and characterization of Ag-doped In2O3 nanoparticles gas sensor. Chem. Phys. Lett. 2017, 682, 140–146. [Google Scholar] [CrossRef]
  116. Ding, M.; Xie, N.; Wang, C.; Kou, X.; Zhang, H.; Guo, L.; Sun, Y.; Chuai, X.; Gao, Y.; Liu, F.; et al. Enhanced NO2 gas sensing properties by Ag-doped hollow urchin-like In2O3 hierarchical nanostructures. Sens. Actuators B Chem. 2017, 252, 418–427. [Google Scholar] [CrossRef]
  117. Al-Hadeethi, Y.; Umar, A.; Ibrahim, A.A.; Al-Heniti, S.H.; Kumar, R.; Baskoutas, S.; Raffah, B.M. Synthesis, characterization and acetone gas sensing applications of Ag-doped ZnO nanoneedles. Ceram. Inter. 2017, 43, 6765–6770. [Google Scholar] [CrossRef]
  118. Hong, C.; Zhou, Q.; Lu, Z.; Umar, A.; Kumar, R.; Wei, Z.; Wu, X.; Xu, L.; Kim, S.H. Ag-doped ZnO nanoellipsoids based highly sensitive gas sensor. Mater. Express 2017, 7, 380–388. [Google Scholar] [CrossRef]
  119. Wei, W.; Guo, S.; Chen, C.; Sun, L.; Chen, Y.; Guo, W.; Ruan, S. High sensitive and fast formaldehyde gas sensor based on Ag-doped LaFeO3 nanofibers. J. Alloy. Compd. 2017, 695, 1122–1127. [Google Scholar] [CrossRef]
  120. Wang, Y.; Wang, Y.; Cao, J.; Kong, F.; Xia, H.; Zhang, J.; Zhu, B.; Wang, S.; Wu, S. Low-temperature H2S sensors based on Ag-doped α-Fe2O3 nanoparticles. Sens. Actuators B Chem. 2008, 131, 183–189. [Google Scholar] [CrossRef]
  121. Natkaeo, A.; Phokharatkul, D.; Hodak, J.H.; Wisitsoraat, A.; Hodak, S.K. Highly selective sub-10 ppm H2S gas sensors based on Ag-doped CaCu3Ti4O12 films. Sens. Actuators B Chem. 2018, 260, 571–580. [Google Scholar] [CrossRef]
  122. Zhang, Y.; Zheng, Z.; Yang, F. Highly sensitive and selective alcohol sensors based on Ag-dped In2O3 Coating. Indus. Eng. Chem. Res. 2010, 49, 3539–3543. [Google Scholar] [CrossRef]
  123. Zhang, R.; Ma, S.Y.; Zhang, Q.X.; Zhu, K.M.; Tie, Y.; Pei, S.T.; Wang, B.J.; Zhang, J.L. Highly sensitive formaldehyde gas sensors based on Ag doped Zn2SnO4/SnO2 hollow nanospheres. Mater. Lett. 2019, 254, 178–181. [Google Scholar] [CrossRef]
  124. Lu, Z.; Zhou, Q.; Xu, L.; Gui, Y.; Zhao, Z.; Tang, C.; Chen, W. Synthesis and characterization of highly sensitive hydrogen (H2) sensing device based on Ag doped SnO2 nanospheres. Materials 2018, 11, 492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Jeong, D.; Kim, K.; Park, S.-i.; Kim, Y.-h.; Kim, S.; Kim, S.-I. Characteristics of Ga and Ag-doped ZnO-based nanowires for an ethanol gas sensor prepared by hot-walled pulsed laser deposition. Res. Chem. Intermed. 2014, 40, 97–103. [Google Scholar] [CrossRef]
  126. Wang, J.; Zou, B.; Ruan, S.; Zhao, J.; Chen, Q.; Wu, F. HCHO sensing properties of Ag-doped In2O3 nanofibers synthesized by electrospinning. Mater. Lett. 2009, 63, 1750–1753. [Google Scholar] [CrossRef]
  127. Adilakshmi, G.; Reddy, R.S.; Reddy, A.S.; Reddy, P.S.; Reddy, C.S. Ag-doped WO3 nanostructure films for organic volatile gas sensor application. J. Mater. Sci. Mater. Electron. 2020, 31, 12158–12168. [Google Scholar] [CrossRef]
  128. Postica, V.; Vahl, A.; Santos-Carballal, D.; Dankwort, T.; Kienle, L.; Hoppe, M.; Cadi-Essadek, A.; De Leeuw, N.H.; Terasa, M.I.; Adelung, R.; et al. Tuning ZnO sensors reactivity toward volatile organic compounds via Ag doping and nanoparticle functionalization. Acs Appl. Mater. Interfaces 2019, 11, 31452–31466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Abideen, Z.U.; Kim, J.-H.; Kim, S.S. Optimization of metal nanoparticle amount on SnO2 nanowires to achieve superior gas sensing properties. Sens. Actuators B Chem. 2016, 238, 374–380. [Google Scholar] [CrossRef]
  130. Meng, F.; Zheng, H.; Sun, Y.; Li, M.; Liu, J. UV-activated room temperature single-sheet ZnO gas sensor. Micro Nano Lett. 2017, 12, 813–817. [Google Scholar] [CrossRef]
  131. Fan, F.; Zhang, J.; Li, J.; Zhang, N.; Hong, R.; Deng, X.; Tang, P.; Li, D. Hydrogen sensing properties of Pt-Au bimetallic nanoparticles loaded on ZnO nanorods. Sens. Actuators B Chem. 2017, 241, 895–903. [Google Scholar] [CrossRef]
  132. Mitri, F.; Iacovo, A.; Luca, M.; Pecora, A.; Colace, L. Lead sulphide colloidal quantum dots for room temperature NO2 gas sensors. Sci. Reports 2020, 10, 12556. [Google Scholar]
  133. Wu, T.; Wang, Z.; Tian, M.; Miao, J.; Zhang, H.; Sun, J. UV excitation NO2 gas sensor sensitized by ZnO quantum dots at room temperature. Sens. Actuators B Chem. 2018, 259, 526–531. [Google Scholar] [CrossRef]
  134. Ayhan, B.; Kwan, C.; Zhou, J.; Kish, L.B.; Benkstein, K.D.; Rogers, P.H.; Semancik, S. Fluctuation enhanced sensing (FES) with a nanostructured, semiconducting metal oxide film for gas detection and classification. Sens. Actuators B Chem. 2013, 188, 651–660. [Google Scholar] [CrossRef]
  135. Kish, L.B.; Smulko, J.; Heszler, P.; Granqvist, C.G. On the sensitivity, selectivity, sensory information and optimal size of resistive chemical sensors. Nanotechnol. Percept. 2007, 3, 43–52. [Google Scholar] [CrossRef]
  136. Ederth, J.; Smulko, J.M.; Kish, L.B.; Heszler, P.; Granqvist, C.G. Comparison of classical and fluctuation-enhanced gas sensing with PdxWO3 nanoparticle films. Sens. Actuators B Chem. 2006, 113, 310–315. [Google Scholar] [CrossRef]
  137. Kish, L.B.; Vajtai, R.; Granqvist, C.G. Extracting information from the noise spectra of chemical sensors: Electronic nose and tongue by one sensor. Sens. Actuators B Chem. 2000, 71, 55–59. [Google Scholar] [CrossRef]
Figure 1. General gas-sensing mechanism of SMO-based gas sensors (a) n- and p-type SMO in air (b) in oxidizing gas atmosphere and (c) in reducing gas atmosphere [19,20].
Figure 1. General gas-sensing mechanism of SMO-based gas sensors (a) n- and p-type SMO in air (b) in oxidizing gas atmosphere and (c) in reducing gas atmosphere [19,20].
Sensors 21 06454 g001
Figure 2. Schematic of chemical and electronic sensitizations in noble metal-decorated SMOs (in the figure, M = Metal, e = electron, O = adsorbed oxygen) [22,31].
Figure 2. Schematic of chemical and electronic sensitizations in noble metal-decorated SMOs (in the figure, M = Metal, e = electron, O = adsorbed oxygen) [22,31].
Sensors 21 06454 g002
Figure 3. Schematic representing the CH3COCH3 sensing mechanism of pure and Ag-doped SnO2. Reprinted from reference [48] with permission from Elsevier.
Figure 3. Schematic representing the CH3COCH3 sensing mechanism of pure and Ag-doped SnO2. Reprinted from reference [48] with permission from Elsevier.
Sensors 21 06454 g003
Figure 4. Energy band diagram and hole transfer in Ag NP-decorated BFO NSs (a) before and (b) after equilibrium [50]. Reproduced from https://pubs.rsc.org/en/content/articlelanding/2018/ra/c8ra06247a (accessed on 26 September 2018) from RSC. (This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence).
Figure 4. Energy band diagram and hole transfer in Ag NP-decorated BFO NSs (a) before and (b) after equilibrium [50]. Reproduced from https://pubs.rsc.org/en/content/articlelanding/2018/ra/c8ra06247a (accessed on 26 September 2018) from RSC. (This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence).
Sensors 21 06454 g004
Figure 5. (a) TEM and (b,c) HRTEM images of the ZG–Ag3 hybrid (inset shows the equivalent SAED pattern). Reprinted from reference [51] with permission from Elsevier.
Figure 5. (a) TEM and (b,c) HRTEM images of the ZG–Ag3 hybrid (inset shows the equivalent SAED pattern). Reprinted from reference [51] with permission from Elsevier.
Sensors 21 06454 g005
Figure 6. Images of (a) cross-section and (b) in-plane view of 8 s Ag-loaded ZnO NRs, (c) the Ag-decorated ZnO flexible gas sensor, and (d) the experimental setup at one bending angle. Reprinted from reference [54] with permission from Elsevier.
Figure 6. Images of (a) cross-section and (b) in-plane view of 8 s Ag-loaded ZnO NRs, (c) the Ag-decorated ZnO flexible gas sensor, and (d) the experimental setup at one bending angle. Reprinted from reference [54] with permission from Elsevier.
Sensors 21 06454 g006
Figure 7. Schematic of the (a,b) TEA gas-sensing mechanism of the Ag–ZnO sensor and (c,d) structures of the microspheres with the illustration of large pores. Reprinted from reference [60] with permission from Elsevier.
Figure 7. Schematic of the (a,b) TEA gas-sensing mechanism of the Ag–ZnO sensor and (c,d) structures of the microspheres with the illustration of large pores. Reprinted from reference [60] with permission from Elsevier.
Sensors 21 06454 g007
Figure 8. Schematics of the band structures of Ag-loaded ZnO in (a) air and (b) HCHO. Reprinted from reference [65] with permission from Elsevier.
Figure 8. Schematics of the band structures of Ag-loaded ZnO in (a) air and (b) HCHO. Reprinted from reference [65] with permission from Elsevier.
Sensors 21 06454 g008
Figure 9. (a) SEM image of Ag–ZnO/MoS2 and (b) schematic of the CO sensing mechanism of Ag–ZnO/MoS2. Reprinted from reference [69] with permission from Elsevier.
Figure 9. (a) SEM image of Ag–ZnO/MoS2 and (b) schematic of the CO sensing mechanism of Ag–ZnO/MoS2. Reprinted from reference [69] with permission from Elsevier.
Sensors 21 06454 g009
Figure 10. Band diagram of the Ag/g-CN sensor, elucidating the C2H5OH sensing mechanism of the sensor [72]. Reprinted from https://pubs.acs.org/doi/10.1021/acsomega.7b00479 (accessed on July 14, 2017) with permission from ACS (further permissions related to the material excerpted should be directed to the ACS).
Figure 10. Band diagram of the Ag/g-CN sensor, elucidating the C2H5OH sensing mechanism of the sensor [72]. Reprinted from https://pubs.acs.org/doi/10.1021/acsomega.7b00479 (accessed on July 14, 2017) with permission from ACS (further permissions related to the material excerpted should be directed to the ACS).
Sensors 21 06454 g010
Figure 11. Schematic of electron mobilization and related reactions of the Ag–ZnO sensor in the presence of NO2 gas [75]. Reproduced from https://iopscience.iop.org/article/10.1149/2.0141807jss (accessed on 10 April 2018) from IOPScience (This is an open access article distributed under the terms of the Creative Commons Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/ (accessed on 10 April 2018))).
Figure 11. Schematic of electron mobilization and related reactions of the Ag–ZnO sensor in the presence of NO2 gas [75]. Reproduced from https://iopscience.iop.org/article/10.1149/2.0141807jss (accessed on 10 April 2018) from IOPScience (This is an open access article distributed under the terms of the Creative Commons Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/ (accessed on 10 April 2018))).
Sensors 21 06454 g011
Figure 12. Schematic of the NO2 gas detection mechanism of pristine and Ag-loaded ZnO NPs irradiated with light of 365–520 nm wavelengths. Reprinted from reference [76] with permission from Elsevier.
Figure 12. Schematic of the NO2 gas detection mechanism of pristine and Ag-loaded ZnO NPs irradiated with light of 365–520 nm wavelengths. Reprinted from reference [76] with permission from Elsevier.
Sensors 21 06454 g012
Figure 13. Schematic of the chemical influence of Ag NPs on the sensing mechanism of the α-Fe2O3 sensor. Reprinted from reference [77] with permission from Elsevier.
Figure 13. Schematic of the chemical influence of Ag NPs on the sensing mechanism of the α-Fe2O3 sensor. Reprinted from reference [77] with permission from Elsevier.
Sensors 21 06454 g013
Figure 14. Schematic of the C8H10 gas-sensing mechanism of (a) the pristine and (b) Ag-loaded TiO2 sensors. Reprinted from reference [80] with permission from Elsevier.
Figure 14. Schematic of the C8H10 gas-sensing mechanism of (a) the pristine and (b) Ag-loaded TiO2 sensors. Reprinted from reference [80] with permission from Elsevier.
Sensors 21 06454 g014
Figure 15. SEM images of Ag-PPy@SiNWs with increasing resolution in front view (ac) and planar views (df). Reprinted from reference [84] with permission from Elsevier.
Figure 15. SEM images of Ag-PPy@SiNWs with increasing resolution in front view (ac) and planar views (df). Reprinted from reference [84] with permission from Elsevier.
Sensors 21 06454 g015
Figure 16. Schematic of the anti-humidity effect caused by Ag NPs. Reprinted from reference [84] with permission from Elsevier.
Figure 16. Schematic of the anti-humidity effect caused by Ag NPs. Reprinted from reference [84] with permission from Elsevier.
Sensors 21 06454 g016
Figure 17. (a) Doping of a metal on metal oxide and (b) doping effect on the gas interaction with the sensing layer. Reprinted from reference [110] with permission from ACS.
Figure 17. (a) Doping of a metal on metal oxide and (b) doping effect on the gas interaction with the sensing layer. Reprinted from reference [110] with permission from ACS.
Sensors 21 06454 g017
Figure 18. Schematic of H2S sensing by Ag-doped graphene. Energy diagram of (A) Ag-doped graphene, (B) interaction of Ag-doped graphene with H2S gas, and (C) the formation of SO2 and H2O. Reprinted from reference [113] with permission from Elsevier.
Figure 18. Schematic of H2S sensing by Ag-doped graphene. Energy diagram of (A) Ag-doped graphene, (B) interaction of Ag-doped graphene with H2S gas, and (C) the formation of SO2 and H2O. Reprinted from reference [113] with permission from Elsevier.
Sensors 21 06454 g018
Table 1. Gas-sensing properties of Ag-decorated SMO-based gas sensors for different gases.
Table 1. Gas-sensing properties of Ag-decorated SMO-based gas sensors for different gases.
Sensing MaterialsGas TypeGC (ppm)T (°C)ResponseRef.
Ag-loaded SnO2 hollow NFsCH3COCH35016042[48]
Ag-decorated TiO2 NRsCH3COCH33.82007.31 (ΔI)/I0[49]
Ag-loaded BiFe2O4 NPsCl21024072.62[50]
Ag-decorated ZnO/G nanocompositeC2H210015021.2[51]
5 wt.% Ag-loaded ZnO NPsC2H2100200≈80%[52]
Ag-loaded ZnO NRsC2H2100020027.2[54]
Ag-loaded ZnO NPsTEA100183.56043[60]
Ag-loaded Porous ZnO NPsHCHO100240180.4[65]
Ag-loaded In2O3 sunflower structureHCHO2024011.3[66]
Ag-loaded ZnO/MoS2CO100256[69]
Ag-loaded ZnO NPsCO10013024.17[70]
Ag-loaded ZnO NPsCH4500020020.15[70]
Ag/g-CNC2H5OH50401.3[72]
Ag-loaded mesoporous WO3NO217540[74]
Ag-loaded ZnO NPsNO2525 (UV light)1.545[76]
Ag-loaded Fe2O3CH3SH802572%[77]
Ag-loaded TiO2 hedgehog-like architectureC8H101003756.9[80]
Ag-loaded SnO2 yolk-shell nanostructuresH2S5350614.9[85]
Ag-loaded ZnO/rGOC2H2100020033[86]
Ag-loaded ZnO NRsC2H2100250255[87]
3.5 wt.% Ag-decorated ZnOC2H5OH5032532.5[88]
Ag-decorated ZnO/Graphene nanocompositeCH3COCH3100017571[89]
Ag-loaded 3D porous flower-like ZnO NPsC2H5OH200300268[90]
Ag-loaded In2O3 NPsHCHO50210156.9[91]
Ag-loaded SnO2-rGO nanocompositeC2H25009026 (%)[92]
Ag-loaded LaFeO3 NPsC7H8521524[93]
Ag-loaded MoO3 nanobeltsTEA10024026.58[94]
Ag-decorated TiO2 QDsNH3202525.1[95]
T = operating temperature of the sensor, GC = gas concentration, and ppm = parts per million
Table 2. Gas-sensing properties of Ag-doped SMO-based gas sensors.
Table 2. Gas-sensing properties of Ag-doped SMO-based gas sensors.
Sensing MaterialsGas TypeGC (ppm)T (°C)ResponseRef.
Ag-doped grapheneH2S5025140%[113]
Ag-doped SnO2H2S4502001.32[114]
3% Ag-doped In2O3C2H5OH1000300175[115]
Ag-doped In2O3 NPsNO2162190.1[116]
Ag-doped ZnO nanoneedlesCH3COCH310037019[117]
Ag-doped ZnO nanoellipsoidsCH3OH20037015.8[118]
Ag-doped LaFeO3 NPsHCHO10023020[119]
Ag-doped Fe2O3 NPsH2S100400220[120]
Ag-doped CaCu3Ti4O12 NPsH2S10250110[121]
Ag-doped In2O3 NPsC2H5OH150100100[122]
Ag-doped Zn2SnO4/SnO2 hollow NPsHCHO1405062.2[123]
Ag-doped SnO2 NPsH25030025[124]
Ag-doped ZnO NWsC2H5OH1300203%[125]
Ag-doped In2O3 NFsHCHO600120130[126]
Ag-doped WO3C2H5OH10030065[127]
Ag-decorated/Ag-doped ZnO columnar filmsC2H5OH100250145[128]
T = operating temperature of the sensor, GC = gas concentration, and ppm = parts per million
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Navale, S.; Shahbaz, M.; Mirzaei, A.; Kim, S.S.; Kim, H.W. Effect of Ag Addition on the Gas-Sensing Properties of Nanostructured Resistive-Based Gas Sensors: An Overview. Sensors 2021, 21, 6454. https://doi.org/10.3390/s21196454

AMA Style

Navale S, Shahbaz M, Mirzaei A, Kim SS, Kim HW. Effect of Ag Addition on the Gas-Sensing Properties of Nanostructured Resistive-Based Gas Sensors: An Overview. Sensors. 2021; 21(19):6454. https://doi.org/10.3390/s21196454

Chicago/Turabian Style

Navale, Sachin, Mehrdad Shahbaz, Ali Mirzaei, Sang Sub Kim, and Hyoun Woo Kim. 2021. "Effect of Ag Addition on the Gas-Sensing Properties of Nanostructured Resistive-Based Gas Sensors: An Overview" Sensors 21, no. 19: 6454. https://doi.org/10.3390/s21196454

APA Style

Navale, S., Shahbaz, M., Mirzaei, A., Kim, S. S., & Kim, H. W. (2021). Effect of Ag Addition on the Gas-Sensing Properties of Nanostructured Resistive-Based Gas Sensors: An Overview. Sensors, 21(19), 6454. https://doi.org/10.3390/s21196454

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