Next Article in Journal
An Adaptive and Robust Control Strategy for Real-Time Hybrid Simulation
Previous Article in Journal
Cloud Data-Driven Intelligent Monitoring System for Interactive Smart Farming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Sensing and Selective Performance Based on Bi-Doped Porous ZnSnO3 Nanospheres for Detection of n-Butanol

School of Materials Science and Engineering, Lanzhou University of Technology, Langongping Road, Lanzhou 730050, China
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(17), 6571; https://doi.org/10.3390/s22176571
Submission received: 31 July 2022 / Revised: 19 August 2022 / Accepted: 29 August 2022 / Published: 31 August 2022
(This article belongs to the Section Chemical Sensors)

Abstract

:
In this study, pure zinc stannate (ZnSnO3) and bismuth (Bi)-doped ZnSnO3 composites (Bi-ZnSnO3) were synthesized via the in situ precipitation method, and their microstructures, morphologies, chemical components, sizes, and specific surface areas were characterized, followed by testing their gas sensing properties. The results revealed that Bi-ZnSnO3 showed superior gas sensing properties to n-butanol gas, with an optimal operating temperature of 300 °C, which was 50 °C lower than that of pure ZnSnO3. At this temperature, moreover, the sensitivity of Bi-ZnSnO3 to n-butanol gas at the concentration of 100 ppm reached as high as 1450.65, which was 35.57 times that (41.01) of ammonia gas, 2.93 times that (495.09) of acetone gas, 6.02 times that (241.05) of methanol gas, 2.54 times that (571.48) of formaldehyde gas, and 2.98 times that (486.58) of ethanol gas. Bi-ZnSnO3 had a highly repeatable performance. The total proportion of oxygen vacancies and chemi-adsorbed oxygen in Bi-ZnSnO3 (4 wt%) was 27.72% to 32.68% higher than that of pure ZnSnO3. Therefore, Bi-ZnSnO3 has considerable potential in detecting n-butanol gas by virtue of its excellent gas-sensing properties.

1. Introduction

n-Butanol, a typical volatile organic compound, has been widely applied to the production of various paint solvents, extractants, and plastic rubber products [1,2]. n-Butanol, which is flammable, gives off a unique pungent smell and tends to form explosive mixtures with air [3]. High-concentration n-butanol gas is hazardous to the human body, causing symptoms such as dizziness, headache, drowsiness, and dermatitis if exposed at >50 ppm for a long time [4,5]. Therefore, n-butanol gas must be rapidly and effectively monitored for the sake of human health and safety, but expensive and cumbersome monitoring instruments are usually required for traditional n-butanol detection methods, which are also marred by their complex operation, long response time(T(res)), etc. [6,7]. Hence, it is worthwhile to develop highly sensitive, highly selective gas sensors that can effectively monitor n-butanol. A variety of gas sensors have been developed, including semiconductor metal oxide gas sensors, primary battery-type oxygen sensors, and catalytic combustion-type sensors [8].
Among semiconductor metal oxides, the most extensively applied gas-sensitive materials at present, n-type semiconductors are the most studied, such as SnO2, TiO2, Fe2O3, In2O3, LaFeO3, WO3, and ZnO, and the gas-sensitive devices made of such materials integrate the merits of low cost, strong stability, easy manufacturing, and convenient use [9,10,11]. Nevertheless, single-metal-oxide gas sensors can no longer meet operational needs due to their high operating temperature and low sensitivity [12]. As a typical n-type metal oxide semiconductor, zinc stannate (ZnSnO3), which can be easily prepared and conveniently used at a low cost, has been used to prepare sensor elements to detect all kinds of volatile organic gases [13]. The gas-sensing properties of ZnSnO3 with different morphologies have been explored, including ZnSnO3 cubes, porous ZnSnO3 nanospheres, and ZnSnO3 nanorods [14]. Zeng et al. [15] successfully synthesized ZnSnO3 nanotubes through the hydrothermal method; Mahmood et al. [16] prepared NiO/ZnSnO3 composites using the electrospinning technique; Zhou et al. [17] prepared hollow ZnSnO3 through the co-precipitation method. Among these ZnSnO3 materials of differing morphology, porous ZnSnO3 nanospheres have aroused the most attention, since porous structures feature a large specific surface area and more activated adsorption sites, making them quite conducive to gas diffusion and mass transfer and facilitating oxygen adsorption, qualities that make ZnSnO3 strongly gas-responsive. However, pure ZnSnO3 gas sensors are prone to the same defects as other semiconductor metal oxide gas sensors, e.g., poor selectivity, weak response, and high operating temperature. Hence, initial efforts have been made to improve the gas-sensing properties of pure ZnSnO3, including doping and compounding with other semiconductor materials.
With an ionic radius of 1.03 Å, which differs greatly from that of Sn (0.69 Å) and Zn (0.74 Å), bismuth (Bi) is very prone to cause lattice distortion if doped into the ZnSnO3 matrix, leading to many crystal defects and further enhancing the gas-sensing properties of ZnSnO3. Mutkule et al. [18] synthesized Bi3+-doped spinel cobalt ferrite, a gas-sensitive material; Cai et al. [19] prepared flower-shaped Bi-doped rGO/Co3O4 nanohybrids; Ma et al. [20] prepared porous Bi-doped SnO2 nanosheets via electrospinning. Therefore, it is feasible to enhance the gas-sensing properties of ZnSnO3-based gas sensors by doping Bi into ZnSnO3.
In this study, pure ZnSnO3 and Bi-doped ZnSnO3 composites (Bi-ZnSnO3) were synthesized via in situ precipitation. Next, the prepared pure ZnSnO3 and Bi-ZnSnO3 were characterized by X-ray diffractometry (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), a Brunauer–Emmett–Teller (BET) analyzer, and X-ray photoelectron spectroscopy (XPS). Then, the gas sensing properties of the two materials to n-butanol were tested, including their optimal operating temperature, sensitivity, T(res)/recovery time T(rec), selectivity, and repeatability.

2. Experiments and Methods

2.1. Materials and Reagents

The reagents used included absolute ethanol, SnCl4•5H2O, Bi(NO3)3•6H2O, and NaOH (purchased from Tianjin Chemical Reagent Factory), as well as Zn(NO3)2•7H2O (bought from Junli Chemical Production Factory). All being analytically pure, the chemical reagents were directly used in the experiments without further purification.

2.2. Material Characterization

SEM (TESCAN MIRA, Brno, Czech Republic) was done to observe material microstructures, before which the test materials were coated with a thin layer of an electrically conducting material to ensure sufficient conductivity. Material microstructures were also observed via TEM (FEI Talos F200X, FEI, Hillsboro, OR, USA), and the crystal structures of samples were analyzed with XRD (D8 ADVANCE/AXS, Bruker, Saarbrücken, Germany). The specific surface area and the average pore size of the test materials were analyzed with the BET analyzer (ASAP-2460, Micromeritics Instrument Corp., Norcross, GA, USA). Then, the chemical element composition and the compound structure in the test materials were analyzed by XPS (K-Alpha, Thermo Fisher Scientific, Waltham, MA, USA). The gas sensing properties of the test materials, such as their sensitivity, operating temperature, T(res)/T(rec), selectivity, and repeatability, were mainly tested via an intelligent gas sensing analysis system (CGS-8, Beijing GIC-Tech Corp., Ltd., Beijing, China).

2.3. Preparation of ZnSnO3 and Bi-ZnSnO3

ZnSnO3 was prepared through an in situ precipitation method. First, ethanol aqueous solution (33.3 wt%) was prepared. Next, 4 mmol SnCl4•5H2O and 4 mmol Zn(NO3)2•7H2O were dispersed into 64 mL of an ethanol aqueous solution, followed by magnetic stirring for 30 min to form the uniform suspension marked as solution A. Afterwards, 48 mmol NaOH was added to a beaker containing 48 mL of deionized water and completely dissolved, forming solution B, which was then slowly added into solution A to be subjected to magnetic stirring for 30 min. Then, 120 mmol NaOH was dissolved in 32 mL of ethanol aqueous solution to form a clear solution, which was added dropwise into the abovementioned mixed solution. After stirring for 5 min, Bi(NO3)3•6H2O with different mass fractions was added at the doping concentration of 0 wt%, 2 wt%, 4 wt%, 5 wt%, or 7 wt%, followed by magnetic stirring for 30 min. It was transferred to a 500 mL round-bottom flask for refluxing at 85 °C for 3 h and left to stand for 12 h. The precipitates were filtered, rinsed, and dried. The obtained powder materials were heated in a 450 °C tube furnace at a heating rate of 5 °C/min for 3 h, yielding Bi-ZnSnO3 with different Bi contents (pure ZnSnO3 was acquired at the Bi doping concentration of 0 wt%). The synthesis process of Bi-ZnSnO3 is displayed in Figure 1.
Under alkaline conditions, Zn2+, Sn4+, and Bi were doped to form the Bi-ZnSn(OH)6 precursor (Formula (1)). Then, Bi-ZnSn(OH)6 was calcinated at a high temperature, and crystals were removed to form Bi-ZnSnO3 (Formula (2)).
  Zn 2 + + Sn 4 + + Bi Bi ZnSn OH   6
Bi ZnSn OH   6 Bi ZnSnO 3

2.4. Preparation of Sensor Device

First, a small amount of the ZnSnO3 and Bi-ZnSnO3 samples were weighed and uniformly ground in a mortar. Second, a moderate amount of absolute ethanol was added into the ground samples to dissolve them into pastes, which were uniformly coated on the outer surface of the gas sensor. Next, the sensor was placed in an 80 °C vacuum-drying oven for 12 h, followed by aging in a 200 °C aging system for 12 h. Finally, the gas-sensing properties of the aging sensor were explored with air-sensing test equipment. The sensitivity of gas sensitive materials was defined as follows:
S = Ra / Rg
where Ra is the resistance (Ω) of the gas-sensing material in air and Rg is the resistance (Ω) of the gas-sensing material in the gas under test.

3. Results and Discussion

3.1. XRD

The XRD patterns of ZnSnO3 and Bi-ZnSnO3 are exhibited in Figure 2. It was clearly observed that pure ZnSnO3 had two weak and wide diffraction peaks at 2θ = 27.25° and 58.59°, indicating that amorphous ZnSnO3 formed after the heat treatment [21]. This was because the calcination and dehydration of the precursor ZnSn(OH)6 destroyed the original H–O bond and changes the internal lattice arrangement, thereby forming amorphous ZnSnO3 [22]. In comparison with the patterns of pure ZnSnO3, new intense characteristic peaks could be observed in Bi-ZnSnO3 at 2θ = 27.47°, 33.16°, 35.01°, 42.42°, 46.47°, 52.38°, and 55.11°, which corresponded to the crystal faces of (021), (410), (240), (241), (431), (540), and (611), respectively. This result coincided with the standard data card (PDF#17-0320) of Bi, indicating that these newly generated diffraction peaks resulted from Bi doping. With the increase in the doping concentration of Bi, the intensity of diffraction peaks rose substantially, which further proved the successful doping of Bi.

3.2. SEM

The SEM images of pure ZnSnO3 and Bi-ZnSnO3 are presented in Figure 3. It can be seen from the SEM image (Figure 3a) of ZnSnO3 that pure ZnSnO3 had a spherical structure with a coarse appearance and many holes on the surface. From the gaps of broken holes, the pure ZnSnO3 was hollowly structured, with a large pore volume and a diameter of about 900 nm, and each ZnSnO3 block consisted of many tiny, uniformly distributed nanoparticles. As shown in Figure 3b, Bi-doped ZnSnO3 had a similar appearance, namely, it remained a spherical hollow porous structure. The element distribution diagram was mainly used to study the composition and concentration of each component in the test materials. As further verified by Figure 3c–g, four elements, O, Bi, Zn, and Sn, mainly existed in Bi-ZnSnO3, among which Sn, Zn, and O were distributed quite uniformly, though the element distribution does not seem obvious in the figures due to too small doping concentrations of Bi. The mass ratios and the atomic percentages of the four elements are listed in Table 1. It can be intuitively seen that the atomic percentages of O, Bi, Zn, and Sn were 47.54%, 3.43%, 20.44%, and 28.58%, respectively. Evidently, the content of Bi was much lower than that of the other three elements.

3.3. TEM

The microstructures of pure ZnSnO3 and Bi-ZnSnO3 were further explored via TEM. The TEM images of pure ZnSnO3 are displayed in Figure 4a,b. The spherically structured ZnSnO3 was 1.92 nm in size and was formed by the aggregation of many particles. In addition, the hollow structure of the sample clearly had many holes on it, in agreement with the SEM characterization results. As observed from the TEM image (Figure 4c) of Bi-ZnSnO3, the sample did change much morphologically, keeping a spherical hollow porous structure with a size of 1.98 nm. The electronic diffraction pattern is illustrated in Figure 4c, presenting regularly arranged diffraction spots. Given this, Bi-ZnSnO3 was judged to be a single-crystal structure within the range of selection.

3.4. BET

Specific surface area is one of the important parameters for measuring gas-sensitive materials. In this study, the specific surface area of prepared materials was determined through the N2 adsorption–desorption method. The N2 adsorption–desorption isothermal diagram of pure ZnSnO3 and Bi-ZnSnO3 is displayed in Figure 5a,b. According to the classification of International Union of Pure and Applied Chemistry (IUPAC), both materials had a type IV isothermal forming an H3 hysteresis loop, which is the characteristic isothermal of mesoporous materials [23,24]. The BET specific surface areas of pure ZnSnO3 and Bi-ZnSnO3 were 10.01 m2/g and 13.75 m2/g, respectively. The pore volume and the average pore size of pure ZnSnO3 were 0.02 cm3/g and 9.26 nm, and those of Bi-ZnSnO3 were 0.03 cm3/g and 25.51 nm. The pores might have been generated by the gaps generated in the formation of ZnSnO3 nanospheres. Furthermore, the mesoporous structure of microspheres and the enlargement of their specific surface area facilitated the gas adsorption and further enhanced the gas transport efficiency [25]. Therefore, we hypothesize that the changes in the specific surface area, the pore volume, and the pore size of Bi-ZnSnO3 exerted important influences on the gas sensitivity of this material.

3.5. XPS

The surface chemical composition and element distribution were analyzed via XPS. The XPS survey spectrum of Bi-ZnSnO3 is displayed in Figure 6a. Characteristic peaks of C, Zn, Sn, O, and Bi existed in the composites, indicating the presence of these five elements in the samples. The high-resolution pattern of Bi4f is exhibited in Figure 6b, where the peaks at the binding energy of 159.08 eV and 164.08 eV corresponded to Bi4f7/2 and Bi4f5/2 of Bi, respectively. The high-resolution patterns of Zn2p in pure ZnSnO3 and Bi-ZnSnO3 are shown in Figure 6c. The binding energies of Zn2p3/2 and Zn2p1/2 in Bi-ZnSnO3 were 1021.68 and 1044.78 eV, respectively, which showed a shift of 0.3 eV in comparison with the binding energy (1021.98 eV and 1045.08 eV) in pure ZnSnO3. This may be because Bi doping changed the electron density on the ZnSnO3 surface and reduced the binding energy. The high-resolution patterns of Sn3d in pure ZnSnO3 and Bi-ZnSnO3 are displayed in Figure 6d. The binding energies of Sn3d5/2 and Sn3d3/2 in Bi-ZnSnO3 were located at 486.38 and 494.78 eV, respectively, showing a shift of 0.2 eV compared with the binding energy (486.58 eV and 494.98 eV) in pure ZnSnO3, another change that we attributed to Bi doping. In addition, the energy gap between two corresponding peaks of Sn3d was 8.4 eV, indicating that Sn existed in the form of +4 valence [26]. The high-resolution patterns of O1s in pure ZnSnO3 and Bi-ZnSnO3 are displayed in Figure 6e,f, where the O1s peak could be divided into three characteristic peaks: lattice oxygen, oxygen vacancies, and chemi-adsorbed oxygen. From Figure 6e, the three characteristic peaks of pure ZnSnO3 were successively located at 530.18, 531.48, and 532.48 eV, respectively, with area ratios of 72.28%, 21.68%, and 6.04%, respectively. As shown in Figure 6f, the three characteristic peaks of Bi-ZnSnO3 were located at 530.38, 531.68, and 532.58 eV, with area ratios of 67.34%, 26.10%, and 6.56%, respectively. It is widely accepted that the higher the proportions of oxygen vacancies and chemi-adsorbed oxygen are, the better the gas sensing properties of the sensor will be [27,28]. In this study, the total proportion of oxygen vacancies and chemi-adsorbed oxygen in Bi-ZnSnO3 was 32.66%, which was higher than 27.72% in pure ZnSnO3, further proving that Bi replaced the lattice atoms in pure ZnSnO3 so as to form oxygen defects.

3.6. The Working Temperature of Bi-ZnSnO3

In general, the properties of metal-oxide gas sensors are closely related to their operating temperature [29]. Too high of an operating temperature not only increases the wear rate of gas-sensitive elements but also raises their application and production costs, so a significant research objective in the field of metal-oxide gas sensors is to reduce their operating temperature [30]. Figure 7 shows the temperature response curves of pure ZnSnO3 and Bi-ZnSnO3 of different mass ratios to 100 ppm n-butanol gas. Under normal circumstances, the sensitivity of gas sensors would first grow and then decline with the rise in ambient temperature. This is because both the adsorption rate and the desorption rate are accelerated by a rise in the ambient temperature, but the former is faster than the latter, thus substantially enhancing the material sensitivity. As the ambient temperature further rises after reaching a certain level, the acceleration of the desorption rate exceeds that of the adsorption rate, thus lowering the sensitivity of the gas-sensitive material [31,32]. First, the temperature response of Bi-ZnSnO3 at the doping concentrations of 0 wt%, 2 wt%, 4 wt%, 5 wt%, and 7 wt% was tested. It can be clearly observed from the temperature response diagram that the five composites operated optimally at 350 °C, 375 °C, 300 °C, 300 °C, and 300 °C, corresponding to the respective sensitivities of 173.25, 339.56, 1450.65, 60.71, and 40.81. Thus, the Bi-ZnSnO3 reached the highest sensitivity at the doping concentration of 4 wt%, being 8.37 times that of pure ZnSnO3. Moreover, the optimal operating temperature of the composite was 300 °C, 50 °C lower than that of pure ZnSnO3. The above results show that the optimal proportion of Bi is doped in ZnSnO3 can effectively improve the gas-sensing properties of the matrix material ZnSnO3 in addition to lowering the optimal operating temperature.

3.7. The Sensitivity Performance of Bi-ZnSnO3 for N-Butanol

As shown in Figure 8a, the sensitivities of ZnSnO3 and Bi-ZnSnO3 with different contents of Bi increased to various extents as the concentration of n-butanol gas increased from 5 to 500 ppm, with Bi-ZnSnO3 (4 wt%) presenting the fastest growth rate in sensitivity, which indicated that the optimal doping concentration of Bi was 4 wt%. In addition, the growth rate of Bi-ZnSnO3 (4 wt%) first increased and then declined with the increase in the concentration of n-butanol gas. This was because a rise in the concentration of n-butanol brings about an increasing number of molecules that occupy a limited number of adsorption sites on the surface of the sensing material, which eventually become saturated and, thus, slow the growth in sensitivity. The concentration–response curves of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) to n-butanol gas are shown in Figure 8b,c, respectively. The sensitivity of pure ZnSnO3 to various concentrations of n-butanol gas was 14.01, 27.02, 93.82, 126.89, 151.29, 168.67, 271.28, 403.45, and 601.87, while that of Bi-ZnSnO3 to n-butanol gas at various concentrations was 82.76, 192.36, 435.06, 693.65, 1055.08, 1450.65, 1953.68, 2506.184, and 3226.02. It was clear that the sensitivity of the two gas-sensitive materials to n-butanol gas increased progressively with the n-butanol gas concentration. That is because more and more available n-butanol molecules participate in the redox reaction, contributing to the gradual increase in the sensitivity of the materials. At a given n-butanol gas concentration, the sensitivity of Bi-ZnSnO3 (4 wt%) was much higher than that of pure ZnSnO3 due to the modification of Bi. Additionally, the repeatability of the gas sensor was excellent, as the sensitivity increased rapidly after adsorbing n-butanol gas and quickly returned to the initial state when it was exposed to the air.
T(res)/T(rec) plays an important role in the evaluation of the properties of sensors [33,34]. T(res) is defined as the time it takes for a gas sensor to reach 90% of its maximum sensitivity to the measured gas, while T(rec) stands for the time it takes for the resistance of a gas sensor to reach 90% of its resistance variation when the sensor is away from the measured gas. We drew the response curves of the two materials to 100 ppm n-butanol gas at their respective optimal operating temperatures (Figure 8d). The T(res) and T(rec) of pure ZnSnO3 were 10 s and 17.5 s, while those of Bi-ZnSnO3 were 8 s and 17 s. By comparison, the T(res) and T(rec) of Bi-ZnSnO3 (4 wt%) were significantly shorter than those of ZnSnO3, meaning that the response and recovery performance of the composites was significantly higher, which in turn improved the application value of the materials.
The linear relationship between the sensitivity of Bi-ZnSnO3 and the concentration of n-butanol gas (5–500 ppm) is shown in Figure 9, from which it can be seen that, with the increase in gas concentration, the sensitivity of Bi-ZnSnO3 (4 wt%) composites increased rapidly in the n-butanol environment with a concentration of less than 100 ppm. When the n-butanol concentration exceeded 100 ppm, the sensitivity of the composites increased rapidly. This indicates that the detection limit of the material gradually approaches a peak, so it may be necessary to modify the material in other ways to better detect higher concentrations of n-butanol gas. The two-stage linear correlation coefficients of the Bi-ZnSnO3 (4 wt%) composite are R12 = 0.9921and R22 = 0.9838, respectively. The results show that the composite can accurately detect n-butanol gas in the range of 5 ppm–500 ppm in practical application.

3.8. Sensitivity of ZnSnO3 and Bi-ZnSnO3 to Different VOCs

Its good selectivity can be attributed to the different adsorption capacities and reducibilities of the sensor surface for various measured gases. To further evaluate the selectivity of the two sensor materials, the sensitivities of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) gas-sensitive elements to six different volatile organic compounds at 100 ppm were measured at their optimal operating temperatures. The results are shown in Figure 10a. The sensitivity of Bi-ZnSnO3 (4 wt%) to n-butanol gas reached 1450.65, which was 35.37 times that (41.01) of ammonia gas, 2.93 times that (495.09) of acetone gas, 6.02 times that (241.05) of methanol gas, 2.54 times that (571.48) of formaldehyde gas, and 2.98 times that (486.58) of ethanol gas. By comparison, the sensitivity of Bi-ZnSnO3 (4 wt%) to n-butanol was much higher than that of other gases, demonstrating a great improvement over pure ZnSnO3. Additionally, Bi-ZnSnO3 presented excellent selectivity for n-butanol gas, probably because Bi doping increased the specific surface area of ZnSnO3 and gave rise to many active sites on the surfaces, thus improving its gas-sensing properties.
The response values of Bi-ZnSnO3 (4 wt%) to a gas concentration of 100 ppm at 300 °C were tested repeatedly within 28 days (four times, once a week) (Figure 10b). The sensitivity obtained from the first test was 1450.65, while that of the fourth test was 1431.60, which indicated that the sensitivity of the composites decreased slightly as the test progressed. Even so, the sensitivity remained above 95% of the initial response value, showing that this gas-sensitive material has good stability for repeated use.
The performance parameters of the Bi-ZnSnO3 gas sensor to n-butanol are compared with those of other gas sensors in Table 2. Despite a higher operating temperature, the Bi-ZnSnO3 gas sensor showed a much better sensitivity to n-butanol gas at 100 ppm than other gas sensors based on different gas-sensitive materials. The optimal operating temperature of the ZnSnO3 sensor was reduced to 300 °C due to the doping of Bi. All these findings indicate that Bi-ZnSnO3, serving as coating materials for gas sensors, is significantly valuable in research and in real-world applications.

3.9. The Sensing Mechanism

The adsorption and desorption of oxygen molecules on the surfaces of n-type semiconductors lead to resistance changes, so the gas-sensing properties of n-type semiconductors were compared [41]. When ZnSnO3 is exposed to air, oxygen molecules in the air will be adsorbed on its surface to form chemi-adsorbed oxygen, generating ionized oxygen O 2 ,   O , and O 2 . As the temperature increases, oxygen molecules capture free electrons from the conduction band of ZnSnO3, resulting in a decrease in electron concentration and the formation of a depletion layer, which increases the resistance of the gas sensor [42]. Hence, the sensitive mechanism of n-butanol detection can be illustrated by the Wolkenstein model [43], as follows:
O 2 gas O 2 ( ads )
O 2 ads + e O 2 ( ads )
O 2 ads + e 2 O ( ads )
O ads + e O 2 ( ads )
The reaction in Formulas (4) and (5) mainly occurred at a temperature less than 100 °C (T < 100 °C); the reaction in Formula (6) mainly occurred when 100 °C < T < 300 °C; and the reaction in Formula (7) mainly occurred when T > 300 °C. A redox reaction between n-butanol gas and ionized oxygen species is initiated by the contact of gas-sensitive elements and n-butanol gas, which releases the captured electrons back to the conduction band of ZnSnO3, reducing the thickness of the electron depletion layer, increasing the carrier concentration, and lowering the resistance of the gas-sensitive elements. According to the test results, the optimal operating temperatures of the prepared materials are above 300 °C, and the chemical reactions in this process are mainly as follows [44,45,46]:
C 4 H 9 OH + 12 O ads 4 CO 2 + 5 H 2 O + 12 e
C 4 H 9 OH + 12 O 2 ads 4 CO 2 + 5 H 2 O + 24 e
Combined with the XPS results, these results show that the total proportion of oxygen vacancies and chemi-adsorbed oxygen in Bi-ZnSnO3 (4 wt%) was from 27.72% to 32.68% higher than that of pure ZnSnO3. The increase in oxygen vacancies indicated that more electrons could be captured from the conduction band of Bi-ZnSnO3 while chemi-adsorbed oxygen directly participated in the reaction of n-butanol gas to release more electrons back to the conduction band of Bi-ZnSnO3, leading to a stronger change in the electron depletion layer. This suggests that the introduction of Bi is beneficial to the surface adsorption of oxygen. In addition, according to the BET characterization, the specific surface area of Bi-ZnSnO3 was also improved, which led to more active sites on the gas sensor surface. Finally, the gas-sensing properties of Bi-ZnSnO3 were promoted by a more complete electron depletion layer due to an increasing number of oxygen molecules adsorbed on the surface [47]. The gas-sensing mechanism of Bi-ZnSnO3 is shown in Figure 11.

4. Conclusions

ZnSnO3 and Bi-ZnSnO3 were first synthesized via the in situ precipitation method and then were characterized as nanospheres, followed by a study on the gas-sensing properties to n-butanol gas. The test results revealed that Bi-ZnSnO3, compared with pure ZnSnO3, was a superior sensor of n-butanol gas. In particular, Bi-ZnSnO3 (4 wt%) possessed the highest sensitivity of 1450.65, approximately 8.37 times that of pure ZnSnO3, at an optimal operating temperature of 300 °C, which was 50 °C lower than that of pure ZnSnO3. Moreover, Bi-ZnSnO3 had better selectivity and repeatability. The total proportion of oxygen vacancies and chemi-adsorbed oxygen in Bi-ZnSnO3 (4 wt%) was from 27.72% to 32.68% higher than that of pure ZnSnO3. The increase in oxygen vacancies indicated that more electrons could be captured from the conduction band of Bi-ZnSnO3, while chemi-adsorbed oxygen directly participated in the reaction of n-butanol gas to release more electrons back to the conduction band of Bi-ZnSnO3, leading to a stronger change in the electron depletion layer. Altogether, our results suggest that Bi-ZnSnO3 have great potential in the detection of n-butanol gas owing to its excellent gas-sensing properties, in contrast to traditional n-butanol sensors.

Author Contributions

Data curation, L.J.; investigation, R.Z.; project administration, L.J.; validation, Q.C. and W.Z.; writing—original draft, L.J.; writing—review & editing, Q.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the financial support from the National Natural Science Foundation of China (52064036), the Talents innovation and entrepreneurship project of Lanzhou (2021-RC-36), and Jieqing Project of Gansu Province.

Conflicts of Interest

The authors declare that there are no competing interests regarding the publication of this paper.

References

  1. Tong, Z.; Rui, Z.; Wang, Z.; Zhao, R.; Wang, Z.; Yue, Y.; Xing, X.; Wang, Y. Sensitive and selective n-butanol gas detection based on ZnO nanocrystalline synthesized by a low-temperature solvothermal method. Phys. E Low Dimens. 2018, 103, 143–150. [Google Scholar] [CrossRef]
  2. Feng, G.Q.; Che, Y.H.; Song, C.W.; Xiao, J.K.; Fan, X.F.; Sun, S.; Huang, G.H.; Ma, Y.C. Morphology-controlled synthesis of ZnSnO3 hollow spheres and their n-butanol gas-sensing performance. Ceram. Int. 2021, 47, 2471–2482. [Google Scholar] [CrossRef]
  3. Sun, L.; Guo, Y.; Hu, Y.; Pan, S.; Jiao, Z. Conductometric n-butanol gas sensor based on Tourmaline@ZnO hierarchical micro-nanostructures. Sens. Actuators B Chem. 2021, 337, 129793. [Google Scholar] [CrossRef]
  4. Xu, Y.; Zheng, L.; Yang, C.; Zheng, W.; Zhang, J. Chemiresistive sensors based on core-shell ZnO@TiO2 nanorods designed by atomic layer deposition for n- butanol detection. Sens. Actuators B Chem. 2020, 310, 127846. [Google Scholar] [CrossRef]
  5. Zhang, S.D.; Yang, M.J.; Liang, K.Y.; Zhang, B.X. An acetone gas sensor based on nanosized Pt-loaded Fe2O3 nanocubes. Sens. Actuators B Chem. 2019, 290, 59–67. [Google Scholar] [CrossRef]
  6. Cheng, C.L.; Che, P.Y.; Chen, B.Y.; Lee, W.J.; Chien, L.J.; Chang, J.S. High yield bio-butanol production by solvent-producing bacterial microflora. Bioresour. Technol. 2012, 113, 58–64. [Google Scholar] [CrossRef]
  7. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconductor metal oxides as sensors for environmentally hazardous gases. Sens. Actuators B Chem. 2011, 160, 580–591. [Google Scholar] [CrossRef]
  8. Kolmakov, A.; Moskovits, M. Chemical Sensing and Catalysis by One-Dimensional Metal-Oxide Nanostructures. ChemInform 2004, 34, 151–180. [Google Scholar] [CrossRef]
  9. Bochenkov, V.E.; Sergeev, G.B. Sensitivity, Selectivity, and Stability of Gas-Sensitive Metal-Oxide Nanostructures; American Scientific Publishers: Valencia, CA, USA, 2010; pp. 31–52. [Google Scholar]
  10. Qin, J.; Cui, Z.; Yang, X.; Zhu, S.; Li, Z.; Liang, Y. Synthesis of three- dimensionally ordered macroporous LaFeO3 with enhanced methanol gas sensing properties. Sens. Actuators B Chem. 2015, 209, 706–713. [Google Scholar] [CrossRef]
  11. Ramgir, N.S.; Ganapathi, S.K.; Kaur, M.; Datta, N.; Muthe, K.P.; Aswal, D.K.; Gupta, S.K.; Yakhmi, J.V. Sub-ppm H2S sensing at room temperature using CuO thin films. Sens. Actuators B Chem. 2010, 151, 90–96. [Google Scholar] [CrossRef]
  12. Parmeshwar, W.; Dipak, B.; Pradip, P. High performance H2 sensor based on ZnSnO3 cubic crystallites synthesized by a hydrothermal method. Talanta 2013, 105, 327–332. [Google Scholar] [CrossRef]
  13. Zheng, J.; Hou, H.; Fu, H.; Gao, L.; Liu, H. Size-controlled synthesis of porous ZnSnO3 nanocubes for improving formaldehyde gas sensitivity. RSC Adv. 2021, 11, 20268–20277. [Google Scholar] [CrossRef]
  14. Song, X.Y.; Jia, H.J.; Tian, X.H.; Zhang, M.G.; Dai, Z. Highly sensitive formaldehyde chemical sensor based on in situ precipitation synthesis of ZnSnO3 microspheres. J. Mater. Sci. Mater. Electron. 2015, 26, 6224–6231. [Google Scholar] [CrossRef]
  15. Zeng, Y.; Zhang, K.; Wang, X.; Sui, Y.; Zou, B.; Zheng, W.; Zou, G. Rapid and selective H2S detection of hierarchical ZnSnO3 nanocages. Sens. Actuators B Chem. 2011, 15, 245–250. [Google Scholar] [CrossRef]
  16. Haq, M.U.; Zhang, Z.Y.; Chen, X.H.; Nasir, R. A two-step synthesis of microsphere-decorated fibers based on NiO/ZnSnO3 composites towards superior ethanol sensitivity performance. J. Alloy Compd. 2019, 777, 73–83. [Google Scholar] [CrossRef]
  17. Wang, L.; Zhang, R.; Zhou, T.T.; Lu, G.Y.; Lou, Z. Highly sensitive sensing platform based on ZnSnO3 hollow cubes for detection of ethanol. Appl. Surf. Sci. 2017, 400, 262–268. [Google Scholar] [CrossRef]
  18. Mutkule, S.U.; Tehare, K.K.; Gore, S.K.; Gunturu, K.C.; Mane, R.S. Ambient temperature Bi-Co ferrite NO2 sensors with high sensitivity and selectivity. Mater. Res. Bull. 2019, 115, 150–158. [Google Scholar] [CrossRef]
  19. Cai, S.X.; Song, X.Q.; Chi, Z.T.; Fu, Y.Q.; Fang, Z.T.; Geng, S.Y.; Kang, Y.R.; Yang, X.X.; Qin, J.F.; Xie, W.F. Rational design of Bi-doped rGO/Co3O4 nanohybrids for ethanol sensing. Sens. Actuators B Chem. 2021, 343, 118–130. [Google Scholar] [CrossRef]
  20. Ma, Z.; Yang, K.; Xiao, C.; Jia, L. Electrospun Bi-doped SnO2 Porous Nanosheets for Highly Sensitive Nitric Oxide Detection. J. Hazard. Mater. 2021, 416, 118–126. [Google Scholar] [CrossRef]
  21. Xie, Q.; Ma, Y.; Zhang, X.; Guo, H.; Lu, A.; Wang, L.; Yue, G.; Peng, D.L. Synthesis of amorphous ZnSnO3-C hollow microcubes as advanced anode materials for lithium ion batteries. Electrochim. Acta. 2014, 141, 374–383. [Google Scholar] [CrossRef]
  22. Yin, Y.Y.; Li, F.; Zhang, N.; Ruan, S.P.; Chen, H.F. Improved gas sensing properties of Silver-functionalized ZnSnO3 hollow nanocubes. Inorg. Chem. Front. 2018, 5, 2123–2131. [Google Scholar] [CrossRef]
  23. Michal, K.; Mietek, J. Gas Adsorption Characterization of Ordered Organic-Inorganic Nanocomposite Materials. Chem. Mater. 2001, 13, 3169–3183. [Google Scholar] [CrossRef]
  24. Sun, G.; Chen, H.L.; Li, Y.W.; Ma, G.Z. Synthesis and triethylamine sensing properties of mesoporous α-Fe2O3 microrods. Mater. Lett. 2016, 178, 213–216. [Google Scholar] [CrossRef]
  25. Jia, T.; Jian, C.; Zhao, D.; Fang, F.; Zhao, J.; Wang, X.; Fei, L. Facile synthesis of Zn-doped SnO2 dendrite-built hierarchical cube-like architectures and their application in lithium storage. J. Mater. Sci. Eng. B 2014, 189, 32–37. [Google Scholar] [CrossRef]
  26. Han, X.G.; He, H.Z.; Kuang, Q.; Zhou, X.; Zhang, X.H. Controlling Morphologies and tuning the related properties of Nano/Microstructured ZnO crystallites. J. Phys. Chem. C 2009, 113, 584–589. [Google Scholar] [CrossRef]
  27. Liu, J.; Wang, T.; Wang, B.; Sun, P.; Yang, Q.; Liang, X.; Song, H.; Lu, G. Highly sensitive and low detection limit of ethanol gas sensor based on hollow ZnO/SnO2 spheres composite material. Sens. Actuators B Chem. 2017, 245, 551–559. [Google Scholar] [CrossRef]
  28. Huang, J.; Wang, L.; Gu, C.; Shim, J.J. Preparation of hollow porous SnO2 microcubes and their gas-sensing property. Mater. Lett 2014, 136, 371–374. [Google Scholar] [CrossRef]
  29. Yu, H.; Li, J.; Tian, Y.; Li, Z. Environmentally friendly recycling of SnO2/Sn3O4 from tin anode slime for application in formaldehyde sensing material by Ag/Ag2O modification. J. Alloy Compd. 2018, 765, 624–634. [Google Scholar] [CrossRef]
  30. Sharma, A.; Tomar, M.; Gupta, V. A low temperature operated NO2 gas sensor based on TeO2/SnO2 p-n heterointerface. Sens. Actuators B Chem. 2013, 176, 875–883. [Google Scholar] [CrossRef]
  31. Liu, X.D.; Liu, J.Y.; Liu, Q.; Chen, R.R.; Zhang, H.S.; Yua, J.; Song, D.L.; Li, J.Q.; Zhang, M.L.; Wang, J. Template-free synthesis of rGO decorated hollow Co3O4 nano/microspheres for ethanol gas sensor. Ceram. Int. 2018, 44, 21091–21098. [Google Scholar] [CrossRef]
  32. Ge, S.; Zheng, H.; Sun, Y.; Jin, Z.; Shan, J.; Wang, C.; Wu, H.; Li, M.; Meng, F. Ag/SnO2/graphene ternary nanocomposites and their sensing properties to volatile organic compounds. J. Alloy Compd. 2015, 659, 127–131. [Google Scholar] [CrossRef]
  33. Wen, Z.; Liu, T.M. Gas-sensing properties of SnO2-TiO2-based sensor for volatile organic compound gas and its sensing mechanism. Phys. B Condens. Matter 2010, 405, 1345–1348. [Google Scholar] [CrossRef]
  34. Lee, E.J.; Lee, D.H.; Yoon, J.; Yin, Y.L.; Lee, Y.N.; Uprety, S.; Yoon, Y.S.; Kim, D.J. Enhanced Gas-Sensing Performance of GO/TiO2 Composite by Photocatalysis. J. Sensors 2018, 18, 3334. [Google Scholar] [CrossRef]
  35. Xu, Y.; Liu, P.; Sun, D.; Sun, Y.; Zhang, G.; Gao, D. Tunable synthesis of uniform ZnO nanospheres and their size-dependent gas sensing performance toward n-butanol. Mater. Lett. 2015, 161, 495–498. [Google Scholar] [CrossRef]
  36. Liu, W.P.; Zhang, X.M.; Wang, Z.F.; Wang, R.J.; Chen, C.; Dong, C.J. Nanoparticles Assembled CdIn2O4 Spheres with High Sensing Properties towards n-butanol. J. Nanomater. 2019, 9, 1714. [Google Scholar] [CrossRef]
  37. Huang, J.; Min, Y.; Gu, C.; Zhai, M.; Sun, Y.; Liu, J. Hematite solid and hollow spindles: Selective synthesis and application in gas sensor and photocatalysis. Mater. Res. Bull. 2011, 46, 1211–1218. [Google Scholar] [CrossRef]
  38. Malik, R.; Tomer, V.K.; Chaudhary, V.; Dahiya, M.S.; Nehra, S.P.; Rana, P.S.; Duhan, S. Ordered mesoporous In-(TiO2/WO3) nanohybrid: An ultrasensitive n-butanol sensor. Sens. Actuators B Chem. 2017, 239, 364–373. [Google Scholar] [CrossRef]
  39. Yang, B.X.; Liu, J.Y.; Qin, H.; Liu, Q.; Jing, X.Y.; Zhang, H.Q.; Li, R.M.; Huang, G.Q.; Wang, J. PtO2-nanoparticles functionalized CuO polyhedrons for n-butanol gas sensor application. Ceram. Int. 2018, 44, 10426–10432. [Google Scholar] [CrossRef]
  40. Jian, W.; Zhi, Z.; An, D.; Tong, X.; Zhou, Q. Highly selective n-butanol gas sensor based on porous In2O3 nanoparticles prepared by solvothermal treatment. Mater. Sci. Semicond. Process. 2018, 83, 139–143. [Google Scholar] [CrossRef]
  41. Husko, C.; Rossi, A.D.; Combrie, S.; Tran, Q.V.; Raineri, F.; Wong, C.W. Ultrafast all-optical modulation in GaAs photonic crystal cavities. Appl. Phys. Lett. 2009, 94, 021111. [Google Scholar] [CrossRef]
  42. Liu, X.; Chen, N.; Xing, X.X.; Li, Y.X.; Xiao, X.C.; Wang, Y.D. A high-performance n-butanol gas sensor based on ZnO nanoparticles synthesized by a low-temperature solvothermal route. RSC Adv. 2015, 5, 54372–54378. [Google Scholar] [CrossRef]
  43. Chen, T.; Liu, Q.J.; Zhou, Z.L.; Wang, Y.D. A high sensitivity gas sensor for formaldehyde based on CdO and In2O3 doped nanocrystalline SnO2. Nanotechnology 2008, 19, 095506. [Google Scholar] [CrossRef] [PubMed]
  44. Yin, L.L.; Zhao, M.; Hu, H.L.; Ye, J.H.; Wang, D.F. Synthesis of graphene/tourmaline/TiO2 composites with enhanced activity for photocatalytic degradation of 2-propanol. J. Catal. 2017, 38, 1307–1314. [Google Scholar] [CrossRef]
  45. Qi, S.Y.; Wu, C.; Wang, D.P.; Xu, H.Y. The effect of Schorl on the photocatalytic properties of the TiO2/Schorl composite materials. Results Phys. 2017, 7, 3645–3647. [Google Scholar] [CrossRef]
  46. Ngo Thi, Y.L.; Hur, S.H. Low-temperature NO2 gas sensor fabricated with NiO and reduced graphene oxide hybrid structur. Mater. Res. Bull. 2016, 84, 168–176. [Google Scholar] [CrossRef]
  47. Wang, M.; Shen, Z.; Zhao, X.; Duanmu, F.; Yu, H.; Ji, H. Rational shape control of porous Co3O4 assemblies derived from MOF and their structural effects on n-butanol sensing. J. Hazard. 2019, 371, 352–361. [Google Scholar] [CrossRef]
Figure 1. Synthesis process of Bi-ZnSnO3 composites.
Figure 1. Synthesis process of Bi-ZnSnO3 composites.
Sensors 22 06571 g001
Figure 2. XRD patterns of ZnSnO3 and Bi-ZnSnO3.
Figure 2. XRD patterns of ZnSnO3 and Bi-ZnSnO3.
Sensors 22 06571 g002
Figure 3. (a) SEM image of ZnSnO3, (b) Bi-ZnSnO3, and (cg) elemental distribution of Bi-ZnSnO3.
Figure 3. (a) SEM image of ZnSnO3, (b) Bi-ZnSnO3, and (cg) elemental distribution of Bi-ZnSnO3.
Sensors 22 06571 g003
Figure 4. TEM image of pure ZnSnO3 (a,b) and TEM image of Bi-ZnSnO3 (4 wt%) (c); inset is the SAED image (c).
Figure 4. TEM image of pure ZnSnO3 (a,b) and TEM image of Bi-ZnSnO3 (4 wt%) (c); inset is the SAED image (c).
Sensors 22 06571 g004
Figure 5. Nitrogen adsorption–desorption isotherms of ZnSnO3 (a) and Bi-ZnSnO3 (4 wt%) composites (b); the inset is the BJH pore size distribution curve.
Figure 5. Nitrogen adsorption–desorption isotherms of ZnSnO3 (a) and Bi-ZnSnO3 (4 wt%) composites (b); the inset is the BJH pore size distribution curve.
Sensors 22 06571 g005
Figure 6. XPS full spectrum of Bi-ZnSnO3 (4 wt%) composite (a), high-resolution spectrum of Bi4f (b), high-resolution spectrum of Zn2p (c), high-resolution spectrum of Sn3d (d), and O1s spectra of ZnSnO3 and Bi-ZnSnO3 (4 wt%) composites (e,f).
Figure 6. XPS full spectrum of Bi-ZnSnO3 (4 wt%) composite (a), high-resolution spectrum of Bi4f (b), high-resolution spectrum of Zn2p (c), high-resolution spectrum of Sn3d (d), and O1s spectra of ZnSnO3 and Bi-ZnSnO3 (4 wt%) composites (e,f).
Sensors 22 06571 g006
Figure 7. Temperature response curves of pure ZnSnO3 and different ratios of Bi-ZnSnO3 to 100 ppm n-butanol gas.
Figure 7. Temperature response curves of pure ZnSnO3 and different ratios of Bi-ZnSnO3 to 100 ppm n-butanol gas.
Sensors 22 06571 g007
Figure 8. The relationship between the sensitivity of different materials and the concentration of n-butanol gas (a), the concentration–response curve of pure ZnSnO3 material to n-butanol gas at 325 °C (b), the concentration–response curve of Bi-ZnSnO3 to n-butanol gas at 300 °C (c), and the response recovery curve of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) materials to 100 ppm n-butanol gas (d).
Figure 8. The relationship between the sensitivity of different materials and the concentration of n-butanol gas (a), the concentration–response curve of pure ZnSnO3 material to n-butanol gas at 325 °C (b), the concentration–response curve of Bi-ZnSnO3 to n-butanol gas at 300 °C (c), and the response recovery curve of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) materials to 100 ppm n-butanol gas (d).
Sensors 22 06571 g008
Figure 9. Linear relationship between material sensitivity and n-butanol concentration of Bi-ZnSnO3 (4 wt%).
Figure 9. Linear relationship between material sensitivity and n-butanol concentration of Bi-ZnSnO3 (4 wt%).
Sensors 22 06571 g009
Figure 10. (a) Comparison of selectivity test results of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) and (b) repeatability test results of Bi-ZnSnO3 (4 wt%).
Figure 10. (a) Comparison of selectivity test results of pure ZnSnO3 and Bi-ZnSnO3 (4 wt%) and (b) repeatability test results of Bi-ZnSnO3 (4 wt%).
Sensors 22 06571 g010
Figure 11. Gas−sensing mechanism of Bi-ZnSnO3 composites.
Figure 11. Gas−sensing mechanism of Bi-ZnSnO3 composites.
Sensors 22 06571 g011
Table 1. Elemental mass ratio and atomic percentage of Bi-ZnSnO3 (4 wt%).
Table 1. Elemental mass ratio and atomic percentage of Bi-ZnSnO3 (4 wt%).
ElementWeight%Atomic%
O13.1847.54
Bi12.433.43
Zn42.0320.44
Sn32.3728.58
Table 2. Performance comparison of n-butanol gas sensor.
Table 2. Performance comparison of n-butanol gas sensor.
MaterialWorking Temperature (°C)Gas Concentration (ppm)Response Time/Recovery Time (s)SensitivityReference
ZnO3401006/17136.00[35]
CdIn2O428001004/1081.20[36]
α-Fe2O32801005/513.90[37]
In-TiO2/WO3200502.2/3127.00[38]
PtO2/CuO1801002.4/5.111.55[39]
In2O314010045/65241.00[40]
Bi-ZnSnO33001008/171450.65this paper
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, L.; Cui, Q.; Zhang, R.; Zhang, W. Highly Sensing and Selective Performance Based on Bi-Doped Porous ZnSnO3 Nanospheres for Detection of n-Butanol. Sensors 2022, 22, 6571. https://doi.org/10.3390/s22176571

AMA Style

Jiang L, Cui Q, Zhang R, Zhang W. Highly Sensing and Selective Performance Based on Bi-Doped Porous ZnSnO3 Nanospheres for Detection of n-Butanol. Sensors. 2022; 22(17):6571. https://doi.org/10.3390/s22176571

Chicago/Turabian Style

Jiang, Lili, Qi Cui, Ruijia Zhang, and Wenqiang Zhang. 2022. "Highly Sensing and Selective Performance Based on Bi-Doped Porous ZnSnO3 Nanospheres for Detection of n-Butanol" Sensors 22, no. 17: 6571. https://doi.org/10.3390/s22176571

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