Next Article in Journal
Productivity Equation of Fractured Vertical Well with Gas–Water Co-Production in High-Water-Cut Tight Sandstone Gas Reservoir
Previous Article in Journal
Optimization of the Flashing Processes in Mineral Metallurgy Based on CFD Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances of Nano-Enabled ZnFe2O4 Based-Gas Sensors for VOC Detection and Their Potential Applications: A Review

by
Murendeni I. Nemufulwi
1,*,
Hendrik C. Swart
2 and
Gugu H. Mhlongo
1,2,*
1
Centre for Nanostructures and Advanced Materials (CeNAM), DSI-CSIR Nanotechnology Innovation Centre, Council for Scientific and Industrial Research, Pretoria ZA0001, South Africa
2
Department of Physics, University of the Free State, Bloemfontein ZA9300, South Africa
*
Authors to whom correspondence should be addressed.
Processes 2023, 11(11), 3122; https://doi.org/10.3390/pr11113122
Submission received: 22 September 2023 / Revised: 26 October 2023 / Accepted: 28 October 2023 / Published: 31 October 2023

Abstract

:
The demand for reliable gas sensing technologies in chemical, manufacturing, environmental, and occupational sites has increased in the last few decades following the global volatile gas sensor market, which is expected to grow further beyond 2025. Currently, several types of sensors have been employed for applications in different fields. Optical sensors are widely implemented in mining and environmental monitoring. Conventional food testing methods are utilized for the detection of any chemical or microbial agent in the food industry. Although robust and sensitive, most sensing technologies are expensive, labor-intensive, and necessitate the use of time-consuming gas sampling pretreatment steps, and these issues impede the achievement of quick, simple detection, portable, and cost-effective gas monitoring. For this reason, researchers around the world are investigating the possibility of using gas sensors as a promising technology that has the potential to alleviate industrial safety concerns. As a highly sensitive semiconducting metal oxide, gas sensors based on ZnFe2O4 have the potential to ensure environmental and occupational safety in real time. This review introduces and highlights recent developments in ZnFe2O4 gas sensors for application in different fields. The challenges limiting the wide application of the ZnFe2O4 sensor are outlined. Furthermore, this review discusses the common strategies adopted to improve the sensing properties of ZnFe2O4 for gas detection. Finally, future perspectives on further improvements of ZnFe2O4 sensing properties are discussed, and integration of ZnFe2O4 sensors into electronic noses to tackle the selectivity issue and how they can feature on the Internet of Things is outlined.

1. Introduction

The detection of different target gases using gas sensors has drawn a lot of attention in recent years. This is mainly due to certain industrial activities in the mining and energy sectors that release various toxic, flammable, and explosive gases (e.g., CH4, H2S, CO2), thus threatening human health and the surrounding environment [1]. As such, environmental monitoring through the detection of gases by reliable sensors remains crucially important. The detection of naturally emitted gases is essential to process and quality control. Furthermore, chemical reactions with gaseous emissions in food production and aging have posed a sharp rise in the demand for gas sensors to analyze complex mixtures. A survey conducted by Grand View Research in 2018 showed that more than 20% of the market share of Europe’s Volatile Organic Compound (VOC) gas sensor, by application, was in the food and beverage sector [2]. Furthermore, the global market size for VOC gas sensors was valued at USD 141.7 million in 2018 and is expected to increase by 4.0% in 2025 [2]. This creates opportunities for sensors targeting VOCs.
In general, sensors are devices that transform a chemical reaction of an analyte or a form of physical property of a system into an analytically useful signal. They operate mainly under two basic functional units. A receptor part is the unit that transforms chemical information into measurable responses, and a transducer takes the response and converts it to a useful signal [3]. The receptor part of sensors operates under three principles, including physical, chemical, and biochemical principles. These principles describe the operating procedure with which the sensors are classified. The physical principle is based on the measurement of temperature, mass, electrical properties, absorbance, etc. On the other hand, the chemical sensor is based on chemical reactions with an analyte to give a certain signal, while the biochemical one gives out a signal due to a biochemical process and is commonly known as a biosensor [4,5]. For this reason, sensors have been used in different fields, including environmental control and monitoring [6], biomedical [7], operation and maintenance in civil construction [8], the food sector [4], and other production industries [9,10,11]. For instance, optical sensors have been widely implemented in the mining and environmental monitoring sectors [12]. In the food industry, conventional methods have been used to detect and analyze several VOCs released by food. These include gas chromatographic techniques, high-performance liquid chromatography (HPLC), polymerase chain reaction (PCR), as well as immunological methods, i.e., enzyme-linked immunosorbent assays (ELISAs) [13,14,15,16,17]. Although robust and sensitive, most sensing technologies are expensive, labor-intensive, and require time-consuming gas sampling pretreatment steps, and these issues impede the achievement of quick, simple detection, portable, and cost-effective gas monitoring. For instance, Mustafa et al. [18] described a novel approach to designing a calorimetric-based biosensor to measure hypoxanthine levels in degraded fish. Carvalho et al. [19] developed a texture sensor capable of differentiating fruits according to skin texture and determining fruit ripeness. However, several sensors come with disadvantages. To obtain more accurate results using the textural sensor developed by Carvalho et al. [19], the fruits should be rubbed at different locations, as textures differ with location, requiring labor and time. In contrast, gas sensors based on variations in electrical properties have several advantages over other sensors. They offer long-term detection at low concentrations without the need for complex measurement techniques. Of particular interest, SMOs have the unique property of reversible gas interaction with pre-adsorbed ambient oxygen, which significantly changes the electrical resistance measured as the output signal. Hence, SMO-based gas sensors have attracted a lot of attention, considering that they show great potential for gas-sensing applications. This type of sensor offers low cost, easy fabrication, simplicity of measurement, and enhanced sensing performance [20]. Moreover, SMO-based gas sensors offer a wide range of applicability. Different SMO materials, including binary oxides (e.g., ZnO, TiO2, SnO2, CuO, etc.) and ternary oxides (e.g., LaCeCoO3, NiFe2O4, etc.), are available for choice to be used in the design and fabrication of gas sensors as active sensing layers. Among them, spinel-type ferrites with an AB2O4 formula are sensitive to a wide range of gases, such as H2S, VOCs, NH3, NO2, and H2O, giving them application possibilities in a wide range of fields. This is owing to their unique electronic and magnetic structures that can be tuned to suit a particular application by the redistribution of cations in the structure. However, a variation in the stoichiometry can cause instability and gas sensing limitations. Several strategies have been developed and applied to improve spinel-ferrite-based sensors to maintain their possibilities for a wide range of applications. Among several spinel ferrite materials, zinc ferrite (ZnFe2O4) has been extensively studied for its potential in gas sensing owing to its excellent chemical stability, low coercivity, and high electromagnetic performance [21,22,23,24,25]. A comparison of commonly used and investigated SMOs for VOC detection shows that ZnFe2O4 and its composites are promising materials for manufacturing as a VOC gas sensor (Table 1). Moreover, the selective properties of ZnFe2O4 sensors towards VOCs make them a good candidate for gas sensing applications in the food sector, occupational sites, chemical, and production industries.
Although ZnFe2O4 sensors show high sensitivity and selectivity to several groups of target gases, their selectivity to specific gas species and high operating temperatures remain a challenge that hinders their wide-range application. Researchers have attempted to improve the selectivity of ZnFe2O4 towards a particular VOC through surface decoration with metal nanoparticles, including Ag, Au, and Pt, and the design of novel hybrid nanocomposites (i.e., ZnO/ZnFe2O4) [26,27]. Furthermore, to tackle the selectivity problems of SMO-based sensors, Industry 4.0 technologies are implemented for more reliable and sustainable devices that automate and modernize sensor applications for all industrial needs.
Therefore, in this review, the recent advances and applicability of ZnFe2O4 gas sensors have been explored. We start by briefly explaining the structure and describing the accepted gas sensing principle of a spinel-type ZnFe2O4 based sensor. We further outline the challenges that hinder practical applications of ZnFe2O4 based sensors and indicate the important gas sensing properties in various industries. Different strategies used to develop and overcome the limitations of practical applications of SMO-based gas sensors in general are also discussed. Finally, different industrial applications of ZnFe2O4 based sensors and machine learning techniques that can be adopted to improve the selectivity and applicability of sensors are described.
Table 1. Gas sensors employed for VOC detection.
Table 1. Gas sensors employed for VOC detection.
MaterialTarget GasResponse (Ra/Rg)/Concentration (ppm)Limit of Detection (ppm)Response/Recovery Times (s)Operating Temperature (°C)Ref
ZnFe2O4Acetone192/90-3/21120[28]
ZnO/ZnFe2O4Acetone92.9/900.187.7/17.3120[29]
ZnFe2O4/SnO2Acetone120/1000.1030/197210[30]
Sn-ZnO/ZnFe2O4Triethylamine28.1/10<0.2-/7270[31]
Au-ZnOIsoprene1371/10.006-350[32]
ZnO-CuOAcetone11.1/10.04-200[33]
Pd-SnO2Ethylene11.1/1000.051/-250[34]
SnO2-CuOEthanol8/100-4/10320[35]
SnO2Ethanol59.6/40-105/100150[36]
CuO/WO3Xylene6.3/50-5.5/16260[37]
TiO2Acetone12/100-3/421320[38]
CoTiO3/TiO2Benzene33.4/500.149/9RT[39]
TiO2Ethanol2.2/100--RT[40]

2. Gas Sensors

2.1. Semiconductor Metal Oxides (SMO) Based Gas Sensors

An SMO-based sensor has a sensing element with a planar arrangement, consisting of a sensitive layer deposited over a substrate with electrodes for the measurement of the electrical properties. The sensing layer is generally heated by a heater located at the back of the substrate. SMO-based sensors are often referred to as chemiresistive sensors due to the changes in electrical resistance in response to changes in the nearby atmospheric environment. This phenomenon was first demonstrated in 1962 by Seiyama using ZnO [41]. Since then, sensitive layers composed of various binary, ternary, and composite SMOs have been studied for their ability to detect gases, making them more relevant in various applications.
The wide range of SMOs employed for gas sensing are characterized as either n-type, with electrons as the majority charge carriers, or p-type, with holes as the majority charge carriers. Unfortunately, most SMO materials exhibit limitations such as cross-sensitivity to untargeted gases, drift, high humidity sensitivity, thermal stability, and longer response times. This makes the choice of a suitable sensor for each application difficult. ZnFe2O4 based sensors are reported to be sensitive to a variety of gases, such as H2S, VOCs, NH3, NO2, and H2O. This places it in a position to be explored for a wide range of applications. Regardless of performance, an ideal SMO should possess unique sensing properties that are suitable for a particular application.

2.2. ZnFe2O4 Spinel Structure and Gas Sensing Mechanism

Since general gas detection involves a change in electrical properties, it is important to understand the structure and conductance of spinel ferrites to fully grasp the sensing mechanism of ZnFe2O4 based sensors. A spinel structure contains sublattices with a general formula of AB2O4. The bulk spinel structure of ZnFe2O4 is typical. Therefore, all the Fe3+ ions prefer to occupy the octahedral (B) sites, while divalent and non-magnetic Zn2+ ions prefer to occupy the tetrahedral (A) sites. The distribution of Zn2+ ions and Fe3+ ions at sites A and B is altered in nanocrystalline ZnFe2O4 spinel ferrite, resulting in the mixed structure described as follows: Z n 1 x 2 + F e x 3 + Z n x 2 + F e 2 x 3 + O 4 where x is the inversion parameter. This inversion of cations introduces interesting magnetic and electrical properties [42]. The concentration of charge carriers is regulated by the non-stoichiometry of cations sitting at the B-site. Therefore, the conductivity is due to the hopping of charge carriers (electrons or holes) between cations with different charge states residing in the B-site. Thus, conductance in ZnFe2O4 occurs by the hopping of electrons between F e 3 + and F e 2 + F e 3 + + e F e 2 + in the octahedral site [43,44] through the so called Verwey conduction model proposed by E.J. Verwey et al. [45] in 1947. Although a high concentration of Fe2+ in the B-site results in high conductance, it is also important to note that n-type spinel ferrite gas sensors have a slightly lower stoichiometric composition and contain couples of Fe2+ and Fe3+ in the octahedral site even prior to oxygen chemisorption [20].
In line with the well-accepted SMO-based sensor mechanism, oxygen species are adsorbed on the surface of the material. At elevated temperatures, the oxygen molecules (O2) convert to atomic oxygen, thus retracting electrons from the surface of the material as described by Equations (1)–(4). Following the above discussion, chemisorbed oxygen on the surface of ZnFe2O4 will retract electrons from the Fe2+ oxidising it to Fe3+, consequently increasing resistance, as illustrated in Figure 1. The interaction of chemisorbed oxygen with a reducing gas then leads to the release of electrons back into the surface, which reverses the process, and a change in resistance is recorded according to Equation (5). The reduction reaction reduces the Fe3+ to Fe2+ which increases the conductivity of ZnFe2O4. The change in the conductivity of the ZnFe2O4 is measured as the sensor response.
O 2 g a s O 2 a d s
O 2 a d s + e O 2 a d s   T < 147   ° C
O 2 a d s + e 2 O a d s   147   ° C < T < 397   ° C
O a d s + e O 2 a d s   T > 397   ° C
T a r g e t   g a s a d s + O G a s   p r o d u c t s + e 147   ° C < T < 397   ° C

3. Challenges and Limitations of ZnFe2O4 Sensors

Patents on preparation methods and prototypes of ZnFe2O4 sensors have been filed; however, there are not commercially available ZnFe2O4-based gas sensors. This could be due to several challenging issues hindering practical applications that are not yet resolved. There are several requirements that a sensor must meet for a specific application. For instance, in an agricultural setting, the environment surrounding raw and processed materials is inherently complex. Gas sensors can be employed to monitor environmental parameters in a changing climate, allowing producers to conduct efficient irrigation and pest control. Therefore, the gas sensors in those environments should be able to function under extreme pressures and temperatures [47]. Similar to the food industry, sensors need to meet certain standards to be employed for a particular application. Such standards can be evaluated by accessing the sensing properties of the developed SMO. Like most SMO-based sensors, ZnFe2O4 based sensors have limitations that hinder their practical applications, including: (i) operating temperature, which is one of the important parameters for identifying the optimal sensing performance of a gas sensor. The operating temperatures of ZnFe2O4 based sensors are reported to be between 100 °C and 400 °C. Such high temperatures can cause thermal instability within the ZnFe2O4 structure, affecting the long-term stability of the sensor. Furthermore, gas sensors that operate at high operating temperatures suffer from degradation of electrode contacts, causing drastic changes in sensor properties and sensor design challenges [48]. (ii) Secondly, the performance of a sensor in harsh, humid conditions is an important aspect for commercial gas sensors. ZnFe2O4 sensors are extremely sensitive to humidity and have been reported to be good humidity sensors. However, this is an issue for the detection of VOCs in applications such as diabetes diagnosis or monitoring of food products in refrigeration systems where moisture cannot be avoided. (iii) Thirdly, the gas selectivity of the ZnFe2O4 sensors is another sensing property that needs careful consideration. The selectivity of a sensing material is evaluated by the ratio of the response toward a particular gas to that of another. SMO materials are not selective in nature, but they can be modified to react more to a specific target gas. ZnFe2O4 already shows good sensing characteristics and is highly selective for H2S and VOCs [49]. The challenge it faces is the selectivity towards a specific gas among a complex mixture of VOCs.

4. Strategies for the Enhancement of ZnFe2O4 Sensing Properties

To date, several studies have been conducted investigating the sensing properties of ZnFe2O4 towards VOCs, including ethanol [50], acetone [25], formaldehyde [51], BTX [52], etc. As outlined, most of the investigated VOCs are found in almost all industries, and this positions ZnFe2O4 as a potential gas sensor for wide application. To meet the requirements for application in these fields, ZnFe2O4 gas sensors should be able to detect very low concentrations of VOC and demonstrate high selectivity. Research and development using different strategies has been conducted in the past few decades targeting specific gases.

4.1. Effects of Zn2+ Metal Ion Substitution

The types of cations, as well as the compositions and site occupancies, determine the physical, chemical, and electrochemical properties of spinel-type ferrite-based metal oxides because they can accommodate a wide range of cations with more than one oxidation state between the tetrahedral and octahedral sites [53,54]. Metal-ions such as Ni2+ and Cu2+ can substitute Zn2+ in the lattice structure of ZnFe2O4. Due to the different properties each metal-ion possesses, substitution of Zn2+ by other metal-ions can fine tune the properties of ZnFe2O4 to suit different applications [55,56,57,58,59]. Depending on the size of the substituted metal-ion, the structural features of ZnFe2O4 can be altered with possible changes in the lattice structure, crystal size, and morphology. For instance, Shen et al. [60] substituted Zn2+ with Co2+ using the formula Co1−xZnxFe2O4 with 0.0 ≤ x ≤ 0.5 and demonstrated that the ‘a’ lattice parameter decreased with an increase in Zn content until x = 0.3, where the lattice parameter increased dramatically thereafter. This dependence of the lattice parameter on the substitution of metal-ions was also observed by Chakrabarty et al. [61], where they substituted Zn2+ with Ni2+. In their work, they showed that both the ‘a’ lattice parameter and crystal size decreased with an increase in Ni2+ concentration, with crystal size changing by 1.86 nanometers at full substitution.
During substitution of Zn2+ in the tetrahedral site, redistribution of cations from the tetrahedral site to the octahedral site may occur, with the substituted metal-ion occupying the octahedral position, migrating the Fe3+ to the tetrahedral site [62]. This alters the conductivity of the spinel and, ultimately, its sensing properties. A common substitution in gas sensing is the Zn2+ by Ni2+ of ZnFe2O4. This substitution has been reported to reduce the operating temperature of the sensor by invoking two possible adsorption pathways at the tetrahedral and octahedral sites [63]. Furthermore, Ni2+ ions preferably occupy octahedral sites and participate in spinel conductivity [55]. A higher concentration of Ni2+ in the octahedral site promotes greater oxygen adsorption, which oxidizes Ni2+ to Ni3+ to maintain electrical neutrality, resulting in more surface oxygen reacting with the target gas and a higher response [20]. In a previous report, the substitution of Zn ions by 10% Ni ions resulted in a unique electronic structure with the correct charge concentration, enhancing the sensing performance [64]. However, to achieve much better sensing properties in Ni2+ substituted ZnFe2O4, it is necessary to control the right amount of Fe2+ and Ni2+ in the octahedral site, and research around this requires further development. Furthermore, different metal-ion substitutions have previously been demonstrated to enhance sensing properties.
VOCs. For example, Wu et al. [65] prepared copper-substituted ZnFe2O4 hollow micro-spheres using facile solvothermal and annealing technology. As displayed in Figure 2, the Cu2+ substituted ZnFe2O4 ferrites showed a high response to a very low concentration of acetone (0.8 ppm). The enhanced sensing properties of the Cu2+ substituted ZnFe2O4 sensors were attributed to the effects of Cu2+ on the lattice cation distribution, which resulted in a higher electron depletion layer and adsorbing capacity. It is important to note that not all substitutions enhance sensing properties. For instance, substitution of Cu2+ in ZnFe2O4 significantly enhances its sensing properties, while substitution of Zn2+ into CuFe2O4 nanostructures reduces the response. However, the relation between sensing properties and the number of substituted metal-ions varies with different target gases [63].

4.2. Effects of Structural Morphology

Metal-ions exposed to different facets have certain adsorption interactions with gas molecules. Due to facets being flat surfaces on a specific geometric shape or morphology, the surface structure, corners, and edges of the SMO have a stronger relationship to the exposed facet [66]. Based on this, different gas molecules prefer to adsorb in different formations on particular surfaces of ZnFe2O4 [67]. To enhance the sensing properties of ZnFe2O4, numerous efforts have been dedicated to synthesizing different structural morphologies such as nanotubes, nanospheres, and nanofibers that will expose the preferred facet for gas adsorption. However, the effects of surface geometry have not been extensively studied in gas sensing due to challenges pertaining to morphology manipulation. On another note, SMOs with a high surface area and low density are preferred for gas sensing, as they generally exhibit good sensitivity [68]. The surface area depends on morphology as a result of variations in the surface-to-volume ratio. Furthermore, the porosity accompanying the surface area allows access to different active sites and facets, which have consequences for sensitivity and selectivity [69]. The effect of porosity on ZnFe2O4 sensing properties was demonstrated by Zhang et al. [68], where the porous structure allowed the detected gas to diffuse through both the outer and inner surfaces of the tubes (Figure 3). This morphology and the small grains that make up the unique interconnected channel structure improved the detection behavior of ZnFe2O4. Due to the dense structure of a compact film, the ZnFe2O4 nanoparticle-based sensor had a relatively low surface-to-volume ratio and could only detect gases effectively using a thin layer of the film near the surface.
In another study conducted by Zhou et al. [70], ZnFe2O4 porous nanospheres displayed excellent selectivity towards 30 ppm acetone and a remarkable response of 11.8, which was 2.5 times higher than that of ZnFe2O4 nanoparticles. The porous structure of the nanospheres that exposes the active surface of different facets was found to be the main reason for the observed remarkable sensing performance. On the other hand, Dong et al. [71] reported on the synthesis of monodisperse ZnFe2O4 nanospheres for the detection of toluene. Due to the large specific surface area, they could detect a low concentration of toluene, attaining a response of 1.41 at 1 ppm. Such enhancement in the sensing performance of porous ZnFe2O4 was attributed to diffusion and the target gas being exposed to a higher active surface area.
Certain facets are more reactive than others due to active sites that are more distinguished by the arrangement of surface atoms. To confirm this concept, Zou et al. [67] applied density functional theory (DFT) to study the adsorption sites and the interaction mechanism of the NH3 molecule on the surface of ZnFe2O4. It was concluded that NH3 molecules preferred to adsorb with the formation of a H3N-Zn bond on the surface (110) of the spinel ZnFe2O4. Thus, the adsorption of target gas molecules depends on the preferred surface and active sites of high-energy surfaces. Studies directly focusing on determining high-energy facets of ZnFe2O4 through experimental and computational studies are still limited; however, research has been focused on preparing morphologies that can expose such facets to enhance the sensing performances. ZnFe2O4 hollow spheres that showed excellent selectivity were prepared by Yang et al. [72] for the detection of ethylene glycol. Li et al. [73] synthesized porous ZnFe2O4 nanorods with a networked pore structure for the detection of acetone. Table 2 compares the performance of different ZnFe2O4 sensors with different morphologies, while Figure 4 displays some of the morphologies synthesized from the literature. Qu et al. [74] conducted a clear comparison of the effect of morphology on ZnFe2O4, which emphasizes the importance of morphology on the sensing properties of ZnFe2O4, the response and selectivity from solid, yolk-shell to double-shell, which were attributed to porous structures (Figure 5).
Figure 4. SEM micrographs of different morphologies of (a) porous ZnFe2O4 olives. (Reproduced with permission from [75], Copyright 2019, Elsevier). (b) porous ZnFe2O4 spheres. (Reproduced with permission from [70], Copyright 2014, Elsevier). (c) ZnFe2O4 tubes (Reproduced with permission from [68], Copyright 2007, Elsevier). and (d) ZnO/ZnFe2O4 nanocubes. (Reproduced with permission from [76], Copyright 2017, RSC Advances).
Figure 4. SEM micrographs of different morphologies of (a) porous ZnFe2O4 olives. (Reproduced with permission from [75], Copyright 2019, Elsevier). (b) porous ZnFe2O4 spheres. (Reproduced with permission from [70], Copyright 2014, Elsevier). (c) ZnFe2O4 tubes (Reproduced with permission from [68], Copyright 2007, Elsevier). and (d) ZnO/ZnFe2O4 nanocubes. (Reproduced with permission from [76], Copyright 2017, RSC Advances).
Processes 11 03122 g004
Table 2. ZnFe2O4 based sensors with different morphologies and their corresponding specific surface area morphologies and performance responses towards different VOCs.
Table 2. ZnFe2O4 based sensors with different morphologies and their corresponding specific surface area morphologies and performance responses towards different VOCs.
ZnFe2O4 MorphologySurface Area (m2g−1)Target GasConcentration (ppm)Reproducibility (Cycles)/Stabilty (Days)Ra/RgRef.
Mesoporous nanostructures103.6Acetone100-/3011.6[77]
Porous Olives96.5Ethanol2008/7223[75]
Porous nanospheres59.0Acetone303/3011.8[70]
Monodisperse nanospheres87.40Toluene1/1005/201.41/9.98[71]
Hollow urchin-like core-shell spheres50.62Toluene1006/3079[78]
Porous nanospheres59.0Acetone1003/3042.1[70]
Porous nanorods82.01Acetone100-/3052.8[73]
Hollow spheres48.1Ethylene glycol100-/8 weeks35.5[72]
Figure 5. (a) Comparison of sensor responses of ZnFe2O4 double-shell, york-shell, and solid microsphere-based sensors to 20 ppm acetone as a function of the operating temperature and (b) ZnFe2O4 double-shell, york-shell, and solid microsphere-based sensors to various gases at 206 °C. Reproduced with permission from [74], Copyright 2018, Elsevier.
Figure 5. (a) Comparison of sensor responses of ZnFe2O4 double-shell, york-shell, and solid microsphere-based sensors to 20 ppm acetone as a function of the operating temperature and (b) ZnFe2O4 double-shell, york-shell, and solid microsphere-based sensors to various gases at 206 °C. Reproduced with permission from [74], Copyright 2018, Elsevier.
Processes 11 03122 g005

4.3. Effects of Generated Surface Defects

For spinel structures, cation mixing can create centers of local charges that serve as electron or hole traps that lead to a point defect (also called an antisite defect) [79]. This is often obtained after a heat treatment process where the spinel structure normally undergoes an order-disorder transition. Two of the most common defects that can occur in ZnFe2O4 are Schottky and Frenkel defects. The Schottky defect involves simultaneous movement of the cation and anion to the surface, while the Frenkel defect occurs either on the cation sub-lattice or on the anion sub-lattice. These two defects are created independently, and one usually dominates the other. El-Sayed [80] investigated the electrical conductivity of Ni1−xZnxFe2O4 (x = 0.1, 0.3, 0.5, 0.7, and 0.9). The type of defect was identified by carefully measuring the relative variation in densities ΔD/D0 and lattice constant Δa/a0, where D0 denotes the true density calculated by XRD analysis and ΔD represents the difference between the true and obtained bulk densities of the prepared samples. If ΔD/D0 > Δa/a0, then the Schottky defect dominates, while the Frenkel defect dominates when ΔD/D0 is approximately equal to Δa/a0 [80]. In this study, the produced material showed a Schottky defect dominating for all values of x.
Structural defects are regarded as the most important parameter in gas sensing, as they influence the chemisorption characteristics of the sensor, leading to enhanced sensing performance. As described above, it is understood that the gas sensing of ZnFe2O4 nanostructures is a result of resistance modulation that comes from electron hoping between iron ions located in the octahedral site (Fe3+ + e ↔ Fe2+). A higher amount of Fe2+ in the octahedral site can increase oxygen chemisorption, which then results in higher resistance modulation. Thus, maintaining the right concentration of Fe2+ in the octahedral site without undermining the structural stability of the ZnFe2O4 FCC structure is crucial to enhancing its sensing properties. Fe2+ can be regulated by synthesizing ZnFe2O4 with the non-stoichiometry of iron ions. Sutka et al. [81] prepared a non-stoichiometry Fe excess ZnFe2O4, and when compared to the stoichiometric ZnFe2O4, the sensitivity towards VOCs increased by threefold. Fe2+ can also be set by controlling surface defects; in particular, oxygen vacancies can be compensated by reducing Fe3+ to Fe2+ which then results in higher oxygen chemisorption. The presence of oxygen vacancies can also narrow the band gap of ZnFe2O4 by raising the position of the valence band, as illustrated in Figure 6. Such a narrow band gap can result in more electron transfer, which in turn results in a higher gas response.

4.4. Effects of Surface Doping

In an attempt to produce sensing materials that exhibit improved sensor performance, researchers have focused their attention on improving the chemical composition through surface decoration. Surface decoration involves the addition of catalytically active sites to the surface of the sensing material. Such active sites will enhance the sensing performance by favoring certain interactions of gas species with the surface and reducing the response times of the sensors [83]. The most common surface decoration species are noble metal nanoparticles (i.e., Au, Ag, Pt, Pd, etc.) due to their multiple sensitization effects [84]. Two sensitization mechanisms can occur that are influenced by surface decorations, including chemical and electronic sensitization. In electronic sensitization, noble metal nanoparticles incorporated on the surface act as a strong electron acceptor, thus increasing the amount of Fe2+ in the octahedral and increasing the depletion around the interface. In chemical sensitization, incorporated noble metal nanoparticles provide preferred adsorption sites, which create active sites that are spilled over onto the ZnFe2O4 surface to react with target gases. Therefore, the observed sensing enhancement in ZnFe2O4 through surface decoration is often attributed to both electronic and chemical sensitization. However, chemical sensitization is a more logical phenomenon for the improvement of sensing properties, given that catalytic reactions on SMO surfaces serve as the receptor function of SMO-based gas sensors and that catalytically active metal nanoparticles are used as additives to strengthen these reactions to increase sensor selectivity and sensitivity of the sensors [85,86].
Au-decorated ZnFe2O4 was recently prepared by Li et al. [26] using the hydrothermal method for the detection of chlorobenzene. In their study, it was found that the Au decorated ZnFe2O4 yolk-shell sphere revealed an improved sensing performance compared to the pristine ZnFe2O4 yolk-shell sphere owing to both electronic and chemical sensitizations induced by Au addition to the surface of ZnFe2O4 yolk-shell spheres. The addition of Au particles particularly resulted in a response of 90.9 toward 10 ppm of chlorobenzene, a detection limit of 100 ppb, and outstanding selectivity. It was then concluded based on this finding that the Au-decorated ZnFe2O4 yolk-shell sphere-based sensor can be used for monitoring chlorobenzene in the food industry. Nemufulwi et al. [87] evaluated the effects of Au on ZnFe2O4 nanoparticles. It was also discovered that the addition of Au metal nanoparticles to the ZnFe2O4 surface enhanced the sensing behavior of nanostructured ZnFe2O4. In another study, Lv et al. [88] doped ZnFe2O4 with an Sb metalloid. They achieved a response of 35.5 from the 0.5% Sb-doped ZnFe2O4, which was twice as high as that of pure ZnFe2O4. Doping was discovered to increase the specific surface area in the Sb-doped ZnFe2O4 samples, increasing the number of active adsorption sites for the target gases. Furthermore, an appropriate amount of element Sb on the surface of ZnFe2O4 increased the concentration of deficient and surface chemisorbed oxygen, further improving the sensing performance. In this case, the element Sb played a significant role by acting as a catalyst that strengthened surface reactions by increasing the amount of chemisorbed oxygen.

4.5. Heterostructures

SMOs come in two varieties: p-type semiconductors, where holes make up the majority of the charge carriers, and n-type semiconductors, where electrons make up the majority of the charge carriers. Numerous research studies have demonstrated that composite metal-oxides can be used to enhance selectivity, sensitivity, and other crucial sensing qualities. Employing SMOs with similar or dissimilar conductivity, metal heterostructures can be formed, whereby different types of SMOs can form a physical interface [89]. Depending on the bandgap of the chosen SMO to form an interface with ZnFe2O4, most of the heterojunctions formed are in the context of type I and type II band-gap alignment. A type I configuration is characterized by straddling band alignment, where at the heterojunction the edges of the valance and conduction band are located within the base SMO or vice versa, forming a nested type I heterostructure, as shown in Figure 7. In type II heterostructures, the adjacent domains of the SMOs form a staggered alignment at the heterojunction. The valence and conduction bands of the added SMO are either higher or lower than those of the base SMO. Heterostructures have attracted attention due to their potential to enhance sensing sensitivity by modulating resistance through heterojunction barriers [90].
Several structural architectures that are used to produce heterojunction between two SMOs have been developed, i.e., those with simple mixtures that can be randomly distributed throughout the materials (e.g., graphene-ZnFe2O4) [91], a base material with an SMO deposited on its surface (e.g., CuO@ZnFe2O4) [92], and a well-defined partition between the two SMOs (e.g., ZnO/ZnFe2O4) [76]. The most common interface formed for gas sensing applications is the p-n heterojunction [93]. However, in each of the structural architectures involving the n-type ZnFe2O4 nanostructure, several heterojunctions can be formed to improve the sensing properties. Enhanced sensor performance can arise from charge transfer and the creation of a charge depletion layer when two distinct SMOs are in close electrical contact and the fermi levels across the interface are equilibrated to the same energy [93]. The charge transfer was perfectly illustrated by W. Li et al. [94] in the case of the ZnO-ZnFe2O4 heterojunction, as illustrated in Figure 5. To separate the excited electron-hole pairs on the ZnFe2O4 side of the heterointerfaces, the excited electrons of ZnFe2O4 are driven to the conduction band of ZnO. However, excited holes in the ZnO valence band are preferentially captured and accumulated in the ZnFe2O4 valence band, which has the most negative valence band of ZnFe2O4. Therefore, a new potential barrier is created between the ZnO-ZnFe2O4 interfaces. In the literature survey conducted here, ZnFe2O4 nanostructures have been used more often to create the n-n heterojunction for gas-sensing applications. Wang et al. [95] prepared ZnO/ZnFe2O4 with hollow nanocages that showed enhanced sensing performance to 100 ppm acetone as compared to ZnO nanocages and ZnFe2O4 nanospheres. Furthermore, the ZnO/ZnFe2O4 heterostructure nanocages exhibited a low detection limit of 1 ppm for acetone, which shows good applicability in the food industry. The improved gas detection performance was later determined to be due to the distinctive structure and heterojunction created at the ZnO and ZnFe2O4 interfaces. In a different investigation, Zhang et al. [92] examined the gas sensing capabilities of ZnFe2O4 both before and after CuO surface modification. The results showed that the CuO@ZnFe2O4 heterostructures had a greater gas response and higher xylene selectivity. In contrast, Li et al. [96] created ZnFe2O4/α-Fe2O4 porous microrods with improved TEA sensing performance using a quick and reliable sacrificial template method. Three types of n-n junctions, including heterojunctions of ZnFe2O4-Fe2O3 and homojunctions of ZnFe2O4-ZnFe2O4 and Fe2O3-Fe2O3, were anticipated to occur in the nanocomposite when ZnFe2O4 nanoparticles were placed in Fe2O4. While only the Fe2O3-Fe2O3 homojunction is present in pure-Fe2O3, the enhanced TEA sensitivity is believed to be mainly due to the newly created ZnFe2O4-Fe2O3 heterojunction. Figure 5 displays the TEM of the microrods and the corresponding SAED displaying the diffraction rings, which were indexed to rhombohedral a-Fe2O4 and cubic ZnFe2O4. The HR-TEM image clearly confirms a heterojunction between ZnFe2O4 and a-Fe2O4, as displayed in Figure 8. Table 3 lists the literature on ZnFe2O4 heterostructures and their responses to several gases.
Figure 7. Schematic energy-level diagrams of hollow ZnO/ZnFe2O4 microspheres with a down/up-shift of actual Fermi-level modulated by thermally dependent ionization reactions: (a) room temperature, (b) low temperature, (c) middle temperature, and (d) high temperature. Reproduced with permission from [94], Copyright 2017, Elsevier.
Figure 7. Schematic energy-level diagrams of hollow ZnO/ZnFe2O4 microspheres with a down/up-shift of actual Fermi-level modulated by thermally dependent ionization reactions: (a) room temperature, (b) low temperature, (c) middle temperature, and (d) high temperature. Reproduced with permission from [94], Copyright 2017, Elsevier.
Processes 11 03122 g007
Figure 8. (a,b) TEM images; (c) SAED pattern; and (d) HRTEM image of the prepared ZnFe2O4/a-Fe2O3 PMRs. Reproduced with permission from [96], Copyright 2018, Elsevier.
Figure 8. (a,b) TEM images; (c) SAED pattern; and (d) HRTEM image of the prepared ZnFe2O4/a-Fe2O3 PMRs. Reproduced with permission from [96], Copyright 2018, Elsevier.
Processes 11 03122 g008
Table 3. Literature of different ZnFe2O4 heterostructures towards different target gases.
Table 3. Literature of different ZnFe2O4 heterostructures towards different target gases.
ZnFe2O4 HeterostructureMorphologyInterface/
Heterojunction
Target GasConcentration (ppm)Reproducibilty (Cycles)/Stability (Days)Ra/RgRef.
ZnO/ZnFe2O4Lychee-like core, shell, and hollow microspheren-n junctionAcetone100-/2036.6[97]
CuO@ZnFe2O4Yolk shell microspheresp-n junctionXylene1005/6024[92]
ZnFe2O4/ZnOFlower-like microstructuresn-n junctionAcetone506/608.3[98]
ZnO/ZnFe2O4Microspheren-p-n junctionBenzene1-/-1.1[94]
ZnO/ZnFe2O4Tetrapod-liken-n junctionEthanol100-/3013.97[73]
ZnFe2O4/ZnONanosheets assembled into microspheresn-n junctionTrimethylamine10010/3031.5[99]
ZnO/ZnFe2O4Hierarchical kiwifruit-liken-n junctionTriethylamine1006/3040.15[98]
ZnFe2O4/α-Fe2O3Porous microrodsn-n junctionTriethylamine1004/3042.4[96]
Fe2O3/ZnFe2O4Porous spindlesn-n junctionTriethylamine203/-60.24[100]
CaFe2O4/ZnFe2O4Walnutp-n junctionIsoprene309/3019.5[101]
ZnFe2O4/(Fe-ZnO)Nanocompositen-n junctionAcetone1003/-30.8[102]

5. Applications of ZnFe2O4 Gas Sensor

Gas sensing is of great importance in various fields such as environmental safety, food safety and fermentation control in the food industry, safety from toxic and explosive gases, detection of combustible gases in households, and patient diagnostics. Several sensing methods have been adopted for various sensing applications in different fields. Numerous findings reported in the literature on ZnFe2O4 based sensors do not describe their direct application in a specific field. Instead, reports are specific about the target gas, and sensing measurements are conducted in practical experiments that simulate real-life applications. Several reports have demonstrated the potential of ZnFe2O4 as an acetone detector [25,70,73,76,87,91,102,103]. In general, acetone is a type of important organic synthesis raw material that is obtained and widely used in agricultural chemicals, plastics, rubber, medicine, coatings, and commercial production, such as spray paint [104]. Although acetone is necessary in industrial production, it is flammable and volatile and can easily ignite when exposed to light or high temperatures, posing a significant risk to commercial production. Furthermore, acetone has a strong impulse with prolonged exposure that causes chronic illness with a paralysis effect produced in the central nervous system, triggering nausea, weakness, dizziness, pharyngitis, bronchitis, and even stupor. As a result, to ensure industrial safety and human health, acetone vapors in commercial production must be detected quickly and accurately. This only requires high sensitivity, a quick response time, and a long-term, stable acetone gas sensor. Furthermore, the breath of a diabetic person releases acetone, suggesting high levels of ketones in their blood. Therefore, acetone can be detected in human breath for the diagnosis of diabetes at concentrations above 1.8 ppm. ZnFe2O4 based sensors have lower acetone detection limits in ppb, making them good candidates for medical diagnosis [74].
In addition, in the food sector, highly sensitive ZnFe2O4 can be used to monitor and adjust the ethanol content in fruit ripening storage chambers online or with an alcoholmeter [75]. Previous reports have also demonstrated the utilization of ZnFe2O4 sensors in the chemical and production industries for the detection of neurotoxic gases such as benzene, toluene, xylene, styrene, and trichloroethylene [105]. Benzene, toluene, and xylene gases, which are also collectively known as BTX, are frequently used in industries and require sensitive gas sensors to detect them. A full discussion of the toxicity levels of these gases was reported by Mary Ellen Fleming-Jones and Robert E. Smith [105].
ZnFe2O4 is also sensitive to triethylamine (TEA), which is considered an organocatalyst and solvent with a wide range of applications. It is often employed in industrial manufacturing as a synthetic dye and preservative due to its relative safety, commercial availability, and low price [106]. Due to its good physical and chemical characteristics, it is also frequently utilized in large amounts in chemical investigations. However, even at modest levels of exposure, TEA can cause a variety of other health problems when the concentration is too high. It can also cause pulmonary swelling and poisoning when inhaled [107]. TEA is also explosive, and sensitive ZnFe2O4 can be used for its detection. Table 4 lists some of the recent developments in ZnFe2O4 sensors targeting different applications.

6. Future Perspective

In the food industry, several molecules can contribute to the aroma and odor of food; thus, the selectivity problem becomes more imperative for sensor application. While standard techniques such as GC/MS can isolate and quantify the molecular information of VOC mixtures, concentration extraction methods are required to detect VOCs below the ppm range, and complex instrumentation makes this impractical for daily use. On the other hand, ZnFe2O4 based sensors suffer several limitations, as outlined in Section 3. Poor selectivity is by far the biggest limitation hindering the wide practical application of SMO-based sensors. This is because SMO-based sensors lack an adequate ability to discriminate between one gas and a mixture of gases. As discussed in previous sections, several attempts have been made to address this selectivity issue. However, while these strategies have been found to increase the response values towards certain gases, the total discrimination of the interfering gas remains a major challenge. Artificial olfactory systems have been proven to be effective methods for resolving the selectivity problem in SMO-based sensors [115]. These systems use cross-sensitive olfactory receptor arrays, olfactory codes, and the brain’s recognition system to imitate the biological olfactory organ’s ability to discriminate [115]. Similar to natural smart sensor systems, artificial smart sensor systems can be built that include output vectors, pattern recognition algorithms, and gas sensor arrays.
Smart sensors can wirelessly capture sensory data from the Internet of Things (IoT) and transmit it to the cloud, enabling machine learning (ML) algorithms to be used for pattern recognition and prediction techniques to address the selectivity problem [116]. Algorithms applicable to SMO sensors focus on three parts: (a) preprocessing of data; (b) dimensionality and feature extraction; and (c) recognition systems to identify and classify different gas types. The first part looks at correcting the base line of the sensor response caused by drift. Most SMOs suffer from drift, whereby the resistance in the air changes after exposure to the target gas. This may be caused by an inability to desorb all target gas molecules when the sensor is no longer exposed to them or the diffusion of oxygen vacancies that causes a change in conductivity in the space charge layer [117]. The second part is to identify meaningful characteristics to describe the sensor output specific to each gas, and then to classify or predict the last part. Figure 9 shows the systematic procedures adopted in the gas sensor array. These algorithms have not been used in the case of ZnFe2O4 gas sensors. In this case, nanostructured ZnFe2O4 sensors can be used in a gas sensor array to improve performance in a specific application.

7. Conclusions

In summary, ZnFe2O4 nanostructures are good candidates for application in different industrial sectors due to their ability to easily sense various gases that are generally found in different industries. Applications such as food quality control and ZnFe2O4 need to be further developed for good stability, excellent selectivity, and sensitivity under specific ambient conditions. Several strategies to achieve enhanced performance can be summarized as follows:
  • The development of novel synthesis procedures with excellent control of morphology can create access to more active sites.
  • Different noble metals (i.e., Ag, Au, and Pd) can improve the sensing properties by chemically or electronically sensitizing ZnFe2O4.
  • Introduction of structural defects by controlling the concertation of Fe2+ in the octahedral site and narrowing the band-gap of ZnFe2O4 to enhance sensing properties.
  • Heterostructure design to form hybrid nanocomposites with either p-type or n-type materials, which will result in a new depletion layer that enhances the sensing response.
  • Integration of several ZnFe2O4 nanostructures with different morphologies to produce a sensor array with enhanced sensing capabilities for specific applications in the food industry.

Author Contributions

Conceptualization, Editing, funding acquisition, Supervision G.H.M.; Research investigation; writing—original draft preparation; M.I.N.; Editing, visualization, co-supervision, H.C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Department of Science and Innovation, South Africa, and the Council for Scientific and Industrial Research (Project No. C2F0099), and the South African Research Chairs Initiative of the Department of Science and Technology and National Research Foundation of South Africa (Grant 84415).

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconducting metal oxides as sensors for environmentally hazardous gases. Sens. Actuators B Chem. 2011, 160, 580–591. [Google Scholar] [CrossRef]
  2. Volatile Organic Compound Gas Sensor Market Size Report, 2019–2025. Available online: https://www.grandviewresearch.com/industry-analysis/volatile-organic-compound-gas-sensor-market (accessed on 11 November 2020).
  3. Mandoj, F.; Nardis, S.; Di Natale, C.; Paolesse, R. Porphyrinoid Thin Films for Chemical Sensing. In Encyclopedia of Interfacial Chemistry; Wandelt, K., Ed.; Elsevier: Oxford, UK, 2018; pp. 422–443. [Google Scholar]
  4. Kassal, P.; Steinberg, M.D.; Steinberg, I.M. Wireless chemical sensors and biosensors: A review. Sens. Actuators B Chem. 2018, 266, 228–245. [Google Scholar] [CrossRef]
  5. Promphet, N.; Ummartyotin, S.; Ngeontae, W.; Puthongkham, P.; Rodthongkum, N. Non-invasive wearable chemical sensors in real-life applications. Anal. Chim. Acta 2021, 1179, 338643. [Google Scholar] [CrossRef]
  6. Oosthuizen, D.N.; Korditis, I.; Swart, H.C.; Motaung, D.E. Facile control of room temperature nitrogen dioxide gas selectivity induced by copper oxide nanoplatelets. J. Colloid Interface Sci. 2020, 560, 755–768. [Google Scholar] [CrossRef] [PubMed]
  7. Zhou, N.; Liu, T.; Wen, B.; Gong, C.; Wei, G.; Su, Z. Recent Advances in the Construction of Flexible Sensors for Biomedical Applications. Biotechnol. J. 2020, 15, e2000094. [Google Scholar] [CrossRef]
  8. Lu, Q.; Xie, X.; Parlikad, A.K.; Schooling, J.M.; Konstantinou, E. Moving from Building Information Models to Digital Twins for Operation and Maintenance. Proc. Inst. Civ. Eng. Smart Infrastruct. Constr. 2020, 174, 46–56. [Google Scholar] [CrossRef]
  9. Tonezzer, M.; Le, D.T.T.; Iannotta, S.; Van Hieu, N. Selective discrimination of hazardous gases using one single metal oxide resistive sensor. Sens. Actuators B Chem. 2018, 277, 121–128. [Google Scholar] [CrossRef]
  10. Yang, B.; Myung, N.V.; Tran, T. 1D Metal Oxide Semiconductor Materials for Chemiresistive Gas Sensors: A Review. Adv. Electron. Mater. 2021, 7, 2100271. [Google Scholar] [CrossRef]
  11. Singh, R.C.; Singh, M.P.; Virk, H.S. Applications of Nanostructured Materials as Gas Sensors. Solid State Phenom. 2013, 201, 131–158. [Google Scholar] [CrossRef]
  12. Duarte, J.; Rodrigues, F.; Castelo Branco, J. Sensing Technology Applications in the Mining Industry—A Systematic Review. Int. J. Environ. Res. Public Health 2022, 19, 2334. [Google Scholar] [CrossRef]
  13. Sarnoski, P.J.; O’Keefe, S.F.; Jahncke, M.L.; Mallikarjunan, P.; Flick, G.J. Analysis of crab meat volatiles as possible spoilage indicators for blue crab (Callinectes sapidus) meat by gas chromatography–mass spectrometry. Food Chem. 2010, 122, 930–935. [Google Scholar] [CrossRef]
  14. Hernández-Mesa, M.; Ropartz, D.; García-Campaña, A.; Rogniaux, H.; Dervilly-Pinel, G.; Le Bizec, B. Ion Mobility Spectrometry in Food Analysis: Principles, Current Applications and Future Trends. Molecules 2019, 24, 2706. [Google Scholar] [CrossRef] [PubMed]
  15. Olafsdottir, G.; Martinsdottir, E.; Jonsson, E.H. Rapid Gas Sensor Measurements to Determine Spoilage of Capelin (Mallotus villosus). J. Agric. Food Chem. 1997, 45, 2654–2659. [Google Scholar] [CrossRef]
  16. Camara, M.; Gharbi, N.; Lenouvel, A.; Behr, M.; Guignard, C.; Orlewski, P.; Evers, D. Detection and Quantification of Natural Contaminants of Wine by Gas Chromatography–Differential Ion Mobility Spectrometry (GC-DMS). J. Agric. Food Chem. 2013, 61, 1036–1043. [Google Scholar] [CrossRef] [PubMed]
  17. Cozzolino, D. Recent Trends on the Use of Infrared Spectroscopy to Trace and Authenticate Natural and Agricultural Food Products. Appl. Spectrosc. Rev. 2012, 47, 518–530. [Google Scholar] [CrossRef]
  18. Mustafa, F.; Othman, A.; Andreescu, S. Cerium oxide-based hypoxanthine biosensor for Fish spoilage monitoring. Sens. Actuators B Chem. 2021, 332, 129435. [Google Scholar] [CrossRef]
  19. Carvalho, M.; Ribeiro, P.; Romão, V.; Cardoso, S. Smart fingertip sensor for food quality control: Fruit maturity assessment with a magnetic device. J. Magn. Magn. Mater. 2021, 536, 168116. [Google Scholar] [CrossRef]
  20. Šutka, A.; Gross, K.A. Spinel ferrite oxide semiconductor gas sensors. Sens. Actuators B Chem. 2016, 222, 95–105. [Google Scholar] [CrossRef]
  21. Allen, G.C.; Harris, S.J.; Jutson, J.A.; Dyke, J.M. A study of a number of mixed transition metal oxide spinels using X-ray photoelectron spectroscopy. Appl. Surf. Sci. 1989, 37, 111–134. [Google Scholar] [CrossRef]
  22. Gao, D.; Shi, Z.; Xu, Y.; Zhang, J.; Yang, G.; Zhang, J.; Wang, X.; Xue, D. Synthesis, Magnetic Anisotropy and Optical Properties of Preferred Oriented Zinc Ferrite Nanowire Arrays. Nanoscale Res. Lett. 2010, 5, 1289–1294. [Google Scholar] [CrossRef]
  23. Xu, X.; Zhou, L.; Zhai, Q.; Lu, C. Synthesis, Properties, and Formation Mechanism of Zinc Ferrite Hollow Spheres. J. Am. Ceram. Soc. 2007, 90, 1959–1962. [Google Scholar] [CrossRef]
  24. Lemine, O.M.; Bououdina, M.; Sajieddine, M.; Al-Saie, A.M.; Shafi, M.; Khatab, A.; Al-hilali, M.; Henini, M. Synthesis, structural, magnetic and optical properties of nanocrystalline ZnFe2O4. Phys. B Condens. Matter 2011, 406, 1989–1994. [Google Scholar] [CrossRef]
  25. Zhang, J.; Song, J.; Niu, H.; Mao, C.; Zhang, S.; Shen, Y. ZnFe2O4 nanoparticles: Synthesis, characterization, and enhanced gas sensing property for acetone. Sens. Actuators B Chem. 2015, 221, 55–62. [Google Scholar] [CrossRef]
  26. Li, K.; Luo, Y.; Gao, L.; Li, T.; Duan, G. Au-Decorated ZnFe2O4 Yolk-Shell Spheres for Trace Sensing of Chlorobenzene. ACS Appl. Mater. Interfaces 2020, 12, 16792–16804. [Google Scholar] [CrossRef]
  27. Li, X.; Han, C.; Lu, D.; Shao, C.; Li, X.; Liu, Y. Highly electron-depleted ZnO/ZnFe2O4/Au hollow meshes as an advanced material for gas sensing application. Sens. Actuators B Chem. 2019, 297, 126769. [Google Scholar] [CrossRef]
  28. Nemufulwi, M.I.; Swart, H.C.; Mhlongo, G.H. Highly selective acetone detection displayed by a surface engineered fiber-like ZnFe2O4 based-sensor following heat-treatment ramping rate variation. Mater. Lett. 2023, 330, 133214. [Google Scholar] [CrossRef]
  29. Nemufulwi, M.I.; Swart, H.C.; Shingange, K.; Mhlongo, G.H. ZnO/ZnFe2O4 heterostructure for conductometric acetone gas sensors. Sens. Actuators B Chem. 2023, 377, 133027. [Google Scholar] [CrossRef]
  30. He, L.; Hu, J.; Yuan, Q.; Xia, Z.; Jin, L.; Gao, H.; Fan, L.; Chu, X.; Meng, F. Synthesis of porous ZnFe2O4/SnO2 core-shell spheres for high-performance acetone gas sensing. Sens. Actuators B Chem. 2023, 378, 133123. [Google Scholar] [CrossRef]
  31. Liu, M.; Wang, C.; Yang, M.; Tang, L.; Wang, Q.; Sun, Y.; Xu, Y. Novel strategy to construct porous Sn-doped ZnO/ZnFe2O4 heterostructures for superior triethylamine detection. Mater. Sci. Semicond. Process. 2021, 125, 105643. [Google Scholar] [CrossRef]
  32. Saito, N.; Haneda, H.; Watanabe, K.; Shimanoe, K.; Sakaguchi, I. Highly sensitive isoprene gas sensor using Au-loaded pyramid-shaped ZnO particles. Sens. Actuators B Chem. 2021, 326, 128999. [Google Scholar] [CrossRef]
  33. Lee, J.E.; Lim, C.K.; Park, H.J.; Song, H.; Choi, S.; Lee, D. ZnO–CuO Core-Hollow Cube Nanostructures for Highly Sensitive Acetone Gas Sensors at the ppb Level. ACS Appl. Mater. Interfaces 2020, 12, 35688–35697. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, Q.; Duan, Z.; Yuan, Z.; Li, X.; Si, W.; Liu, B.; Zhang, Y.; Jiang, Y.; Tai, H. High performance ethylene sensor based on palladium-loaded tin oxide: Application in fruit quality detection. Chin. Chem. Lett. 2020, 31, 2045–2049. [Google Scholar] [CrossRef]
  35. Zhang, J.; Ma, S.; Wang, B.; Pei, S. Hydrothermal synthesis of SnO2-CuO composite nanoparticles as a fast-response ethanol gas sensor. J. Alloys Compd. 2021, 886, 161299. [Google Scholar] [CrossRef]
  36. Motsoeneng, R.G.; Kortidis, I.; Ray, S.S.; Motaung, D.E. Designing SnO2 Nanostructure-Based Sensors with Tailored Selectivity toward Propanol and Ethanol Vapors. ACS Omega 2019, 4, 13696. [Google Scholar] [CrossRef]
  37. Guo, M.; Luo, N.; Chen, Y.; Fan, Y.; Wang, X.; Xu, J. Fast-response MEMS xylene gas sensor based on CuO/WO3 hierarchical structure. J. Hazard. Mater. 2022, 429, 127471. [Google Scholar] [CrossRef]
  38. Cao, S.; Sui, N.; Zhang, P.; Zhou, T.; Tu, J.; Zhang, T. TiO2 nanostructures with different crystal phases for sensitive acetone gas sensors. J. Colloid Interface Sci. 2022, 607, 357–366. [Google Scholar] [CrossRef]
  39. Wang, D.; Zhang, D.; Mi, Q. A high-performance room temperature benzene gas sensor based on CoTiO3 covered TiO2 nanospheres decorated with Pd nanoparticles. Sens. Actuators B Chem. 2022, 350, 130830. [Google Scholar] [CrossRef]
  40. Wang, M.; Zhu, Y.; Meng, D.; Wang, K.; Wang, C. A novel room temperature ethanol gas sensor based on 3D hierarchical flower-like TiO2 microstructures. Mater. Lett. 2020, 277, 128372. [Google Scholar] [CrossRef]
  41. Seiyama, T.; Kato, A.; Fujiishi, K.; Nagatani, M. A New Detector for Gaseous Components using semiconductive thin films. Anal. Chem. 1962, 34, 1502–1503. [Google Scholar] [CrossRef]
  42. Yadav, R.S.; KuritKa, I.; Vilcakova, J.; Urbanek, P.; Machovsky, M.; Masar, M.; Holek, M. Structural, magnetic, optical, dielectric, electrical and modulus spectroscopic characteristics of ZnFe2O4 spinel ferrite nanoparticles synthesized via honey mediated sol gel combustion method. J. Phys. Chem. Solids 2017, 110, 87–99. [Google Scholar] [CrossRef]
  43. Devan, R.S.; Kolekar, Y.D.; Chougule, B.K. Effect of cobalt substitution on the properties of nickel copper ferrite. J. Physics Condens. Matter 2006, 18, 9809–9821. [Google Scholar] [CrossRef]
  44. Hosseinpour, A.; Sadeghi, H.; Morisako, A. Simulation of DC-hopping conduction in spinel ferrites using free electron gas model. J. Magn. Magn. Mater. 2007, 316, e283–e286. [Google Scholar] [CrossRef]
  45. Verwey, E.J.; Haayman, P.W.; Romeijn, F.C. Physical Properties and Cation Arrangement of Oxides with Spinel Structures II. Electronic Conductivity. J. Chem. Phys. 1947, 15, 181–187. [Google Scholar] [CrossRef]
  46. Nemufulwi, M.I.; Swart, H.C.; Mdlalose, W.B.; Mhlongo, G.H. Size-tunable ferromagnetic ZnFe2O4 nanoparticles and their ethanol detection capabilities. Appl. Surf. Sci. 2020, 508, 144863. [Google Scholar] [CrossRef]
  47. Neethirajan, S.; Jayas, D.S.; Sadistap, S. Carbon Dioxide (CO2) Sensors for the Agri-food Industry—A Review. Food Bioprocess Technol. 2008, 2, 115–121. [Google Scholar] [CrossRef]
  48. Korotcenkov, G.; Cho, B.K. Instability of metal oxide-based conductometric gas sensors and approaches to stability improvement (short survey). Sens. Actuators B Chem. 2011, 156, 527–538. [Google Scholar] [CrossRef]
  49. Darshare, S.L.; Deshmulch, R.G.; Surgavanshi, S.S.; Mulla, I.S. Gas sensing properties of Zinc Ferrite Nanoparticles Synthesized by the Molten-Salt Route. Am. Ceram. Soc. 2008, 91, 2724–2726. [Google Scholar] [CrossRef]
  50. Sutka, A.; Zavickis, J.; Mezinskis, G.; Jakovlevs, D.; Barloti, J. Ethanol monitoring by ZnFe2O4 thin film obtained by spray pyrolysis. Sens. Actuators B Chem. 2013, 176, 330–334. [Google Scholar] [CrossRef]
  51. Rahman, M.M.; Khan, S.B.; Faisal, M.; Asiri, A.M.; Alamry, K.A. Highly sensitive formaldehyde chemical sensor based on hydrothermally prepared spinel ZnFe2O4 nanorods. Sens. Actuators B Chem. 2012, 171–172, 932–937. [Google Scholar] [CrossRef]
  52. Jiang, Y.; Song, W.; Xie, C.; Wang, A.; Zeng, D.; Hu, M. Electrical conductivity and gas sensitivity to VOCs of V-doped ZnFe2O4 nanoparticles. Mater. Lett. 2006, 60, 1374–1378. [Google Scholar] [CrossRef]
  53. Park, M.; Kim, J.; Kim, K.J.; Lee, J.; Kim, J.H.; Yamauchi, Y. Porous nanoarchitectures of spinel-type transition metal oxides for electrochemical energy storage systems. Phys. Chem. Chem. Phys. PCCP 2015, 17, 3963–3977. [Google Scholar] [CrossRef] [PubMed]
  54. O’Neill, H.S.C.; Navrotsky, A. Simple spinels; crystallographic parameters, cation radii, lattice energies, and cation distribution. Am. Mineral. 1983, 68, 181. [Google Scholar]
  55. Sutka, A.; Mezinskis, G.; Lusis, A.; Stingaciu, M. Gas sensing properties of Zn-doped p-type nickel ferrite. Sens. Actuators B Chem. 2012, 171–172, 354–360. [Google Scholar] [CrossRef]
  56. Patil, R.P.; Delekar, S.D.; Mane, D.R.; Hankare, P.P. Synthesis, structural and magnetic properties of different metal ion substituted nanocrystalline zinc ferrite. Results Phys. 2013, 3, 129–133. [Google Scholar] [CrossRef]
  57. Ebrahimi, H.R.; Parish, M.; Amiri, G.R.; Bahraminejad, B.; Fatahian, S. Synthesis, characterization and gas sensitivity investigation of Ni0.5Zn0.5Fe2O4 nanoparticles. J. Magn. Magn. Mater. 2016, 414, 55–58. [Google Scholar] [CrossRef]
  58. Mukherjee, C.; Mondal, D.; Sarkar, M.; Das, J. Nanocrystalline nickel zinc ferrite thick film as an efficient alcohol sensor at room temperature. Int. J. F Environ. Agric. Biotechnol. 2017, 2, 799–804. [Google Scholar]
  59. El-Sayed, A.M.; Hamzawy, E.M.A. Structure and Magnetic Properties of Nickel–Zinc Ferrite Nanoparticles Prepared by Glass Crystallization Method. Monatsh. Chem. 2006, 137, 1119–1125. [Google Scholar] [CrossRef]
  60. Shen, X.; Xiang, J.; Song, F.; Liu, M. Characterization and magnetic properties of electrospun Co1−xZnxFe2O4 nanofibers. Appl. Phys. A 2010, 99, 189–195. [Google Scholar] [CrossRef]
  61. Chakrabarty, S.; Pal, M.; Dutta, A. Structural, optical and electrical properties of chemically derived nickel substituted zinc ferrite nanocrystals. Mater. Chem. Phys. 2011, 127, v–viii. [Google Scholar] [CrossRef]
  62. Issa, B.; Obaidat, I.M.; Albiss, B.A.; Haik, Y. Magnetic nanoparticles: Surface effects and properties related to biomedicine applications. Int. J. Mol. Sci. 2013, 14, 21266–21305. [Google Scholar] [CrossRef]
  63. Rathore, D.; Kurchania, R.; Pandey, R.K. Fabrication of Ni1−xZnxFe2O4 (x=0, 0.5 and 1) nanoparticles gas sensor for some reducing gases. Sens. Actuators A Phys. 2013, 199, 236–240. [Google Scholar] [CrossRef]
  64. Nemufulwi, M.I.; Swart, H.C.; Mhlongo, G.H. Enhanced Propanol Response Behavior of ZnFe2O4 NP-Based Active Sensing Layer Induced by Film Thickness Optimization. Processes 2021, 9, 1791. [Google Scholar] [CrossRef]
  65. Wu, K.; Lu, Y.; Liu, Y.; Liu, Y.; Shen, M.; Debliquy, M.; Zhang, C. Synthesis and acetone sensing properties of copper (Cu2+) substituted zinc ferrite hollow micro-nanospheres. Ceram. Int. 2020, 46, 28835–28843. [Google Scholar] [CrossRef]
  66. Pal, J.; Pal, T. Faceted Metal and Metal Oxide Nanoparticles: Design, Fabrication and Catalysis. Proc. Natl. Acad. Sci. USA 2020, 117, 19168–19177. [Google Scholar] [CrossRef] [PubMed]
  67. Zou, C.; Ji, W.; Shen, Z.; Tang, Q.; Fan, M. NH3 molecule adsorption on spinel-type ZnFe2O4 surface: A DFT and experimental comparison study. Appl. Surf. Sci. 2018, 442, 778–786. [Google Scholar] [CrossRef]
  68. Zhang, G.; Li, C.; Cheng, F.; Chen, J. ZnFe2O4 tubes: Synthesis and application to gas sensors with high sensitivity and low-energy consumption. Sens. Actuators B Chem. 2007, 120, 403–410. [Google Scholar] [CrossRef]
  69. Tiemann, M. Porous Metal Oxides as Gas Sensors. Chem. A Eur. J. 2007, 13, 8376–8388. [Google Scholar] [CrossRef]
  70. Zhou, X.; Liu, J.; Wang, C.; Sun, P.; Hu, X.; Li, X.; Shimanoe, K.; Yamazoe, N.; Lu, G. Highly sensitive acetone gas sensor based on porous ZnFe2O4 nanospheres. Sens. Actuators B Chem. 2015, 206, 577–583. [Google Scholar] [CrossRef]
  71. Dong, C.; Liu, X.; Xiao, X.; Du, S.; Wang, Y. Monodisperse ZnFe2O4 nanospheres synthesized by a nonaqueous route for a highly slective low-ppm-level toluene gas sensor. Sens. Actuators B Chem. 2017, 239, 1231–1236. [Google Scholar] [CrossRef]
  72. Yang, H.; Bai, X.; Hao, P.; Tian, J.; Bo, Y.; Wang, X.; Liu, H. A simple gas sensor based on zinc ferrite hollow spheres: Highly sensitivity, excellent selectivity and long-term stability. Sens. Actuators B Chem. 2019, 280, 34–40. [Google Scholar] [CrossRef]
  73. Li, L.; Tan, J.; Dun, M.; Huang, X. Porous ZnFe2O4 nanorods with net-worked nanostructure for highly sensor response and fast response acetone gas sensor. Sens. Actuators B Chem. 2017, 248, 85–91. [Google Scholar] [CrossRef]
  74. Qu, F.; Shang, W.; Thomas, T.; Ruan, S.; Yang, M. Self-template derived ZnFe2O4 double-shell microspheres for chemresistive gas sensing. Sens. Actuators B Chem. 2018, 265, 625–631. [Google Scholar] [CrossRef]
  75. Sun, K.; Song, X.; Wang, X.; Li, X.; Tan, Z. Annealing temperature-dependent porous ZnFe2O4 olives derived from bimetallic organic frameworks for high-performance ethanol gas sensing. Mater. Chem. Phys. 2020, 241, 122379. [Google Scholar] [CrossRef]
  76. Ma, X.; Zhou, X.; Gong, Y.; Han, N.; Liu, H.; Chen, Y. MOF-derived hierarchical ZnO/ZnFe2O4 hollow cubes for enhanced acetone gas-sensing performance. RSC Adv. 2017, 7, 34609–34617. [Google Scholar] [CrossRef]
  77. Wang, Y.; Liu, F.; Yang, Q.; Gao, Y.; Sun, P.; Zhang, T.; Lu, G. Mesoporous ZnFe2O4 prepared through hard template and its acetone sensing properties. Mater. Lett. 2016, 183, 378–381. [Google Scholar] [CrossRef]
  78. Zhang, H.; Hu, J.; Li, M.; Li, Z.; Yuan, Y.; Yang, X.; Guo, L. Highly efficient toluene gas sensor based on spinel structured hollow urchin-like core-shell ZnFe2O4 spheres. Sens. Actuators B Chem. 2021, 349, 130734. [Google Scholar] [CrossRef]
  79. Gritsyna, V.T.; Afanasyev-Charkin, I.V.; Kobyakov, V.A.; Sickafus, K.E. Structure and electronic states of defects in spinel of different compositions MgO.nAI2O3:Me. J. Am. Ceram. Soc. 1999, 82, 3365. [Google Scholar] [CrossRef]
  80. El-Sayed, A.M. Electrical conductivity of nickel–zinc and Cr substituted nickel–zinc ferrites. Mater. Chem. Phys. 2003, 82, 583–587. [Google Scholar] [CrossRef]
  81. Sutka, A.; Mezinskis, G.; Lusis, A.; Jakovlevs, D. Influence of iron non-stoichiometry on spinel zinc ferrite gas sensing properties. Sens. Actuators B Chem. 2012, 171–172, 204–209. [Google Scholar] [CrossRef]
  82. Peng, S.; Wang, Z.; Liu, R.; Bi, J.; Wu, J. Controlled oxygen vacancies of ZnFe2O4 with superior gas sensing properties prepared via a facile one-step self-catalyzed treatment. Sens. Actuators B Chem. 2019, 288, 649–655. [Google Scholar] [CrossRef]
  83. Franke, M.E.; Koplin, T.J.; Simon, U. Metal and Metal Oxide Nanoparticles in Chemiresistors: Does the Nanoscale Matter? Small 2006, 2, 36–50. [Google Scholar] [CrossRef]
  84. 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 Chem. 2019, 301, 127055. [Google Scholar] [CrossRef]
  85. Korotcenkov, G. The role of morphology and crystallographic structure of metal oxides in response of conductometric-type gas sensors. Mater. Sci. Eng. R 2008, 61, 1–39. [Google Scholar] [CrossRef]
  86. Korotcenkov, G.; Cho, B.K. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B Chem. 2017, 244, 182–210. [Google Scholar] [CrossRef]
  87. Nemufulwi, M.I.; Swart, H.C.; Mhlongo, G.H. Evaluation of the effects of Au addition into ZnFe2O4 nanostructures on acetone detection capabilities. Mater. Res. Bull. 2021, 142, 111395. [Google Scholar] [CrossRef]
  88. Lv, L.; Cheng, P.; Wang, Y.; Xu, L.; Zhang, B.; Lv, C.; Ma, J.; Zhang, Y. Sb-doped three-dimensional ZnFe2O4 macroporous spheres for N-butanol chemiresistive gas sensors. Sens. Actuators B Chem. 2020, 320, 128384. [Google Scholar] [CrossRef]
  89. Kwon, H.; Yoon, J.; Lee, Y.; Kim, D.Y.; Baek, C.; Kim, J.K. An array of metal oxides nanoscale hetero p-n junctions toward designable and highly-selective gas sensors. Sens. Actuators B Chem. 2018, 255, 1663–1670. [Google Scholar] [CrossRef]
  90. Walker, J.M.; Akbar, S.A.; Morris, P.A. Synergistic effects in gas sensing semiconducting oxide nano-heterostructures: A review. Sens. Actuators B Chem. 2019, 286, 624–640. [Google Scholar] [CrossRef]
  91. Liu, F.; Chu, X.; Dong, Y.; Zhang, W.; Sun, W.; Shen, L. Acetone gas sensors based on graphene-ZnFe2O4 composite prepared by solvothermal method. Sens. Actuators B Chem. 2013, 188, 469–474. [Google Scholar] [CrossRef]
  92. Zhang, N.; Ruan, S.; Han, J.; Yin, Y.; Li, X.; Liu, C.; Adimi, S.; Wen, S.; Xu, Y. Oxygen vacancies dominated CuO@ZnFe2O4 yolk-shell microspheres for robust and selective detection of xylene. Sens. Actuators B Chem. 2019, 295, 117–126. [Google Scholar] [CrossRef]
  93. 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]
  94. Li, W.; Wu, X.; Chen, Y.; Chen, J.; Gong, Y.; Han, N. Abnormal n-p-n type conductivity transition of hollow ZnO/ZnFe2O4 nanostructures during gas sensing process: The role of ZnO-ZnFe2O4 hetero-interface. Sens. Actuators B Chem. 2017, 253, 144–155. [Google Scholar] [CrossRef]
  95. Wang, X.; Zhang, S.; Shao, M.; Huang, J.; Deng, X.; Hou, P.; Xu, X. Fabrication of ZnO/ZnFe2O4 hollow nanocages through metal organic frameworks route with enhanced gas sensing properties. Sens. Actuators B Chem. 2017, 251, 27–33. [Google Scholar] [CrossRef]
  96. Li, Y.; Luo, N.; Sun, G.; Zhang, B.; Ma, G.; Jin, H.; Wang, Y.; Cao, J.; Zhang, Z. Facile synthesis of ZnFe2O4/α-Fe2O3 porous microrods with enhanced TEA-sensing performance. J. Alloys Compd. 2018, 737, 255–262. [Google Scholar] [CrossRef]
  97. Hu, Y.; Wang, H.; Liu, D.; Lin, G.; Wan, J.; Jiang, H.; Lai, X.; Hao, S.; Liu, X. Lychee-like ZnO/ZnFe2O4 core-shell hollow microsphere for improving acetone gas sensing performance. Ceram. Int. 2020, 46, 5960–5967. [Google Scholar] [CrossRef]
  98. Liu, C.; Wang, B.; Wang, T.; Liu, J.; Sun, P.; Chuai, X.; Lu, G. Enhanced gas sensing characteristics of the flower-like ZnFe2O4/ZnO microstructures. Sens. Actuators B Chem. 2017, 248, 902–909. [Google Scholar] [CrossRef]
  99. Zheng, C.; Zhang, C.; He, L.; Zhang, K.; Zhang, J.; Jin, L.; Asiri, A.M.; Alamry, K.A.; Chu, X. ZnFe2O4/ZnO nanosheets assembled microspheres for high performance trimethylamine gas sensing. J. Alloys Compd. 2020, 849, 156461. [Google Scholar] [CrossRef]
  100. Wei, Q.; Sun, J.; Song, P.; Li, J.; Yang, Z.; Wang, Q. Spindle-like Fe2O3/ZnFe2O4 porous nanocomposites derived from metal-organic frameworks with excellent sensing performance towards triethylamine. Sens. Actuators B Chem. 2020, 317, 128205. [Google Scholar] [CrossRef]
  101. Guo, W.; Huang, L.; Liu, X.; Wang, J.; Zhang, J. Enhanced isoprene gas sensing performance based on p-CaFe2O4/n-ZnFe2O4 heterojunction composites. Sens. Actuators B Chem. 2022, 354, 131243. [Google Scholar] [CrossRef]
  102. Cao, E.; Guo, Z.; Song, G.; Zhang, Y.; Hao, W.; Sun, L.; Nie, Z. MOF-derived ZnFe2O4/(Fe-ZnO) nanocomposites with enhanced acetone sensing performance. Sens. Actuators B Chem. 2020, 325, 128783. [Google Scholar] [CrossRef]
  103. Zhou, X.; Li, X.; Sun, H.; Sun, P.; Liang, X.; Liu, F.; Hu, X.; Lu, G. Nanosheet-assembled ZnFe2O4 hollow microspheres for high-sensitive acetone sensor. ACS Appl. Mater. Interfaces 2015, 7, 15414. [Google Scholar] [CrossRef]
  104. Morgott, D.A. Acetone. In Patty’s Toxicology; Wiley Online Library: New York, NY, USA, 2001; pp. 735–752. [Google Scholar]
  105. Fleming-Jones, M.E.; Smith, R.E. Volatile Organic Compounds in Foods:  A Five Year Study. J. Agric. Food Chem. 2003, 51, 8120–8127. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, Z.; Lin, R.; Fang, W.; Li, G.; Guo, Y.; Qin, Z. Triethylamine as an initiator for cracking of heptane. Energy 2006, 31, 2773–2790. [Google Scholar] [CrossRef]
  107. Brahmachari, G.; Nayek, N.; Nurjamal, K.; Karmakar, I.; Begam, S. Triethylamine—A Versatile Organocatalyst in Organic Transformations: A Decade Update. Synthesis 2018, 50, 4145–4164. [Google Scholar] [CrossRef]
  108. Karmakar, M.; Das, P.P.; Pal, M.M.; Mondal, B.B.; Majumder, S.B.; Mukherjee, K.K. Acetone and ethanol sensing characteristics of magnesium zinc ferrite nano-particulate chemi-resistive sensor. J. Mater. Sci. 2014, 49, 5766. [Google Scholar] [CrossRef]
  109. Nemufulwi, M.I.; Swart, H.C.; Mhlongo, G.H. A comprehensive comparison study on magnetic behaviour, defects-related emission and Ni substitution to clarify the origin of enhanced acetone detection capabilities. Sens. Actuators B Chem. 2021, 339, 129860. [Google Scholar] [CrossRef]
  110. Zhang, C.; Wu, Q.; Zheng, B.; You, J.; Luo, Y. Synthesis and acetone gas sensing properties of Ag activated hollow sphere structured ZnFe2O4. Ceram. Int. 2018, 44, 20700–20707. [Google Scholar] [CrossRef]
  111. Zhai, C.; Zhao, Q.; Gu, K.; Xing, D.; Zhang, M. Ultra-fast response and recovery of triethylamine gas sensors using a MOF-based ZnO/ZnFe2O4 structures. J. Alloys Compd. 2019, 784, 660–667. [Google Scholar] [CrossRef]
  112. Li, S.; Zhang, Y.; Han, L.; Li, X.; Xu, Y. Hierarchical kiwifruit-like ZnO/ZnFe2O4 heterostructure for high-sensitive triethylamine gaseous sensor. Sens. Actuators B Chem. 2021, 344, 130251. [Google Scholar] [CrossRef]
  113. Liu, S.; Guan, M.; Li, X.; Guo, Y. Light irradiation enhanced triethylamine gas sensing materials based on ZnO/ZnFe2O4 composites. Sens. Actuators B Chem. 2016, 236, 350–357. [Google Scholar] [CrossRef]
  114. Arshak, K.; Moore, E.; Cunniffe, C.; Nicholson, M.; Arshak, A. Preparation and characterisation of ZnFe2O4/ZnO polymer nanocomposite sensors for the detection of alcohol vapours. Superlattices Microstruct. 2007, 42, 479–488. [Google Scholar] [CrossRef]
  115. Hayasaka, T.; Lin, A.; Copa, V.C.; Lopez, L.P.; Loberternos, R.A.; Ballesteros, L.I.M.; Kubota, Y.; Liu, Y.; Salvador, A.A.; Lin, L. An electronic nose using a single graphene FET and machine learning for water, methanol, and ethanol. Microsyst. Nanoeng. 2020, 6, 50. [Google Scholar] [CrossRef] [PubMed]
  116. Zhu, J.; Cho, M.; Li, Y.; He, T.; Ahn, J.; Park, J.; Ren, T.; Lee, C.; Park, I. Machine learning-enabled textile-based graphene gas sensing with energy harvesting-assisted IoT application. Nano Energy 2021, 86, 106035. [Google Scholar] [CrossRef]
  117. Chen, Z.; Chen, Z.; Song, Z.; Ye, W.; Fan, Z. Smart gas sensor arrays powered by artificial intelligence. J. Semicond. 2019, 40, 111601. [Google Scholar] [CrossRef]
  118. Feng, S.; Farha, F.; Li, Q.; Wan, Y.; Xu, Y.; Zhang, T.; Ning, H. Review on Smart Gas Sensing Technology. Sensors 2019, 19, 3760. [Google Scholar] [CrossRef]
Figure 1. Illustration of an ethanol gas sensing mechanism depicting the cations involved (Fe3+ + e = Fe2+) in the conduction of ZnFe2O4 NPs. Reproduced with permission from [46], Copyright 2019, Elsevier.
Figure 1. Illustration of an ethanol gas sensing mechanism depicting the cations involved (Fe3+ + e = Fe2+) in the conduction of ZnFe2O4 NPs. Reproduced with permission from [46], Copyright 2019, Elsevier.
Processes 11 03122 g001
Figure 2. (a) Sensor responses of CuxZn1−xFe2O4 (S1 to S5, x = 0, 0.25, 0.5, 0.75, 1, respectively) to 1 ppm acetone at 100–225 °C; (b) Dynamic sensing performance of S1, S4, and S5 gas sensors to 0.8–10 ppm acetone at 125 °C; (c) response values versus acetone concentration. Reproduced with permission from [65], Copyright 2020, Elsevier.
Figure 2. (a) Sensor responses of CuxZn1−xFe2O4 (S1 to S5, x = 0, 0.25, 0.5, 0.75, 1, respectively) to 1 ppm acetone at 100–225 °C; (b) Dynamic sensing performance of S1, S4, and S5 gas sensors to 0.8–10 ppm acetone at 125 °C; (c) response values versus acetone concentration. Reproduced with permission from [65], Copyright 2020, Elsevier.
Processes 11 03122 g002
Figure 3. Schematic diagrams of (a) the structure of a ceramic tube gas sensor, (b) the sample film based on ZnFe2O4 tubes, and (c) the sample film based on ZnFe2O4 nanoparticles. Reproduced with permission from [68], Copyright 2020, Elsevier.
Figure 3. Schematic diagrams of (a) the structure of a ceramic tube gas sensor, (b) the sample film based on ZnFe2O4 tubes, and (c) the sample film based on ZnFe2O4 nanoparticles. Reproduced with permission from [68], Copyright 2020, Elsevier.
Processes 11 03122 g003
Figure 6. Schematic diagram of the energy bands for ZnFe2O4 and ZnFe2O4 having oxygen vacancy concentrations: (a) in air and (b) towards acetone. The proposed reaction mechanism of the ZnFe2O4 nanoparticle-based sensor to acetone is (c) ZnFe2O4 and (d) ZnFe2O4 with oxygen vacancy contributions. Reproduced with permission from [82], Copyright 2019, Elsevier.
Figure 6. Schematic diagram of the energy bands for ZnFe2O4 and ZnFe2O4 having oxygen vacancy concentrations: (a) in air and (b) towards acetone. The proposed reaction mechanism of the ZnFe2O4 nanoparticle-based sensor to acetone is (c) ZnFe2O4 and (d) ZnFe2O4 with oxygen vacancy contributions. Reproduced with permission from [82], Copyright 2019, Elsevier.
Processes 11 03122 g006
Figure 9. Systematic steps for smart gas sensing [118].
Figure 9. Systematic steps for smart gas sensing [118].
Processes 11 03122 g009
Table 4. Different applications of ZnFe2O4 based gas sensors.
Table 4. Different applications of ZnFe2O4 based gas sensors.
GasSource of Emission/ProductDetection Test/ApplicationSMO UsedDetection Range (ppm)/Environmental SettingRef.
Acetone and EthanolHousehold and industrial products, laboratories, and chemical industriesSafety purposesMg0.5Zn0.5 Fe2O420–200/dry air RH, RT[108]
PropanolAlcoholic beveragesQuality and classification of winesZnFe2O4
NPs
0.5–40/10–80%, 120 °C[64]
AcetoneHuman breath, fish productsDiabetes diagnostics and spoilage detectionZnFe2O4 nanoparticles0.5–40/0–60%, 120 °C[109]
n-butanolpetroleum refineries and insect repellents,Human health and safetySb-doped 3D ZnFe2O4 MPs0.049–200/25–90 RH, 250 °C[88]
Toluenelacquers, medicine, pesticides, leather manufacturing, and explosivesEnvironmental and human safety, Human healthurchin-like hollow core-shell ZnO/ZnFe2O4
ZnFe2O4
0.2–100/10–98% RH, 275 °C
0.2–100/10–98% RH, 250 °C
[78]
TolueneChemical IndustryEnvironmental safety and human safetyMonodisperse ZnFe2O4 nanospheres1–100/Dry air. 300 °C[71]
AcetoneHuman breathAcetone-threat or diabetes-breathalyzer testsDouble-shelled ZnFe2O4 microsphere0.13–200/Dry air, 206 °C[74]
AcetoneHuman breath, industryDiagnosis of diabetes, industrial processes, and health controlZnO/ZnFe2O4/Au
0.125%Graphene-ZnFe2O4
ZnFe2O4 hollow sphere and Ag-ZnFe2O4
0.3–200/33–95% RH, 225 °C
1–1000/Dry air, 275 °C
0.8–500/25–100% Rh, 175 °C
[27,91,110]
EthanolHuman breath; alcohol beveragesMedical or clinical applications; brewing process control.ZnFe2O4 thin film1–50/Dry air, 390 °C[50]
Triethylamine (TEA)Meat, wastewater, dead fish, and marine productsMeat spoilage, environmental monitoring, and human health protectionMOF-ZnO/ZnFe2O42–100/Dry air, 170 °C[111]
TEAIndustrial productionIndustrial monitoringkiwifruit-like ZnO/ZnFe2O4
ZnO/ZnFe2O4
1–200/dry air, 200 °C
5–1000/dry air, 80 and 240 °C
[112,113]
PropanolIndustrial plantsLeak detectionZnFe2O4/ZnO polymer nanocomposite500–5000/Dry air, N/A[114]
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

Nemufulwi, M.I.; Swart, H.C.; Mhlongo, G.H. Advances of Nano-Enabled ZnFe2O4 Based-Gas Sensors for VOC Detection and Their Potential Applications: A Review. Processes 2023, 11, 3122. https://doi.org/10.3390/pr11113122

AMA Style

Nemufulwi MI, Swart HC, Mhlongo GH. Advances of Nano-Enabled ZnFe2O4 Based-Gas Sensors for VOC Detection and Their Potential Applications: A Review. Processes. 2023; 11(11):3122. https://doi.org/10.3390/pr11113122

Chicago/Turabian Style

Nemufulwi, Murendeni I., Hendrik C. Swart, and Gugu H. Mhlongo. 2023. "Advances of Nano-Enabled ZnFe2O4 Based-Gas Sensors for VOC Detection and Their Potential Applications: A Review" Processes 11, no. 11: 3122. https://doi.org/10.3390/pr11113122

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