Next Article in Journal
Highly Sensitive and Selective SnO2-Gr Sensor Photoactivated for Detection of Low NO2 Concentrations at Room Temperature
Previous Article in Journal
Carbonized Plant Powder Gel for Rapid Hemostasis and Sterilization in Regard to Irregular Wounds
Previous Article in Special Issue
The Structures and Compositions Design of the Hollow Micro–Nano-Structured Metal Oxides for Environmental Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

VOC Gas Sensors Based on Zinc Stannate Nanoparticles Decorated with Silver

NT-MDT BV, 7335 Apeldoorn, The Netherlands
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(24), 1993; https://doi.org/10.3390/nano14241993
Submission received: 25 October 2024 / Revised: 27 November 2024 / Accepted: 10 December 2024 / Published: 12 December 2024
(This article belongs to the Collection Metallic and Metal Oxide Nanohybrids and Their Applications)

Abstract

:
Today, air pollution is a global environmental problem. A huge amount of explosive and combustible gas emissions that negatively affect nature and human health. Gas sensors are one of the ways to prevent this impact. Several types of sensors have been developed, but the main problem with them is the high operating temperature. This leads to a decrease in reproducibility and stability over time. The aim of the work is to synthesize zinc stannate nanoparticles, study their phase composition, and modify the structure with silver nanoparticles to improve gas-sensing characteristics. This paper presents the synthesis of zinc stannate nanoparticles by the coprecipitation method and annealing at varying temperatures. A method of decorating zinc stannate with silver nanoparticles was proposed. Using XRD, it was found that a change in the annealing temperature leads to a change in the initial phase composition of the sample. Decoration with silver nanoparticles allows for increasing the sensor response of zinc stannate layers to isopropyl alcohol by 20 times. The corresponding increase in response to ethanol is 16 times.

1. Introduction

Today, gas sensors play a crucial role in various industries, including medicine and environmental protection. A typical chemiresistive gas sensor consists of a substrate, metal electrodes, sensing material and a heating element [1]. In order to detect a signal, a resistance measurement system is required. The main characteristics of semiconductor gas sensors include selectivity to the target gas, sensitivity, response and recovery times, as well as stability over time and temperature [2].
The most common semiconductors used in gas sensors are zinc oxide (ZnO) and tin dioxide (SnO2). In previous studies [2,3], it has been found that the ternary oxide structure of zinc stannate is more effective. Zinc stannate, with its band gap of 3.6 eV and n-type conductivity, has high electron mobility and electrical conductivity, as well as a low coefficient of thermal expansion. This material has a wide range of applications [4], including gas sensors, transparent electrodes for solar cells and lithium–ion batteries.
Zinc stannate can exist in two different crystal structures [5], as shown in Figure 1. One structure is perovskite (orthorhombic symmetry) [6], ZnSnO3, and the other is inverse spinel (cubic symmetry), Zn2SnO4 [7]. At low synthesis temperatures, the perovskite structure is formed. However, this structure is unstable, and under external influences, such as a temperature increase to 500 °C, it transforms into an amorphous phase, leading to the formation of oxygen vacancies that increase the sensitivity of gas sensors [8,9,10]. As the temperature continues to rise, recrystallization takes place and stable zinc orthostannate (Zn2SnO4) with inclusions of SnO2 forms. Typically, synthesized powders contain a mixture of both perovskite and inverse spinel phases in varying proportions [11].
Currently, there are several methods for producing zinc stannate, including hydrothermal synthesis [12,13], ion exchange [14], sol–gel technique [15], and thermal precipitation [16,17]. One of these methods, coprecipitation [18], has been chosen for this study.
The main goal of further development is to reduce operating temperatures while maintaining sensor sensitivity [19]. This is important because high temperatures can lead to several problems, such as shorter sensor lifetime due to material degradation and deformation caused by differences in thermal expansion coefficients between the sensor layer and substrate. Additionally, using heating elements can complicate the device, and VOC decomposition at high temperatures can affect the stability and accuracy of the sensor.
To address these issues, noble metals can be used for decoration [20], which can help to reduce operating temperatures. The decoration with noble metal nanoparticles increases the specific surface area of the sensor, providing an additional active adsorption site [21]. This, in turn, enhances the sensor’s sensitivity and selectivity. The use of noble metals, such as palladium (Pd), in the decoration of a semiconductor allows the sensor to selectively detect hydrogen (H2) molecules [22,23]. This is because Pd has a strong affinity for hydrogen and can selectively adsorb and dissociate hydrogen molecules from other gases. Upon interaction with hydrogen, Pd undergoes a specific reaction that allows for its detection:
2Pd + xH2 → 2PdHx.
PdH has a lower work function than Pd. This means that the Schottky barrier between the two materials is lower. As a result, more electrons can be injected into the semiconductor. This leads to a larger change in resistance.
The aim of this work is to synthesize zinc stannate layers at different annealing temperatures, modify them with silver nanoparticles, and study the gas-sensing properties of the resulting structures.

2. Materials and Methods

2.1. Sample Fabrication

Coprecipitation in an aqueous solution was selected as a technique for producing zinc stannate. This method allows for the production of a nanopowder with a high degree of phase uniformity. Zinc acetate dihydrate (Zn(CH3COO)2·2H2O) and sodium stannate trihydrate (Na2SnO3·3H2O) were chosen as precursors. Using an ultrasonic bath, 0.02 M Zn(CH3COO)2·2H2O and 0.02 M Na2SnO3·3H2O were dissolved in 100 mL of distilled water and stirred at 50 °C for 20 min on a magnetic stirrer. Furthermore, the resulting solutions were mixed at the same temperature for 2 h. A white precipitate was separated using a centrifuge. Then it was washed three times with distilled water and dried at 80 °C in an air atmosphere. The resulting powder was annealed at 300 °C, 500 °C and 700 °C for two hours in a muffle furnace [18]. The corresponding samples were marked as ZTO-300, ZTO-500, ZTO-700.

2.2. Preparation of Ag Nanoparticles and Decoration of Zinc Stannate

To produce silver nanoparticles, ascorbic acid, silver nitrate and sodium citrate were used as starting materials [24]. The synthesis process involved the following steps. Aqueous solutions of silver nitrate (2 M), ascorbic acid (1.5 M) and sodium citrate (0.05 M) were prepared. A total of 10 mL of acetone and 10 mL of ethanol were mixed using a magnetic stirrer. Silver nitrate (0.5 mL), ascorbic acid (0.25 mL) and sodium citrate (0.5 mL) solutions were added to the solution of acetone and ethanol and mixed for 15 min. The precipitate was washed three times with distilled water and dried in the air at 70 °C.
A total of 50 mg of silver nanoparticles were dispersed in 50 mL of distilled water. Then, 1 g of ZTO-500 was added to the resulting suspension, which was mixed for 2 h. The precipitate was collected using a centrifuge and dried in the air at 70 °C. The corresponding sample was marked as ZTO_Ag-500.

2.3. Characterization of Sample

The crystal structure of the samples was studied by X-ray difraction (XRD) with Cu-Kα radiation (D2 Phaser, Bruker AXS, Karlsruhe, Germany). Scanning electron microscopy (SEM) was chosen to study the morphology of the samples, elemental composition was analyzed by energy dispersive spectroscopy (Tescan VEGA3LMH, Tescan, Brno, Czech Republic). X-ray photoelectron spectroscopy (XPS) was used to study the elemental composition of the samples (K-Alpha, Thermo Scientific, Waltham, MA, USA) using a monochromatic Al Kα X-ray source (1486.6 eV), and the residual pressure in the analytical chamber was approximately 4 × 10−9 mbar. Atomic force microscopy (AFM) was used to estimate the average size of silver nanoparticles (NTEGRA Aura, NT-MDT, Apeldoorn, The Netherlands).

2.4. Sensor Fabrication and Measurement

The resulting powders were deposited onto ceramic sensor platforms with gold electrodes in an interdigitated configuration (Figure S1). The distance between the electrodes was 200 μm, and their width was 200 μm. The substrates were first cleaned with acetone, isopropyl alcohol, and distilled water to remove any impurities. The powders were dispersed in 10 mL of water and deposited on the substrates by spin-coating. The coated substrates were then annealed at 300 °C (ZTO-300), 500 °C (ZTO-500) and (ZTO-500_Ag) and 700 °C (ZTO-700) for 15 min in an air atmosphere.
Gas-sensing properties were measured using a special stand (Figure 2). All measurements were performed at a voltage of 5 V.
The current through the sample was measured using a Keithley 6485 picoammeter (Keithley, Cleveland, OH, USA). An electromechanical valve, controlled by a computer, allowed switching between two flow rates. When the valve is off, dry air is supplied to the measurement chamber, and when it is on, a mixture of dry air and the target gas is introduced. The flow rates were monitored using rotameters. The vapor concentration of the target gas was determined by the saturation vapor pressure at room temperature:
C = P g a s F g a s P a t m F g a s + F a i r ,
where Pgas is the saturated vapor pressure of the bubbled liquid, Fgas is the air flow rate through the bubbler, Patm is atmospheric pressure, and Fair is the diluent air flow rate. Measurements were taken when the samples were exposed to acetone, ethanol, and isopropanol.
The response was calculated as the ratio of the sensor’s resistance in air to its resistance in the presence of a reducing gas at a given concentration:
S = R a i r R g a s ,
where Rair is the resistance of the sensor in the air atmosphere, Rgas is the resistance of the sensor in the atmosphere of the target gas.
Response and recovery times refer to the time it takes for the sensor’s resistance to change by 90% after target gas and air have been supplied, respectively.

3. Results

Figure 3 shows the results of X-ray diffraction of non-annealed powder, ZTO-300, ZTO-500 and ZTO-700. Then it can be seen that the non-annealed powder is zinc hydroxostannate ZnSn(OH)6. All of the peaks correspond to the standard cubic phase of ZnSn(OH)6 (JCPDS file No. 73-2384, space group: Pn-3 (201)) [25]. No diffraction peaks due to impurities or other phases were observed. The sharp and narrow diffraction peaks suggest the high crystallinity of the zinc hydroxostannate.
During the annealing of the precursor powder at 300 °C and 500 °C, the crystallinity has been disrupted. Analysis of XRD patterns of powders annealed at 300 °C and 500 °C showed the formation of low crystalline structures at these temperatures. The wide peaks at 34° and 60°, corresponding to the crystallographic planes (110) and (211) of face-centered perovskite ZnSnO3 (JCPDS 11-0274) [26], indicate the presence of zinc stannate nanoparticles with a small size. The XRD pattern of the ZTO-700 sample reveals a well-defined crystal structure. It was found that the cubic inverse spinel Zn2SnO4 phase (JCPDS, No. 73-1725) and tetragonal SnO2 phase (JCPDS, No. 77-0451) were formed at 700 °C [27]. In this way, a phase transition can be observed.
The results of the studies on the morphology of zinc stannate powders using SEM are shown in Figure 4. As can be seen, the particles have a cubic shape. The average particle size is the same after annealing both at 300 °C and 500 °C.
The uniform distribution of Sn, Zn and O in the synthesized nanoparticles after annealing at 500 °C is shown in Figure 5.
The sizes of the synthesized silver nanoparticles were studied using AFM, as shown in Figure 6. The average particle size was found to be 70 nm.
The EDS spectrum (Figure 7a) confirms the presence of Ag in the modified sample. At the same time, the percentages of elements determined by the spectrum are 51.5 wt.% (Sn), 24.2 wt.% (Zn), 15.4 wt.% (O), and 8.9 wt.% (Ag). The EDX scan (Figure 7b) shows the uniform distribution of silver particles over the surface of the sensor layer.
The surface elemental composition and electronic structure of the synthesized ZTO-300, ZTO-500 and ZTO-700 powders were analyzed by XPS. Distinct peaks can be seen in the XPS survey spectrum (Figure 8). These peaks correspond to the binding energies of zinc, tin, oxygen and carbon [28].
In the O1s spectra (Figure 9), two peaks can be distinguished, corresponding to energies of 530.3 eV and 531.5 eV. The more intense peak of 530.3 eV corresponds to oxygen in the crystal lattice (Olat) [29], and the second peak of 531.5 eV refers to oxygen vacancies on the surface (Ovac) [30]. The percentage ratio between different forms of oxygen on the surface of samples was analyzed. A decrease in the proportion of oxygen, corresponding to oxygen vacancies, was found as the annealing temperature increased (47% Ovac for ZTO-300, 34% Ovac for ZTO-500, 29% Ovac for ZTO-700).
The successful decoration of Ag nanoparticles on the surface of zinc stannate nanoparticles was confirmed by XPS. The study of the Ag 3d X-ray photoelectron spectrum (Figure 10) revealed the presence of two peaks with binding energies at 374.4 eV and 368.4 eV, corresponding to the 3d3/2 and 3d5/2 states of metallic silver, respectively [31,32].
The gas-sensing properties of the samples heated to 250 °C were studied and are presented in Table 1. The results showed that ZTO-300 had the highest response to isopropyl alcohol (6.26), while ZTO-500 had the highest response to acetone (7.1). ZTO-700, however, did not exhibit distinct selectivity.
Based on the results obtained, it can be concluded that the samples consisted of ZnSnO3 show higher responses compared to multiphase sample of Zn2SnO4 and SnO2. Therefore, it is likely that the Zn2SnO4 and SnO2 phases formed at 700 °C do not form a heterojunction. Another possible explanation for this is the nonuniform distribution of these phases in the powder.
A semiconductor acts as the active layer in a gas sensor. When it is exposed to different gases, the resistance of the sensor changes. This change depends on the type of semiconductor and the nature of the gas [33].
For instance, zinc stannate is an n-type semiconductor. If reducing gases such as CO, NH3, and volatile organic compounds are present, resistance decreases. Conversely, if oxidizing gases like O2 and NO2 are present, resistance increases. P-type semiconductors have the opposite reaction.
This type of resistance change can be explained by the formation of a “core–shell” structure as a result of oxygen adsorption. The electrical resistance of the sensor is due to the formation of a contact between the shells of grains (Figure 11).
The gas detection process consists of several stages, including adsorption, oxidation, and desorption. When a sensor is exposed to air, oxygen molecules are adsorbed onto its surface. Oxygen has a higher electron affinity than the work function of the semiconductor, so electrons move from the conduction band to the oxygen. This causes the surface to become negatively charged. The form of adsorbed oxygen depends on the number of free electrons and temperature [34]:
O 2 g a s O 2 a d s ,
O 2 a d s + e O 2 a d s T < 150   ° C ,
O 2 a d s + 2 e 2 O a d s 150   ° C   <   T <   400   ° C ,
O 2 a d s + 4 e 2 O 2 a d s T > 400   ° C .
Then, in the near-surface layer of the semiconductor, the concentration of free charge carriers decreases and a depleted layer is formed. This layer has increased resistance, creating a shell. As a result, free charge carriers are redistributed in the semiconductor, leading to bending of energy bands and the formation of a potential barrier. This causes the resistance of the semiconductor to increase.
The thickness of depleted layer and corresponding bending of the energy bands depends on the Debye length [35,36]:
L D = 2 ε ε 0 U s q N ,
where ε is the relative permittivity of the semiconductor, ε0 is the vacuum permittivity, Us is the height of the potential barrier, q is the elementary charge, N is the concentration of ionized donors.
When reducing gases are supplied, their molecules are adsorbed onto the surface of the semiconductor. Then redox reactions occur between the adsorbed gas particles and the adsorbed oxygen:
  V O C s   g a s V O C s a d s ,
V O C s a d s + O \ O 2 C O 2 + H 2 O g a s + e .
This reduces the resistance of the material, as well as the height of the potential barrier and thickness of the depleted layer.
A possible explanation for the observed dependencies may be related to the different mechanisms of interaction between metal oxides, acetone, and alcohols. According to XRD results, ZTO-300 and ZTO-500 samples have a similar crystal structure. However, there may be differences in the ratio of zinc, tin, and oxygen ions on the surfaces of these samples. It is known that acetone molecules can be adsorbed on metal cations [37,38], while alcohol adsorption can occur on both metal cation sites and hydroxyl groups [39,40]. The high concentration of metal cation sites on ZTO-500 surface, which serve as adsorption sites for acetone molecules, could explain why the response to acetone is higher than the response to alcohol. Conversely, the higher response of ZTO-300 sample to isopropyl alcohol may be due to a higher content of oxygen vacancies on its surface, which can adsorb hydroxyl groups from ambient air. The lower responses of the ZTO-700 sample are associated with the formation of two phases, Zn2SnO4 and SnO2.
The ZTO-500 sample, which showed a high gas-sensing response, was chosen for decoration with silver nanoparticles. After decoration, it was also tested for gas sensitivity. The results are presented in Table 2. The typical response curve of ZTO-500_Ag to isopropyl alcohol is shown in Figure 12. The effect of water vapor on the response of ZTO-500_Ag when detecting isopropyl alcohol at a concentration of 1000 ppm showed that at a relative humidity of 10%, the response was reduced by 17%. The response of the ZTO-500_Ag to isopropyl alcohol at 250 °C depends on the concentration of the target gas, as shown in Figure 13. There is a linear increase in response as the concentration increases. The stability of the baseline resistance was studied at a temperature of 250 °C for 12 h. The results are presented in Figure S2. During this time, the maximum deviation of the resistance from its initial value was 4%.
When comparing the results of gas sensitivity tests, it is can be seen that the responses to isopropyl alcohol and ethanol have increased by more than an order of magnitude. This confirms the theory of increasing the sensitivity of gas sensors through decoration with noble metals. Silver nanoparticles have acted as catalysts for chemical adsorption, as expected.
In addition to the fact that silver particles may increase sensitivity, they may also lower operating temperatures. To test this, a gas sensitivity to isopropyl alcohol was studied at a lower temperature of 150 °C (Figure 14). Under these conditions, the response was six, with a response time of 210 s and a recovery time of 450 s.
Thus, when the temperature was reduced by 100 °C, ZTO-500_Ag exhibited a response similar in magnitude to that of ZTO-500 before modification with silver nanoparticles at 250 °C.
The improvement of gas-sensitive properties is associated with two main mechanisms, chemical and electronic sensitization [41,42]. Electronic sensitization occurs when charge carriers (electrons) transfer at the metal–semiconductor interface due to the difference in work functions between the electrons in the metal and semiconductor. When electrons diffuse, they redistribute charge carriers and disrupt the electrical neutrality at the interface. With this diffusion of electrons, the charge carriers are redistributed and the electrical neutrality of the regions at the interface is disrupted.
Consequently, a contact potential difference arises, which causes the bending of the bands and the formation of a potential barrier. The height of the barrier formed depends on the contact potential difference, which is determined by the difference between the work functions of the metal and semiconductor. Figure 15 illustrates the band diagram for the metal–semiconductor junction. The work function of a noble metal is higher than that of a semiconductor, so electrons will transfer from the conduction band of a semiconductor to the metal. As a result, a Schottky barrier forms, which increases the electron-depleted layer. This can prevent the recombination of electrons and holes, and cause a higher change in resistance when the sensor is exposed to reducing gases. As a result, the response of the sensor increases.
In chemical sensitization (Figure 16) noble metals act as catalysts for chemical adsorption, reducing the adsorption energy of gases and contributing to the dissociation of oxygen molecules into more reactive oxygen atoms. These formed ions then transport from the noble metal to the surface of the semiconductor, where they react with the target gas. When the target gas is present, a larger number of oxygen ions enter redox reactions, releasing a greater number of electrons. Consequently, the response of the sensor increases.
The observed decrease in the response of ZTO-500 to acetone after modification with silver nanoparticles can be caused by the influence of chemical sensitization. As a result, the concentration of ionosorbed oxygen ions, occupying free metal cation sites on the surface, increases. The content of adsorption sites for acetone molecules decreases; therefore, the response to acetone also decreases.
At the same time, as discussed earlier, the adsorption of alcohol molecules can occur with the participation of surface hydroxyl groups, so the concentration of adsorbed alcohol molecules remains high. An increase in the concentration of ionosorbed oxygen leads to an increase in the number of electrons formed as a result of the reaction (10) and an increase in the responses to isopropyl alcohol and ethanol.
A comparison of the gas-sensing characteristics achieved in this study with the results of other researchers on the development of silver-modified sensors based on metal oxide materials is presented in Table 3.
The table shows silver modification is an effective approach for developing highly efficient sensors. The findings from this study demonstrate the ability to achieve response and recovery at temperatures significantly lower than those reported in most previous studies.

4. Conclusions

Three structures of zinc stannate were obtained using the coprecipitation method in an aqueous solution. When studying their gas-sensitive properties, it was found that ZTO-300 was selective to isopropyl alcohol vapors, ZTO-500 to acetone vapors and ZTO-700 showed no selectivity. XRD analysis found that ZT0-300 and ZTO-500 had an amorphous-like structure, while ZTO-700 had a crystalline structure of Zn2SnO4 and SnO2. SEM analysis revealed that all samples had a cubic structure, and the average particle size was 180 nm, regardless of annealing temperature. A sample decorated with silver nanoparticles showed a 20-fold improvement in gas sensitivity to isopropyl alcohol, and reducing operating temperatures to 150 °C was possible by decorating with silver nanoparticles. Using AFM, it was found the average size of the silver nanoparticles was 70 nm.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14241993/s1.

Author Contributions

Conceptualization, S.S.N. and V.A.M.; methodology, S.S.N. and Z.V.S.; investigation, A.A.R., S.S.B., C.D.B., Z.V.S. and I.A.N.; resources, D.A.K.; data curation, S.S.N., Z.V.S., V.A.M. and D.A.K.; writing—original draft preparation, A.A.R., S.S.B. and S.S.N.; writing—review and editing, C.D.B., Z.V.S., V.A.M., I.A.N. and D.A.K.; visualization, A.A.R., S.S.B. and C.D.B.; supervision, S.S.N., Z.V.S., V.A.M. and D.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by NT-MDT BV.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

All authors are employed by the company NT-MDT BV. The authors declare that this study received funding from NT-MDT BV. The funder was not involved in the study design, collection, analysis, interpretation of data, the writing of this article or the decision to submit it for publication.

References

  1. Korotcenkov, G. Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 2: Gas Sensors and Their Advantages and Limitations. Nanomaterials 2021, 11, 1555. [Google Scholar] [CrossRef] [PubMed]
  2. Hashtroudi, H.; Mackinnon, I.D.R.; Shafiei, M. Emerging 2D hybrid nanomaterials: Towards enhanced sensitive and selective conductometric gas sensors at room temperature. J. Mater. Chem. C 2020, 8, 13108–13126. [Google Scholar] [CrossRef]
  3. Wawrzyniak, J. Advancements in Improving Selectivity of Metal Oxide Semiconductor Gas Sensors Opening New Perspectives for Their Application in Food Industry. Sensors 2023, 23, 9548. [Google Scholar] [CrossRef] [PubMed]
  4. Petrov, V.V.; Sysoev, V.V.; Starnikova, A.P.; Volkova, M.G.; Kalazhokov, Z.K.; Storozhenko, V.Y.; Khubezhov, S.A.; Bayan, E.M. Synthesis, Characterization and Gas Sensing Study of ZnO-SnO2 Nanocomposite Thin Films. Chemosensors 2021, 9, 124. [Google Scholar] [CrossRef]
  5. Nalimova, S.S.; Maksimov, A.I.; Matyushkin, L.B.; Moshnikov, V.A. Current State of Studies on Synthesis and Application of Zinc Stannate (Review). Glass Phys. Chem. 2019, 45, 251–260. [Google Scholar] [CrossRef]
  6. Anitha, A.; Ponnusamy, V. Optical and electrochemical studies on single-phase ZnSnO3 nanostructures—A photosensitive approach. Surf. Interfaces 2024, 51, 104747. [Google Scholar] [CrossRef]
  7. Hanh, N.H.; Van Duy, L.; Hung, C.M.; Van Duy, N.; Heo, Y.-W.; Van Hieu, N.; Hoa, N.D. VOC gas sensor based on hollow cubic assembled nanocrystal Zn2SnO4 for breath analysis. Sens. Actuators A 2020, 302, 111834. [Google Scholar] [CrossRef]
  8. Shomakhov, Z.V.; Nalimova, S.S.; Rybina, A.A.; Buzovkin, S.S.; Kalazhokov, Z.; Moshnikov, V.A. Improving the sensor characteristics of binary and ternary oxide nanosystems. Phys. Chem. Asp. Study Clust. Nanostruct. Nanomater. 2023, 15, 879–887. [Google Scholar] [CrossRef]
  9. Xiong, Y.; Lin, Y.; Wang, X.; Zhao, Y.; Tian, J. Defect engineering on SnO2 nanomaterials for enhanced gas sensing performances. Adv. Powder Mater. 2022, 1, 100033. [Google Scholar] [CrossRef]
  10. Zhang, J.; Li, J. The Oxygen Vacancy Defect of ZnO/NiO Nanomaterials Improves Photocatalytic Performance and Ammonia Sensing Performance. Nanomaterials 2022, 12, 433. [Google Scholar] [CrossRef]
  11. Bora, T.; Al-Hinai, M.H.; Al-Hinai, A.T.; Dutta, J. Phase Transformation of Metastable ZnSnO3 Upon Thermal Decomposition by In-Situ Temperature-Dependent Raman Spectroscopy. J. Am. Ceram. Soc. 2015, 98, 4044–4049. [Google Scholar] [CrossRef]
  12. Zeng, Y.; Zhang, T.; Fan, H.; Lu, G.; Kang, M. Synthesis and gas-sensing properties of ZnSnO3 cubic nanocages and nanoskeletons. Sens. Actuators B 2009, 143, 449–453. [Google Scholar] [CrossRef]
  13. Xu, J.; Jia, X.; Lou, X.; Shen, J. One-step hydrothermal synthesis and gas sensing property of ZnSnO3 microparticles. Solid-State Electron. 2006, 50, 504–507. [Google Scholar] [CrossRef]
  14. Kovacheva, D.; Petrov, K. Preparation of crystalline ZnSnO3 from Li2SnO3 by low-temperature ion exchange. Solid State Ion. 1998, 109, 327–332. [Google Scholar] [CrossRef]
  15. Ibrahim, D.M.; Gaber, A.A.; Reda, A.E.; Abdel Aziz, D.A.; Ajiba, N.A. Structural, optical, and dielectric properties of sol-gel derived perovskite ZnSnO3 nanomaterials. J. Sol-Gel Sci. Technol. 2024, 112, 703–714. [Google Scholar] [CrossRef]
  16. Jie, J.; Wang, G.; Han, X.; Fang, J.; Yu, Q.; Liao, Y.; Xu, B.; Wang, Q.; Hou, J. Growth of Ternary Oxide Nanowires by Gold-Catalyzed Vapor-Phase Evaporation. J. Phys. Chem. B 2004, 108, 8249–8253. [Google Scholar] [CrossRef]
  17. Wang, L.; Zhang, X.; Liao, X.; Yang, W. A simple method to synthesize single-crystalline Zn2SnO4 (ZTO) nanowires and their photoluminescence properties. Nanotechnology 2005, 16, 2928–2931. [Google Scholar] [CrossRef]
  18. Ge, Q.; Liu, C.; Zhao, Y.; Wang, N.; Zhang, X.; Feng, C.; Zhang, S.; Wang, H.; Jiang, W.; Liu, S.; et al. Phase evolution in preparing ZnSnO3 powders by precipitation method. Appl. Phys. A 2021, 127, 89. [Google Scholar] [CrossRef]
  19. Chang, C.-J.; Lin, C.-Y.; Chen, J.-K.; Hsu, M.-H. Ce-doped ZnO nanorods based low operation temperature NO2 gas sensors. Ceram. Int. 2014, 40, 10867–10875. [Google Scholar] [CrossRef]
  20. Zhu, L.-Y.; Ou, L.-X.; Mao, L.-W.; Wu, X.-Y.; Liu, Y.-P.; Lu, H.-L. Advances in Noble Metal-Decorated Metal Oxide Nanomaterials for Chemiresistive Gas Sensors: Overview. Nano-Micro Lett. 2023, 15, 89. [Google Scholar] [CrossRef]
  21. Lai, H.-Y.; Chen, C.-H. Highly sensitive room-temperature CO gas sensors: Pt and Pd nanoparticle-decorated In2O3 flower-like nanobundles. J. Mater. Chem. 2012, 22, 13204. [Google Scholar] [CrossRef]
  22. Meng, X.; Bi, M.; Xiao, Q.; Gao, W. Ultra-fast response and highly selectivity hydrogen gas sensor based on Pd/SnO2 nanoparticles. Int. J. Hydrogen Energy 2022, 47, 3157–3169. [Google Scholar] [CrossRef]
  23. Kafil, V.; Sreenan, B.; Hadj-Nacer, M.; Wang, Y.; Yoon, J.; Greiner, M.; Chu, P.; Wang, X.; Fadali, M.S.; Zhu, X. Review of noble metal and metal-oxide-semiconductor based chemiresistive hydrogen sensors. Sens. Actuators A 2024, 373, 115440. [Google Scholar] [CrossRef]
  24. Nogami, K.; Kishimoto, K.; Hashimoto, Y.; Watanabe, H.; Hishii, Y.; Ma, Q.; Niki, T.; Kotani, T.; Kiwa, T.; Shoji, S.; et al. Self-growth of silver tree-like fractal structures with different geometries. Appl. Phys. A 2022, 128, 860. [Google Scholar] [CrossRef]
  25. He, Q.; Zi, J.; Huang, B.; Yan, L.; Fa, W.; Li, D.; Zhang, Y.; Gao, Y.; Zheng, Z. Controlled growth and thermal decomposition of well-dispersed and uniform ZnSn(OH)6 submicrocubes. J. Alloys Compd. 2014, 607, 193–197. [Google Scholar] [CrossRef]
  26. Fan, H.; Zeng, Y.; Xu, X.; Lv, N.; Zhang, T. Hydrothermal synthesis of hollow ZnSnO3 microspheres and sensing properties toward butane. Sens. Actuators B 2011, 153, 170–175. [Google Scholar] [CrossRef]
  27. Li, B.; Luo, L.; Xiao, T.; Hu, X.; Lu, L.; Wang, J.; Tang, Y. Zn2SnO4–SnO2 heterojunction nanocomposites for dye-sensitized solar cells. J. Alloys Compd. 2011, 509, 2186–2191. [Google Scholar] [CrossRef]
  28. Nalimova, S.S.; Shomakhov, Z.V.; Moshnikov, V.A.; Bobkov, A.A.; Ryabko, A.A.; Kalazhokov, Z.K. An X-ray Photoelectron Spectroscopy Study of Zinc Stannate Layer Formation. Tech. Phys. 2020, 65, 1087–1090. [Google Scholar] [CrossRef]
  29. Sá, B.S.; Zito, C.A.; Perfecto, T.M.; Volanti, D.P. Porous ZnSnO3 nanocubes as a triethylamine sensor. Sens. Actuators B 2021, 338, 129869. [Google Scholar] [CrossRef]
  30. Wang, Y.-C.; Wu, J.M. Effect of Controlled Oxygen Vacancy on H2-Production through the Piezocatalysis and Piezophototronics of Ferroelectric R3C ZnSnO3 Nanowires. Adv. Funct. Mater. 2020, 30, 1907619. [Google Scholar] [CrossRef]
  31. Naberezhnyi, D.; Rumyantseva, M.; Filatova, D.; Batuk, M.; Hadermann, J.; Baranchikov, A.; Khmelevsky, N.; Aksenenko, A.; Konstantinova, E.; Gaskov, A. Effects of Ag Additive in Low Temperature CO Detection with In2O3 Based Gas Sensors. Nanomaterials 2018, 8, 801. [Google Scholar] [CrossRef] [PubMed]
  32. Barathy, T.R.; Yadav, P.V.K.; Mondal, A.; Ramakrishnan, K.; Jarugala, J.; Liu, C.; Reddy, Y.A.K. Enhanced response of WO3 thin film through Ag loading towards room temperature hydrogen gas sensor. Chemosphere 2024, 353, 141545. [Google Scholar] [CrossRef]
  33. Korotcenkov, G.; Cho, B.K. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B 2017, 244, 182–210. [Google Scholar] [CrossRef]
  34. Ma, S.; Shen, L.; Ma, S.; Wen, J.; Xu, J. Emerging zinc stannate and its application in volatile organic compounds sensing. Coord. Chem. Rev. 2023, 490, 215217. [Google Scholar] [CrossRef]
  35. Park, S.; Ko, H.; Kim, S.; Lee, C. Role of the interfaces in multiple networked one-dimensional core-shell nanostructured gas sensors. ACS Appl. Mater. Interfaces 2014, 6, 9595–9600. [Google Scholar] [CrossRef]
  36. Zhu, X.; Zhang, X.; Chang, X.; Li, J.; Pan, L.; Jiang, Y.; Gao, W.; Gao, C.; Sun, S. Metal-organic framework-derived porous SnO2 nanosheets with grain sizes comparable to Debye length for formaldehyde detection with high response and low detection limit. Sens. Actuators B 2021, 347, 130599. [Google Scholar] [CrossRef]
  37. Marikutsa, A.; Rumyantseva, M.; Konstantinova, E.A.; Gaskov, A. The Key Role of Active Sites in the Development of Selective Metal Oxide Sensor Materials. Sensors 2021, 21, 2554. [Google Scholar] [CrossRef]
  38. Zhou, T.; Sang, Y.; Sun, Y.; Wu, C.; Wang, X.; Tang, X.; Zhang, T.; Wang, H.; Xie, C.; Zeng, D. Gas Adsorption at Metal Sites for Enhancing Gas Sensing Performance of ZnO@ZIF-71 Nanorod Arrays. Langmuir 2019, 35, 3248. [Google Scholar] [CrossRef]
  39. Cheng, I.-K.; Wang, J.-H.; Tsai, C.-Y.; Chen, Y.-S.; Pan, F.-M. Correlation of surface processes with characteristic sensing responses of PdO thin films to ethanol. Appl. Surf. Sci. 2019, 473, 589. [Google Scholar] [CrossRef]
  40. Chiang, H.; Bhan, A. Catalytic consequences of hydroxyl group location on the rate and mechanism of parallel dehydration reactions of ethanol over acidic zeolites. J. Catal. 2010, 271, 251. [Google Scholar] [CrossRef]
  41. Barbosa, M.S.; Suman, P.H.; Kim, J.J.; Tuller, H.L.; Orlandi, M.O. Investigation of electronic and chemical sensitization effects promoted by Pt and Pd nanoparticles on single-crystalline SnO nanobelt-based gas sensors. Sens. Actuators B 2019, 301, 127055. [Google Scholar] [CrossRef]
  42. Cheng, P.; Wang, Y.; Wang, C.; Ma, J.; Xu, L.; Lv, C.; Sun, Y. Investigation of doping effects of different noble metals for ethanol gas sensors based on mesoporous In2O3. Nanotechnology 2021, 32, 305503. [Google Scholar] [CrossRef] [PubMed]
  43. Agarwal, S.; Kumar, S.; Agrawal, H.; Moinuddin, M.G.; Kumar, M.; Sharma, S.K.; Awasthi, K. An efficient hydrogen gas sensor based on hierarchical Ag/ZnO hollow microstructures. Sens. Actuators B 2021, 346, 130510. [Google Scholar] [CrossRef]
  44. Wang, S.; Jia, F.; Wang, X.; Hu, L.; Sun, Y.; Yin, G.; Zhou, T.; Feng, Z.; Kumar, P.; Liu, B. Fabrication of ZnO Nanoparticles Modified by Uniformly Dispersed Ag Nanoparticles: Enhancement of Gas Sensing Performance. ACS Omega 2020, 5, 5209. [Google Scholar] [CrossRef]
  45. 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]
  46. Zhang, C.; Huan, Y.; Li, Y.; Luo, Y.; Debliquy, M. Low concentration isopropanol gas sensing properties of Ag nanoparticles decorated In2O3 hollow spheres. J. Adv. Ceram. 2022, 11, 379. [Google Scholar] [CrossRef]
  47. Kim, M.Y.; Hwang, J.Y.; Mirzaei, A.; Choi, S.W.; Kim, S.I.; Kim, H.S.; Kim, S.J.; Roh, J.W.; Choi, M.S.; Lee, K.H.; et al. NO2 Gas Sensing Properties of Ag-Functionalized Porous ZnO Sheets. Adsorpt. Sci. Technol. 2023, 2023, 9021169. [Google Scholar] [CrossRef]
  48. Sima, Z.; Song, P.; Wang, Q. Ag nanoparticles decorated ZnSnO3 hollow cubes for enhanced formaldehyde sensing performance at low temperature. Appl. Surf. Sci. 2023, 614, 156215. [Google Scholar] [CrossRef]
  49. Kamble, C.; Narwade, S.; Mane, R. Detection of acetylene (C2H2) gas using Ag-modified ZnO/GO nanorods prepared by a hydrothermal synthesis. Mater. Sci. Semicond. Proc. 2023, 153, 107145. [Google Scholar] [CrossRef]
  50. Guo, L.; Liang, H.; An, D.; Yang, H. One-step hydrothermal synthesis of uniform Ag-doped SnO2 nanoparticles for highly sensitive ethanol sensing. Physica E 2023, 151, 115717. [Google Scholar] [CrossRef]
  51. Wang, Z.; Haidry, A.A.; Xie, L.; Zavabeti, A.; Li, Z.; Yin, W.; Fomekong, R.L.; Saruhan, B. Acetone sensing applications of Ag modified TiO2 porous nanoparticles synthesized via facile hydrothermal method. Appl. Surf. Sci. 2020, 533, 147383. [Google Scholar] [CrossRef]
  52. Anusha; Ani, A.; Poornesh, P.; Antony, A.; Bhaghyesh; Shchetinin, I.V.; Nagaraja, K.K.; Chattopadhyay, S.; Vinayakumar, K.B. Impact of Ag on the Limit of Detection towards NH3-Sensing in Spray-Coated WO3 Thin-Films. Sensors 2022, 22, 2033. [Google Scholar] [CrossRef]
Figure 1. Crystal structures of ZnSnO3 (a) and Zn2SnO4 (b).
Figure 1. Crystal structures of ZnSnO3 (a) and Zn2SnO4 (b).
Nanomaterials 14 01993 g001
Figure 2. The scheme of the measuring stand: 1—compressor, 2—dehumidifier, 3, 4—rotameters, 5—bubbler, 6—electromechanical valve, 7—measuring cell, 8—controller, 9, 10—picoammeter and voltage source, 11—personal computer.
Figure 2. The scheme of the measuring stand: 1—compressor, 2—dehumidifier, 3, 4—rotameters, 5—bubbler, 6—electromechanical valve, 7—measuring cell, 8—controller, 9, 10—picoammeter and voltage source, 11—personal computer.
Nanomaterials 14 01993 g002
Figure 3. XRD patterns of: (a) non-annealed powder; (b) ZTO-300; (c) ZTO-500; (d) ZTO-700.
Figure 3. XRD patterns of: (a) non-annealed powder; (b) ZTO-300; (c) ZTO-500; (d) ZTO-700.
Nanomaterials 14 01993 g003
Figure 4. SEM images of: (a) ZTO-300; (b) ZTO-500.
Figure 4. SEM images of: (a) ZTO-300; (b) ZTO-500.
Nanomaterials 14 01993 g004
Figure 5. Elemental mapping for ZTO-500.
Figure 5. Elemental mapping for ZTO-500.
Nanomaterials 14 01993 g005
Figure 6. AFM image of silver nanoparticles.
Figure 6. AFM image of silver nanoparticles.
Nanomaterials 14 01993 g006
Figure 7. (a) EDS spectrum of ZTO-500_Ag; (b) Elemental mapping of Ag in ZTO-500_Ag.
Figure 7. (a) EDS spectrum of ZTO-500_Ag; (b) Elemental mapping of Ag in ZTO-500_Ag.
Nanomaterials 14 01993 g007
Figure 8. Survey X-ray photoelectron spectrum of ZTO-500.
Figure 8. Survey X-ray photoelectron spectrum of ZTO-500.
Nanomaterials 14 01993 g008
Figure 9. X-ray photoelectron spectra of O1s level: (a) ZTO-300; (b) ZTO-500; (c) ZTO-700.
Figure 9. X-ray photoelectron spectra of O1s level: (a) ZTO-300; (b) ZTO-500; (c) ZTO-700.
Nanomaterials 14 01993 g009
Figure 10. X-ray photoelectron spectra of Ag3d level in ZTO-500_Ag.
Figure 10. X-ray photoelectron spectra of Ag3d level in ZTO-500_Ag.
Nanomaterials 14 01993 g010
Figure 11. The mechanism of gas sensitivity in zinc stannate.
Figure 11. The mechanism of gas sensitivity in zinc stannate.
Nanomaterials 14 01993 g011
Figure 12. ZTO-500_Ag and ZTO-500 response curves to isopropyl alcohol with a concentration of 1000 ppm at 250 °C.
Figure 12. ZTO-500_Ag and ZTO-500 response curves to isopropyl alcohol with a concentration of 1000 ppm at 250 °C.
Nanomaterials 14 01993 g012
Figure 13. Concentration dependence of ZTO-500_Ag response to isopropyl alcohol at 250 °C.
Figure 13. Concentration dependence of ZTO-500_Ag response to isopropyl alcohol at 250 °C.
Nanomaterials 14 01993 g013
Figure 14. ZTO-500_Ag response curve to isopropyl alcohol with a concentration of 1000 ppm at 150 °C.
Figure 14. ZTO-500_Ag response curve to isopropyl alcohol with a concentration of 1000 ppm at 150 °C.
Nanomaterials 14 01993 g014
Figure 15. Band diagram of the metal–semiconductor interface: (a) in air; (b) in the presence of reducing gas [20].
Figure 15. Band diagram of the metal–semiconductor interface: (a) in air; (b) in the presence of reducing gas [20].
Nanomaterials 14 01993 g015
Figure 16. Chemical sensitization mechanism [20].
Figure 16. Chemical sensitization mechanism [20].
Nanomaterials 14 01993 g016
Table 1. Experimental data on the gas-sensing properties of ZTO for different VOCs at a concentration of 1000 ppm.
Table 1. Experimental data on the gas-sensing properties of ZTO for different VOCs at a concentration of 1000 ppm.
StructureGasResponseResponse/Recovery Time, s
ZTO-300isopropyl alcohol6.26143/313
acetone1.56143/88
ethanol1.5961/110
ZTO-500isopropyl alcohol3108/254
acetone7.118/282
ethanol2.234/235
ZTO-700isopropyl alcohol1.4862/60
acetone2.37106/159
ethanol1.75179/120
Table 2. Comparison of the results from measuring the gas sensitivity of ZTO-500 before and after decoration with silver nanoparticles.
Table 2. Comparison of the results from measuring the gas sensitivity of ZTO-500 before and after decoration with silver nanoparticles.
Isopropyl Alcohol (1000 ppm)Acetone (1000 ppm)Ethanol (1000 ppm)
ZTO-500ZTO_Ag-500ZTO-500ZTO_Ag-500ZTO-500ZTO_Ag-500
Response3627.14.222.235.6
Response time, s17758602203473
Recovery time, s254457282238235442
Table 3. Gas-sensing performances of the developed sensor and reported in the literature.
Table 3. Gas-sensing performances of the developed sensor and reported in the literature.
Sensor MaterialTarget Gas (Concentration)Response, Ra/RgTemperature, °CReference
Ag-ZnOH2, 300 ppm5.79250[43]
Ag-ZnOEthanol, 100 ppm148320[44]
Ag-ZnO/In2O3HCHO, 100 ppm186260[45]
Ag-In2O3Isopropanol, 5 ppm5.2300[46]
Ag-ZnONO2, 10 ppm17.18200[47]
Ag-ZnSnO3HCHO, 50 ppm26.7100[48]
Ag–ZnO/GOC2H2, 100 ppm2.87250[49]
Ag/SnO2Ethanol, 50 ppm135180[50]
Ag-TiO2Acetone, 100 ppm13.9275[51]
Ag-WO3NH3, 5 ppm3.29200[52]
Ag-ZnSnO3Isopropanol, 1000 ppm6150This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nalimova, S.S.; Shomakhov, Z.V.; Kozodaev, D.A.; Rybina, A.A.; Buzovkin, S.S.; Bui, C.D.; Novikov, I.A.; Moshnikov, V.A. VOC Gas Sensors Based on Zinc Stannate Nanoparticles Decorated with Silver. Nanomaterials 2024, 14, 1993. https://doi.org/10.3390/nano14241993

AMA Style

Nalimova SS, Shomakhov ZV, Kozodaev DA, Rybina AA, Buzovkin SS, Bui CD, Novikov IA, Moshnikov VA. VOC Gas Sensors Based on Zinc Stannate Nanoparticles Decorated with Silver. Nanomaterials. 2024; 14(24):1993. https://doi.org/10.3390/nano14241993

Chicago/Turabian Style

Nalimova, Svetlana S., Zamir V. Shomakhov, Dmitry A. Kozodaev, Arina A. Rybina, Sergey S. Buzovkin, Cong D. Bui, Ivan A. Novikov, and Vyacheslav A. Moshnikov. 2024. "VOC Gas Sensors Based on Zinc Stannate Nanoparticles Decorated with Silver" Nanomaterials 14, no. 24: 1993. https://doi.org/10.3390/nano14241993

APA Style

Nalimova, S. S., Shomakhov, Z. V., Kozodaev, D. A., Rybina, A. A., Buzovkin, S. S., Bui, C. D., Novikov, I. A., & Moshnikov, V. A. (2024). VOC Gas Sensors Based on Zinc Stannate Nanoparticles Decorated with Silver. Nanomaterials, 14(24), 1993. https://doi.org/10.3390/nano14241993

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