Next Article in Journal
A Facile Approach to Solid-State White Emissive Carbon Dots and Their Application in UV-Excitable and Single-Component-Based White LEDs
Previous Article in Journal
First-Principles Analysis of Vibrational Properties of Type II SiGe Alloy Clathrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of a Flower-Like g-C3N4/ZnO Hierarchical Structure with Improved CH4 Sensing Properties

1
School of Materials Science and Engineering, Cultivating Base for Key Laboratory of Environment-Friendly Inorganic Materials in University of Henan Province, Henan Polytechnic University, Jiaozuo 454000, China
2
State Key Laboratory Cultivation Bases Gas Geology and Gas Control, Henan Polytechnic University, Jiaozuo 454000, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(5), 724; https://doi.org/10.3390/nano9050724
Submission received: 17 April 2019 / Revised: 6 May 2019 / Accepted: 7 May 2019 / Published: 10 May 2019

Abstract

:
In this paper, a hierarchical structure of graphite carbon nitride (g-C3N4) modified ZnO (g-C3N4/ZnO) was synthesized using a simple precipitation-calcination method. Through this method, g-C3N4 nanosheets with a controlled content were successfully decorated on the petals of flower-like ZnO. Various techniques were used to confirm the successful formation of the g-C3N4/ZnO hierarchical structure. The methane (CH4) sensing properties of g-C3N4/ZnO sensor were investigated. The result exhibited that after decorating ZnO with g-C3N4, the CH4 sensing performances of the fabricated sensor were remarkably improved. At the optimum operating temperature of 320 °C, the response of the sensor fabricated with CNZ-3 (the sample with an optimum content of g-C3N4) towards 1000 ppm CH4 was as high as 11.9 (Ra/Rg), which was about 2.2 times higher than that of the pure ZnO sensor (5.3). In addition, the CNZ-3 sensor also exhibited a fast response/recovery speed (15/28 s) and outstanding long-term stability. The enhancing CH4 sensing mechanism may be contributed to enlarged surface area, pore structure, and g-C3N4-ZnO n-n junction.

1. Introduction

Methane (CH4), the main constituent of natural gas, has been widely used for cooking, domestic heating, and industrial applications. However, CH4 is odorless, colorless, and highly flammable, which can cause some potential dangers in daily life and industrial production. Especially, CH4 can easily react with other gases in the air to produce dangerous explosives when its concentration in air is 5% to 15% [1,2,3]. In order to protect the safety of human life and property, developing effective method for fast and real-time detection of TEA is of great importance [4].
Over the past several decades, metal oxide semiconductor (MOS) based gas sensors have received much research interest because of their superiorities of low cost, easy fabrication, and fast response and recovery speed [5]. Now, various MOSs, such as ZnO [6], SnO2 [7], Fe2O3 [8], In2O3 [9], WO3 [10,11], TiO2 [12], Co3O4 [13], SnS2 [14], and CeO2 [15], have been developed as active materials for gas detection. However, as compared with the abundant research on the gas sensing properties of MOS for volatile organic compounds’ (VOCs) detection, the reports on the design and preparation of MOS nanomaterials for CH4 detection are relatively rare. Among the various reported MOSs, ZnO, as a typical n-type semiconductor, has been investigated as one potential gas sensing material owing to its good ability to respond to different gases [6,16]. In previous literature, ZnO was also found to be a potential material for CH4 detection [16]. However, due to the high chemical stability of CH4 molecules, most of the reported CH4 sensors suffered from the drawbacks of a low sensor response and long response-recover time.
Since the micro/nanostructures of MOS based on sensors have an important role in gas sensing behavior, controlling the synthesis of novel morphology and architecture of MOS nanomaterials has been a popular strategy to ameliorate gas sensing performances [17]. Recently, a three-dimensional (3D) hierarchical structure of MOS with high porosity and large specific surface area has attracted increasing interest and has been identified as a promising candidate for achieving prominent gas sensing performance [18]. In such a special structure, the high porosity and large specific surface area can endow the material with more effective channels for gas transmission and diffusion, as well as abundant active sites for gas adsorption. So, MOS with a 3D hierarchical structure always exhibits superior gas sensing properties [18,19,20]. Wang et al. synthesized flower-like WO3 hierarchical nanostructures through a simple two-step method of hydrothermal and calcination, which achieved a ppb-level sensitivity to NO2 [19]. Yang and co-works reported an ultrafast response/recovery time (0.9/1.5 s) towards trimethylamine gas through the preparation of snowflake-like α-Fe2O3 hierarchical architectures [20]. Besides the fabrication of the 3D hierarchical structure of MOS, the combination of two different nanomaterials to construct a heterostructure has also been testified to be an effective method to ameliorate gas sensing performance [21]. Many previous studies have demonstrated that heterostructured MOS materials can exhibit improved gas sensing properties as compared with their pure phase counterpart due to the synthetic effect between different sensitive materials [21,22,23,24]. Yang et al. reported the synthesis of branched SnO2/ZnO heterostructures and their highly enhanced ethanol gas sensing performance [22]. Hu et al. fabricated a CuO/CuFe2O4 heterostructure and found that the response value of the CuO/CuFe2O4 sensor to 10 ppm H2S gas was approximately 20 times higher than pure CuO [23]. Yan et al. researched the synthesis of an In2O3/ZnS heterostructure with spherical morphology via a two-step hydrothermal route. The prepared In2O3/ZnS nanocomposite showed an apparent enhancement of the ethanol sensing properties compared with pure ZnS and In2O3 [25].
G-C3N4, a novel two-dimensional material composed of carbon and nitrogen atoms, has emerged as a promising sensitizer that can improve the gas sensing properties of MOS due to its n-type semiconducting property, high specific surface area, good catalytic properties, and significant two-dimensional material characteristics [25,26,27,28,29]. Zhang et al. proved that the sensor of α-Fe2O3/g-C3N4 nanocomposites showed superior ethanol sensing performance than the pure α-Fe2O3 sensor, such as a faster response/recovery time and excellent selectivity [30]. Our group synthesized SnO2/g-C3N4 composites by a hydrothermal method, which showed a higher response (360) than pure SnO2 (230) to 500 ppm ethanol at 300 °C [28]. In previous literature, research investigating the improvement of the gas sensing properties of ZnO through combining with g-C3N4 has also been reported [29]. While most of these reports on g-C3N4/MOS are about the VOC sensing characteristics, in contrast, there are few reports focused on improving the CH4 properties of MOS by decorating with g-C3N4, especially to that decorating g-C3N4 on a pore sheet assembled ZnO hierarchical structure to ameliorate the CH4 sensing properties.
In this work, flower-like hierarchical structures of g-C3N4/ZnO nanocomposites with different mass ratios of g-C3N4 were successfully synthesized through a simple precipitation-calcination method. Through this method, different amounts of g-C3N4 were easily decorated on the flower-like ZnO hierarchical structure that assembled with pore petals. Various techniques, including X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectrometer (FTIR), field emission scanning electron microscope (FESEM), energy dispersive X-ray spectroscopic (EDS), and field emission transmission electron microscope (FETEM), were used to confirm the successful preparation of the g-C3N4/ZnO hierarchical structure. The (CH4) sensing performance of the prepared g-C3N4/ZnO was investigated and the results indicated that after decorating with g-C3N4, the ZnO sensor sensing properties towards CH4 were remarkably improved, especially regarding its higher response and faster response/recovery speed.

2. Materials and Methods

2.1. Synthesis of g-C3N4

Graphite carbon nitride (g-C3N4) was prepared via a previously reported method [30,31]. Simply, a desired amount of urea was put in a horizontal tubular furnace and continuously heated in air at 250, 350, and 550 °C and maintained for 2 h, respectively. After cooling to room temperature, the last yellow powder was g-C3N4.

2.2. Synthesis of g-C3N4/ZnO Nanocomposite

As shown in Figure 1, in order to synthesize the g-C3N4/ZnO nanocomposite, a desired amount of as-prepared g-C3N4 was dispersed in 20 mL of deionized water under ultrasonic irradiation for 2 h. Then, 0.571 g of Zn(NO3)2·6H2O was dissolved into the above suspension under vigorous stirring and 20 mL aqueous solution of NaOH (0.026 M) was added. The obtained solution was then maintained at 60 °C for 1.5 h in a water bath kettle to complete the precipitation reaction. The precipitate was obtained by centrifugation, washed with water and ethanol three times, and dried at 60 °C for 12 h. Finally, the white powder was annealed in air at 400 °C for 2 h to obtain the g-C3N4/ZnO nanocomposite. By simply adjusting the amount of g-C3N4 used, g-C3N4/ZnO nanocomposites with a g-C3N4 content of 1, 3, and 5 wt % were synthesized, which were labeled as CNZ-1, CNZ-3, and CNZ-5, respectively. For comparison, pure ZnO was also prepared under the same condition.

2.3. Characterization

The crystallographic structure of the samples was identified by powder X-ray diffraction (XRD, Bruker-AXS D8, Bruker, Madison, MI, USA). FESEM (QuantaTM250 FEG, FEI, Eindhoven, The Netherlands), EDS (FEI, Eindhoven, The Netherlands), and FSTEM (JEOL, Tokyo, Japan) were used to investigate the morphology of the prepared samples. Nitrogen adsorption-desorption isotherms were obtained on a Quantachrnme Autosorb-iQ sorption analyzer (Quantachrome, Boynton Beach, FL, USA). XPS (Ferkin Elmer Limited, Waltham Mass, Waltham, MA, USA) was measured to explode the analysis the elemental valence and surface. FTIR was measured on a Bruker Tensor 27 (Beijing, China).

2.4. Sensor Fabrication and Measurement

The CGS-4TPS (Beijing Elite Tech. Co., Ltd., Beijing, China), an intelligent gas sensing analysis system, was used to test the gas sensing properties, and the complete device is exhibited in Figure 2. The structure of the gas sensor is clearly shown in Figure 2, consisting of four parts, a sample heater (direct contact heating device), ceramic substrate (Al2O3), Ag-Pd comb-like electrodes (resistance signal of the transmission device), and sensing materials. In order to reduce the error of the test, the gas sensor was used strictly in accordance with the quantitative requirements during the manufacturing process. Firstly, 10 mg of prepared powers of CNZ was mixed with two drops of deionized water to form a homogenous paste and the paste was brush-coated on the surface of a ceramic substrate. Then, the substrates were dried at 60 °C for 24 h to obtain the sensor. Finally, the fabricated sensors were aged at 200 °C for 4 h to improve their stability. During testing, a static volumetric method was adopted to obtain different concentrations of CH4 in the closed test chamber. A certain amount of CH4 was injected into the test chamber (1.8 L) by a syringe and the relative humidity was around 35% in this measurement.

3. Results and Discussion

3.1. Morphology and Structure of the ZnO and g-C3N4/ZnO Nanocomposites

The compositions of the sample were first analyzed by XRD. Figure 3a shows the pure ZnO sample XRD pattern. All diffraction peaks matched well with the standard date of hexagonal ZnO (JCPDS: no. 89-1397), indicating the purity of the ZnO sample. Diffraction peaks of the prepared g-C3N4 are displayed in Figure 3f. In this Figure, two diffraction peaks located at 12.9° and 27.5° are clearly observed, which can be assigned to the (100) and (002) planes of g-C3N4, respectively. The XRD patterns of the prepared g-C3N4/ZnO composites are shown in Figure 3b–d, in which the characteristic peaks arising from ZnO can be clearly observed, but no obvious diffraction peaks of g-C3N4 are detected. The absence of the peaks from g-C3N4 can be attributed to the relatively low content of g-C3N4 in the composite and the results are similar with the studies of Cao [28]. In addition, the g-C3N4/ZnO composites with a g-C3N4 content of 20 wt % were prepared and the XRD are shown in Figure 3e. To be sure, the g-C3N4/ZnO composites were successfully synthesized in this work.
From Figure 4a, one can observe that g-C3N4 nanosheets were successfully prepared, and in Figure 4b, it can be seen that many 3D hierarchical microstructures with a flower-like morphology formed in the pure ZnO sample. Similar flower-like microstructures were also observed in the g-C3N4/ZnO nanocomposites, as shown in Figure 4c,d, indicating that the introduction of g-C3N4 nanosheets has almost no influence on the morphology of the products. Meanwhile, the TEM of ZnO petals are exhibited in Figure 4e. A large number of randomly dispersed nanopores with a size of about 1 to 5 nm were embedded on the petals, resulting in the formation pore petals in the flower-like architecture [32]. The existence of these nanopores on the ZnO petals can not only enlarge the specific surface area of the composite sensing materials, but also promote gas diffusion and transmission in the materials, both of which are considered to be helpful for improving the gas sensing properties. Detailed structural information of the prepared flower-like CNZ-3 was revealed by TEM observation. Figure 4f gives a typical TEM image recorded from CNZ-3, in which the g-C3N4/ZnO composites are clearly observed, consistent with the results given by FESEM observation. Figure 4g presents an enlarged TEM image of the observed flower-like composites and the lattice fringe spacing was measured to be 0.282 nm, in good agreement with the (100) plane of ZnO (JCPDS: no. 89-1397). Around the crystalline ZnO sheets, some g-C3N4 phase was also observed. From the high magnification FESEM images of CNZ-3 shown in Figure 4f,g, we can clearly see that the g-C3N4 nanosheets closely attached to the ZnO petals, suggesting that (n)ZnO-(n)g-C3N4 junctions can be created in the composite material.
In order to confirm the successful introduction of g-C3N4 into the flower-like ZnO, EDS measurement was performed on the prepared composite sample. Figure 4(h1,h2,h3,h4) show the EDS mappings recorded from an isolated flower observed in Figure 4c. Besides the Zn and O, the signals from C and N were also observed, suggesting that the observed flower was composed of g-C3N4 and ZnO.
The FTIR spectra of the prepared pure ZnO and CNZ-3 were recorded. As shown in Figure 5, compared with the pure ZnO sample, two additional peaks located at 1448 cm−1 and 880 cm−1 were observed in the CNZ-3 sample, which can be attributed to the typical stretching vibration of CN heterocycles [33] and the characteristic breathing mode of the triazine units of g-C3N4, respectively.
The specific surface area and pore size distribution of the ZnO and CNZ-3 samples were investigated by N2-sorption measurements and the results are shown in Figure 6. In Figure 6, typical type-IV isotherm curves are shown. The BET surface areas were calculated to be 37.9 m2/g and 20.1 m2/g for CNZ-3 and ZnO, respectively. The larger specific surface area of CNZ-3 could be attributed to the introduction of g-C3N4.
The chemical bond state of the prepared samples was investigated with XPS. Figure 7a shows the full survey scan spectrum recorded from CNZ-3, in which four elements, including Zn, O, C, and N, were identified, further proving the successful introduction of g-C3N4 into the flower-like ZnO. A weak N 1s peak at about 498.4 eV was observed in the high resolution spectra of N1s (Figure 7b), similar with the results reported by Kouki et al. [34]. The low intensity of the N 1s peak could be attributed to the relatively low content of g-C3N4. Figure 7c and d show the O 1s spectra of the pure ZnO and CNZ, respectively. Three kinds of oxygen species were obtained by fitting the O 1s peaks, including lattice oxygen (OL), oxygen vacancy (OV), and chemisorbed oxygen (OC), and the relative percentages of the OV and OC components were 32.9% and 34.9% for CNZ-3, and 25.3 and 24.8% for ZnO, respectively.

3.2. Gas Sensor Fabrication and Analysis

The decoration of g-C3N4 nanosheets on the porous petals-assembled flower-like ZnO was expected to bring improved gas sensing properties, thus the CH4 sensing performance of the g-C3N4/ZnO nanocomposites was investigated to explore their possible application in CH4 detection. During gas sensing tests, the pure ZnO sample was used as a reference. Considering that the operating temperature influences MOS sensors, the temperature-dependent response was first tested to obtain the optimum operating temperature. Figure 8a shows the response values of the sensors to 1000 ppm CH4 at different operating temperatures (ranging from 260 to 360 °C). All sensors showed the highest response value at 320 °C. Such a result means that the gas adsorption and desorption on the surface of sensitive materials is balanced at this temperature. While, in the relatively low temperature range (260–320 °C), the gas adsorption speed will increase with the increase of the operating temperature, resulting in a continuously increased response. Conversely, when the temperature is over 320 °C, the speed of gas desorption will exceed that of gas adsorption, which will lead to a decreased response with further increases of the temperature from 320 to 360 °C. So, 320 °C is determined as the optimum operating temperature. Meanwhile, the optimal g-C3N4 content of 3 wt% in the present composite system is exhibited.
The dynamic response–recover curves as well as the concentration-dependent response of the sensors based on ZnO and CNZ-3 are presented in Figure 9. As shown in Figure 9a,c, once exposed to 500 ppm CH4, both sensors gave a fast decrease in resistance, exhibiting characteristic n-type semiconducting properties. With a further increase of the CH4 concentration, the resistances of the two sensors decreased correspondingly. While, as compared with the pure ZnO sensor, the CNZ-3 sensor gave clear and regular response steps to different concentrations of CH4 (insets in Figure 9a,c), revealing its better ability to respond to CH4. The response variation depending on the CH4 concentration of the ZnO and CNZ-3 sensors is shown in Figure 9b,d, respectively. With successive increases of the CH4 concentration, both sensors achieved an obvious enhancement of the response and a good response linearity within 100 to 500 ppm. While, the response values of the CNZ-3 sensor to various concentrations of CH4 were always higher than the ZnO sensor, demonstrating the sensitization effect of g-C3N4 on ZnO during the sensing of CH4. Based on the fitting lines shown in the insets of Figure 9b,d, the detection limit (DL) of the sensors was estimated, which reference the literature [35]. The sensor noise was extracted using 20 points and the slope was obtained, as shown in Figure 9b [35]. Using the above equation, the DL of the pure ZnO and CNZ-3 sensors were calculated to be 10.2 ppm and 0.4 ppm, respectively. The lower DL value of CNZ-3 implies its better ability to detect low concentrations of CH4.
Figure 10a shows the sensors’ transient response-recovery curves to 100 ppm CH4. Based on this figure, the measured response/recovery (Tres/Trec) times of the ZnO and CNZ-3 sensors are 14/18 s and 26/33 s. The dynamic response of the sensors to four different concentrations of CH4 at 320 °C are shown in Figure 10b. The response amplitudes of the sensors increased correspondingly with the CH4 concentration increasing (from 100 ppm to 400 ppm). Importantly, the CNZ-3 sensor can give similar response amplitudes to the same concentration of CH4, demonstrating its good repeatability.
In addition, the long-term stability of the sensors based on CNZ-3 was also tested, as shown in Figure 11. It can be seen that the CNZ-3 sensor showed a small fluctuation in its response in the first month perhaps due to the sensor aging and a disappearance of unstable adsorption sites. While, after 90 days, the sensor could still maintain 97.7% of the initial response value (stability = 97.7%), which indicates that the sensor has excellent long-term stability. In order to further evaluate the quality of the CNZ-3 sensor, its CH4 sensing properties weree compared with different kinds of sensitive materials, as shown in Table 1. Clearly, the present g-C3N4/ZnO sensor showed a remarkable improvement in sensor response and response/recovery speed.

3.3. Gas Sensing Mechanism

The gas sensing mechanism of ZnO can be explained by the varied thickness of the electronic depletion layer (EDL) formed on the surface of ZnO when the senor is switched from air to reducing gas atmospheres [22,28,31]. In our experiment, when the sensor was exposed to air, oxygen molecules adsorbed on the surface of ZnO and then captured electrons from the conduction band of ZnO to form various types of chemisorbed oxygen anions, such as O, O2−, and O2, resulting in the formation of a thick EDL, as shown in Figure 12a. In this case, a relatively higher sensor resistance (Ra) can be obtained because of the low conductivity of the EDL. When CH4 was injected into the test chamber, the redox reaction between CH4 molecules and chemisorbed oxygen anions will occur on the ZnO surface. Meanwhile, the electrons will be released back to ZnO, which are captured by oxygen anions. So, a lower sensor resistance (Rg) will be exhibited due to the decrease of the EDL thickness, as shown in Figure 12b.
In our experiments, it was found that after decorating ZnO with g-C3N4, the response of the sensor to CH4 was remarkably improved. The improved CH4 sensing behaviors regarding CNZ-3 can be mainly attributed to three factors. Firstly, the results of the N2 sorption analysis indicated that after decorating ZnO with g-C3N4, the specific surface area of the composite material was remarkably increased. In fact, the specific surface area of CNZ-3 was found to be 37.9 m2/g, which was much larger than that of pure ZnO (20.1 m2/g). The larger specific surface area of CNZ-3 means that during the gas sensing process, more active sites can be provided for gas adsorption, thus leading to a higher response. In addition, the result of the O 1s XPS spectra (Figure 7) shows that the relative percentage of the Oc component in CNZ-3 (32.9%) is more than that of pure ZnO (25.3%). The higher Oc value in CNZ-3 indicates that the decoration of g-C3N4 on ZnO can produce more Oc anions on the surface of the composite material, resulting in, during the gas sensing process, more chemisorbed oxygen anions provided to react with CH4 molecules.
Another important reason for the improved CH4 response of the composite nanomaterial sensors should be the n-n heterojunction formed between g-C3N4 and ZnO. As observed by HRSEM and HRTEM (Figure 4), (n)ZnO-(n)g-C3N4 junctions were formed. The n-n junctions may bring two positive effects. First, the different lattice parameters between ZnO and g-C3N4, leading to a great number of defects, formed between ZnO and g-C3N4, which became potential active sites, thus causing a higher response of the composite sensor as compared with the pure ZnO sensor. Form the result given by XPS analysis, it can be found that the OV value in CNZ-3 (34.9%) was obviously higher than that in pure ZnO (24.8%), revealing that due to the formation of the g-C3N4-ZnO n-n heterojunction, more defects (oxygen vacancy) were created in the CNZ-3 sample. Secondly, because the Fermi levels of g-C3N4 are higher than ZnO [45,46], at the interface of g-C3N4 and ZnO, electrons will flow from g-C3N4 to ZnO until their Fermi levels reache equalization. Thus, once the CNZ-3 sensor is exposed to the CH4 vapor, the electrons that were trapped by chemisorbed oxygen anions will be released back to ZnO, and additional electrons coming from g-C3N4 can also be poured into ZnO through the n-n heterojunction, leading to a much thinner EDL (Figure 12d). Thus, a higher CH4 response can be obtained on the CNZ-3 sensor.
In addition, the good catalytic property of g-C3N4 was also considered to be helpful for the improved CH4 response of the g-C3N4/ZnO nanocomposites. In the air atmosphere, g-C3N4 can promote oxygen molecules to form functional groups, such as hydroxyl (*OH), which can facilitate the oxidation of CH4 molecules [47,48].

4. Conclusions

In summary, a g-C3N4 decorated ZnO hierarchical structure was synthesized via a facile precipitation-calcination method, through which g-C3N4 nanosheets with a controlled content were successfully decorated on the ZnO hierarchical structure assembled with pore petals. The results displayed that after decorating with g-C3N4, the ZnO sensor CH4 sensing properties were remarkably enhanced. The response of the sensor fabricated with CNZ-3 towards 1000 ppm CH4 was as high as 11.9, which was about 2.2 times higher than that of the pure ZnO sensor (5.3) at the optimum operating temperature of 320 °C. Furthermore, the CNZ-3 sensor also exhibited a fast response/recovery (15/28 s) speed and good long-term stability, demonstrating its potential application in CH4 detection. The present research demonstrates that decorating ZnO with two-dimensional g-C3N4 to construct n-n junctions is a practical method to enhance the gas sensing performance of ZnO. In future work, we will consider other porous nanomaterials [49] with g-C3N4 to improve the CH4 gas sensor performance.

Author Contributions

X.L. designed the materials, finished the experimental and prepared original draft. Y.L. guided and solved the problems in the experimental. G.S. investigated the gas sensing mechanism and revised the manuscript. Z.Z. reviewed the manuscript. N.L. analyzed the data. B.Z. character the materials.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (U1404613, U1704255), Program for Science & Technology Innovation Talents in Universities of Henan Province (18HASTIT010, 17HASTIT029), Our group thankful to the Young Core Instructor Project of Colleges and Universities in Henan Province (2015GGJS-063, 2016GGJS-040) and Foundation of Henan Scientific and Technology key project (182102310892).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vuong, N.M.; Hieu, N.M.; Hieu, H.N.; Yi, H.; Kim, D.; Han, Y.S. Ni2O3-decorated SnO2 particulate films for methane gas sensors. Sens. Actuators B 2014, 192, 327–333. [Google Scholar] [CrossRef]
  2. Shemshad, J.; Aminossadati, S.M.; Kizil, M.S. A review of developments in near infrared methane detection based on tunable diode laser. Sens. Actuators B 2012, 171, 77–92. [Google Scholar] [CrossRef]
  3. Lu, W.B.; Ding, D.G.; Xue, Q.Z.; Du, Y.G.; Xiong, Y.; Zhang, J.Q.; Pan, X.L.; Xing, W. Great enhancement of CH4 sensitivity of SnO2 based nanofibers by heterogeneous sensitization and catalytic effect. Sens. Actuators B 2018, 254, 393–401. [Google Scholar] [CrossRef]
  4. Ye, W.L.; Zheng, C.T.; Yu, X.; Zhao, C.X.; Song, Z.W.; Wang, Y.D. Design and performances of a mid-infrared CH4 detection device with novel three-channel-based LS-FTF self-adaptive denoising structure. Sens. Actuators B 2011, 155, 37–45. [Google Scholar] [CrossRef]
  5. Wang, Y.; Liu, C.Y.; Wang, Z.; Song, Z.W.; Zhou, X.Y.; Han, N.; Chen, Y.F. Sputtered SnO2 NiO thin films on self-assembled Au nanoparticle arrays for MEMS compatible NO2 gas sensors. Sens. Actuators B 2019, 278, 28–38. [Google Scholar] [CrossRef]
  6. Bian, H.Q.; Ma, S.Y.; Sun, A.M.; Xu, X.L.; Yang, G.J.; Yan, S.H.; Gao, J.M.; Zhang, Z.M.; Zhu, H.B. Improvement of acetone gas sensing performance of ZnO nanoparticles. J. Alloys Compd. 2016, 658, 629–635. [Google Scholar] [CrossRef]
  7. Jia, C.C.; Dong, T.; Li, M.; Wang, P.; Yang, P. Preparation of anatase/rutile TiO2/SnO2 hollow heterostructures for gas sensor. J. Alloys Compd. 2018, 769, 521–532. [Google Scholar] [CrossRef]
  8. Sun, X.H.; Ji, H.M.; Li, X.L.; Cai, S.; Zheng, C.M. Open-system nanocasting synthesis of nanoscale α-Fe2O3 pores structure with enhanced acetone-sensing properties. J. Alloys Compd. 2014, 600, 111–117. [Google Scholar] [CrossRef]
  9. Qurashi, A.; El-Maghraby, E.M.; Yamazaki, T.; Shen, Y.B.; Kikuta, T. A generic approach for controlled synthesisof In2O3 nanostructures for gas sensing applications. J. Alloys Compd. 2009, 481, 35–39. [Google Scholar] [CrossRef]
  10. Chen, L.; Huang, L.; Lin, Y.J.; Sai, L.; Chang, Q.H.; Shi, W.Z.; Chen, Q. Fully gravure-printed WO3/Pt-decorated rGO nanosheets composite film for detection of acetone. Sens. Actuators B 2018, 255, 1482–1490. [Google Scholar] [CrossRef]
  11. Ou, J.Z.; Yaaco, M.H.; Campbell, J.L.; Breedon, M.; Kalantar-zadeh, K.; Wlodarski, W. H2 sensing performance of optical fiber coated with nano-platelet WO3 film. Sens. Actuators B 2012, 1, 166–167. [Google Scholar] [CrossRef]
  12. Ou, J.Z.; Ge, W.Y.; Carey, B.J.; Daeneke, T.; Rotbart, A.; Shan, W.; Wang, Y.C.; Fu, Z.Q.; Chrimes, A.F.; Wlodarski, W.S.; et al. Physisorption-based charge transfers in two-dimensional SnS2 for selective reversible NO2 gas sensing. ACS Nano 2015, 9, 10313–10323. [Google Scholar] [CrossRef] [PubMed]
  13. Alberti, A.; Renna, L.; Sanzaro, S.; Smecca, E.; Mannino, G.; Bongiorno, C.; Galati, C.; Gervasi, L.; Sanangelo, A.; Santangelo, A.; et al. Innovative spongy TiO2 layers for gas detection at low working temperature. Sens. Actuators B 2018, 259, 658–667. [Google Scholar] [CrossRef]
  14. Zhao, G.Y.; Tang, L.; Zhang, L.; Chen, X.; Mao, Y.C.; Sun, K.N. Well-developed capacitive-capacity of metal-organic framework derived Co3O4 films in Li ion battery anodes. J. Alloys Compd. 2018, 746, 277–284. [Google Scholar] [CrossRef]
  15. Xie, W.Y.; Duan, X.C.; Deng, J.W.; Nie, J.X.; Wang, T.H. CeO2/ionic liquid hybrid materials with enhanced humidity performance. Sens. Actuators B 2017, 252, 870–876. [Google Scholar] [CrossRef]
  16. Zhang, D.; Yin, N.; Xia, B. Facile fabrication of ZnO nanocrystalline-modified graphene hybrid nanocomposite toward methane gas sensing application. J. Mater. Sci. Mater. Electron. 2015, 26, 5937–5945. [Google Scholar] [CrossRef]
  17. Zou, Y.; Wang, H.; Yang, R.; Lai, X.; Wan, J.; Lin, G.; Liu, D. Controlled synthesis and enhanced toluene-sensing properties of mesopores NixCo1-xFe2O4 nanostructured microspheres with tunable composite. Sens. Actuators B 2019, 280, 227–234. [Google Scholar] [CrossRef]
  18. Xu, H.; Li, W.; Han, R.; Zhai, T.; Yu, H.; Yu, Z.; Chen, Z.; Wu, X.; Wang, J.; Cao, B. Enhanced triethylamine sensing properties by fabricating Au@SnO2/α-Fe2O3 core-shell nanoneedles directly on alumina tubes. Sens. Actuators B 2018, 262, 70–78. [Google Scholar] [CrossRef]
  19. Wang, C.; Sun, R.Z.; Li, X.; Sun, Y.F.; Sun, P.; Liu, F.M.; Lu, G.Y. Hierarchical flower-like WO3 nanostructures and their gas sensing properties. Sens. Actuators B 2017, 204, 224–230. [Google Scholar] [CrossRef]
  20. Yang, T.Y.; Du, L.Y.; Zhai, C.B.; Li, Z.F.; Zhao, Q.; Luo, Y.; Xing, D.J.; Zhang, M.Z. Ultrafast response and recovery trimethylamine sensor based on α-Fe2O3 snowflake-like hierarchical architectures. J. Alloys Compd. 2017, 718, 396–404. [Google Scholar] [CrossRef]
  21. Motaung, D.E.; Mhlongo, G.H.; Makgwane, P.R.; Dhonge, B.P.; Cummings, F.R.; Swart, H.C.; Ray, S.S. Ultra-high sensitive and selective H2 gas sensor manifested by interface of n–n heterostructure of CeO2-SnO2 nanoparticles. Sens. Actuators B 2018, 254, 984–995. [Google Scholar] [CrossRef]
  22. Yang, X.; Zhang, S.; Zhao, L.; Sun, P.; Wang, T.; Liu, F.; Yan, X.; Gao, Y.; Liang, X.; et al. One step synthesis of branched SnO2/ZnO heterostructures and their enhanced gas-sensing properties. Sens. Actuators B 2019, 281, 415–423. [Google Scholar] [CrossRef]
  23. Hu, X.B.; Zhu, Z.G.; Li, Z.H.; Xie, L.L.; Wu, Y.H.; Zheng, L.Y. Heterostructure of CuO microspheres modified with CuFe2O4 nanoparticles for highly sensitive H2S gas sensor. Sens. Actuators B 2018, 264, 139–149. [Google Scholar] [CrossRef]
  24. Chen, Q.; Ma, S.Y.; Xu, X.L.; Jiao, H.Y.; Zhang, G.H.; Liu, L.W.; Wang, P.Y.; Gengzang, D.J.; Yao, H.H. Optimization ethanol detection performance manifested by gas sensor based on In2O3/ZnS rough microspheres. Sens. Actuators B 2018, 264, 263–278. [Google Scholar] [CrossRef]
  25. Wang, X.; Blechert, S.; Antonietti, M. Polymeric graphitic carbon nitride for heterogeneous photocatalysis. ACS Catal. 2012, 254, 1596–1606. [Google Scholar] [CrossRef]
  26. Das, D.; Shinde, S.L.; Nanda, K.K. Temperature-dependent photoluminescence of g-C3N4 implication for temperature sensing. ACS Appl. Mater. Interfaces 2016, 8, 2181–2186. [Google Scholar] [CrossRef]
  27. Zhang, Y.J.; Zhang, D.K.; Guo, W.M.; Chen, S.J. The a-Fe2O3/g-C3N4 heterostructural nanocomposites with enhanced ethanol gas sensing performance. J. Alloys Compd. 2016, 685, 84–90. [Google Scholar] [CrossRef]
  28. Cao, J.; Qin, C.; Wang, Y.; Zhang, B.; Gong, Y.; Zhang, H.; Sun, G.; Bala, H.; Zhang, Z. Calcination method synthesis of SnO2/g-C3N4 composites for a high-performance ethanol gas sensing application. Nanomaterials 2017, 7, 98. [Google Scholar] [CrossRef] [PubMed]
  29. Zhai, J.L.; Wang, T.; Wang, C.; Liu, D.C. UV-light-assisted ethanol sensing characteristics of g-C3N4/ZnO composites at room temperature. Appl. Surf. Sci. 2017, 206, 406–416. [Google Scholar]
  30. Dong, F.; Wang, Z.Y.; Sun, Y.J.; Ho, W.-K.; Zhang, H.D. Engineering the nanoarchitecture and texture of polymeric carbon nitride semiconductor for enhanced visible light photocatalytic activity. J. Colloid Interface Sci. 2013, 401, 70–79. [Google Scholar] [CrossRef]
  31. Yang, X.L.; Yu, Q.; Zhang, S.F.; Sun, P.; Lu, H.Y.; Yan, X.; Liu, F.M.; Zhou, X.; Liang, X.H.; Gao, Y.; et al. Highly sensitive and selective triethylamine gas sensor based on pores SnO2/Zn2SnO4 composites. Sens. Actuators B Chem. 2018, 266, 213–220. [Google Scholar] [CrossRef]
  32. Butburee, T.; Bai, Y.; Wang, H.J.; Chen, H.J.; Wang, Z.L.; Liu, G.; Zou, J.; Khemthong, P.; Lu, G.Q.M.; Wang, L.Z. 2D pores TiO2 single-crystalline nanostructure demonstrating high photo-electrochemical water splitting performance. Adv. Mater. 2018, 30, 1705666. [Google Scholar] [CrossRef] [PubMed]
  33. Mao, Z.; Chen, J.; Yang, Y.; Wang, D.; Bie, L.; Fahlman, B.D. Novel g-C3N4/CoO nanocomposites with significantly enhanced visible-light photocatalytic activity for H2 evolution. ACS Appl. Mater. Interfaces 2017, 9, 12427–12435. [Google Scholar] [CrossRef] [PubMed]
  34. Akaike, K.; Aoyama, K.; Dekubo, S.; Onishi, A.; Kanai, K. Characterizing electronic structure near the energy gap of graphitic carbon nitride based on rational interpretation of chemical analysis. Chem. Mater. 2018, 30, 2341–2352. [Google Scholar] [CrossRef]
  35. Li, Y.W.; Luo, N.; Sun, G.; Zhang, B.; Jin, H.H.; Lin, L.; Bala, H.; Cao, J.L.; Zhang, Z.Y.; Wang, Y. Synthesis of porous nanosheets-assembled ZnO/ZnCo2O4 hierarchical structure for TEA detection. Sens. Actuators B Chem. 2019, 287, 199–208. [Google Scholar] [CrossRef]
  36. Sung, J.H.; Lee, Y.S.; Lim, J.W.; Hong, Y.H.; Lee, D.D. Sensing characteristics of tin dioxide/gold sensor prepared by coprecipitation method. Sens. Actuators B Chem. 2000, 1, 149–152. [Google Scholar] [CrossRef]
  37. Bose, S.; Chakraborty, S.; Ghosh, B.K.; Das, D.; Sen, A.; Maiti, H.S. Methane sensitivity of Fe-doped SnO2 thick films. Sens. Actuators B Chem. 2005, 2, 346–350. [Google Scholar] [CrossRef]
  38. Min, B.; Choi, S. Undoped and 0.1wt % Ca-doped Pt-catalyzed SnO2 sensors for CH4 detection. Sens. Actuators B Chem. 2005, 1, 119–124. [Google Scholar] [CrossRef]
  39. Waitz, T.; Becker, B.; Wagner, T.; Sauerwald, T.; Kohl, C.D.; Tiemann, M. Ordered nanopores SnO2 gas sensors with high thermal stability. Sens. Actuators B Chem. 2010, 2, 788–793. [Google Scholar] [CrossRef]
  40. Wagner, T.; Bauer, M.; Sauerwald, T.; Kohl, C.D.; Tiemann, M. X-Ray absorption near-edge spectroscopy investigation of the oxidation state of Pd species in nanopores SnO2 gas sensors for methane detection. Thin Solid Films 2011, 3, 909–912. [Google Scholar] [CrossRef]
  41. Barazzouk, S.; Tandon, R.P.; Hotchandani, S. MoO3-based sensor for NO, NO2 and CH4 detection. Sens. Actuators B Chem. 2006, 2, 691–694. [Google Scholar] [CrossRef]
  42. Suzuki, T.; Kunihara, K.; Kobayashi, M.; Tabata, S.; Higaki, K.; Ohnishi, H. A micromachined gas sensor based on a catalytic thick film/SnO2 thin film bilayer and thin film heater. Sens. Actuators B Chem. 2005, 2, 185–189. [Google Scholar] [CrossRef]
  43. Nasresfahani, S.; Sheikhi, M.H.; Tohidi, M.; Zarifkar, A. Methane gas sensing properties of Pd-doped SnO2/reduced graphene oxide synthesized by a facile hydrothermal route. Mater. Res. Bull. 2017, 89, 161–169. [Google Scholar] [CrossRef]
  44. Xue, D.P.; Wang, Y.; Cao, J.L.; Sun, G.; Zhang, Z.Y. Improving methane gas sensing performance of flower-like SnO2 decorated by WO3 nanoplates. Talanta 2019, 199, 603–611. [Google Scholar] [CrossRef]
  45. Dong, F.; Zhao, Z.; Xiong, T.; Ni, Z.; Zhang, W.; Sun, Y.; Ho, W.K. In situ construction of g-C3N4/g-C3N4 Metal-Free heterojunction for enhanced visible-light photocatalysis. Appl. Mater. Interface 2013, 13, 11392–11401. [Google Scholar] [CrossRef] [PubMed]
  46. Jo, W.-K.; Selvam, N.C.S. Enhanced visible light-driven photocatalytic performance of ZnO/g-C3N4 coupled with graphene oxide as a novel ternary. J. Hazard. Mater. 2015, 299, 462–470. [Google Scholar] [CrossRef]
  47. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Origin of the electrocatalytic oxygen reduction activity of graphene-based catalysts a roadmap to achieve the best performance. J. Am. Chem. Soc. 2014, 136, 4394–4403. [Google Scholar] [CrossRef]
  48. Xu, J.; Markus, A. The performance of nanoparticulate graphitic carbon nitride as an amphiphile. J. Am. Chem. Soc. 2017, 139, 6026–6029. [Google Scholar] [CrossRef]
  49. Haque, F.; Zavabeti, A.; Zhang, B.Y.; Datta, R.S.; Yin, Y.F.; Yi, Z.F.; Wang, Y.C.; Mahmood, N.; Pillai, N.; Syed, N. Ordered intracrystalline pores in planar molybdenum oxide for enhanced alkaline hydrogen evolution. J. Mater. Chem. A 2019, 7, 257. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration for the synthesis of pure ZnO and g-C3N4/ZnO.
Figure 1. Schematic illustration for the synthesis of pure ZnO and g-C3N4/ZnO.
Nanomaterials 09 00724 g001
Figure 2. Schematic presentation of the gas sensing analysis system and 3D schematic diagram of the sensor element.
Figure 2. Schematic presentation of the gas sensing analysis system and 3D schematic diagram of the sensor element.
Nanomaterials 09 00724 g002
Figure 3. XRD patterns of the prepared samples: (a) pure ZnO, (b) CNZ-1, (c) CNZ-3, (d) CNZ-5, and (e) g-C3N4/ZnO composites with a g-C3N4 content of 20 wt %, (f) g-C3N4.
Figure 3. XRD patterns of the prepared samples: (a) pure ZnO, (b) CNZ-1, (c) CNZ-3, (d) CNZ-5, and (e) g-C3N4/ZnO composites with a g-C3N4 content of 20 wt %, (f) g-C3N4.
Nanomaterials 09 00724 g003
Figure 4. FESEM images of g-C3N4 (a), pure ZnO (b), and CNZ-3 (c,d), (e,f) TEM and (g) HRTEM image of CNZ-3, (h1,h2,h3,h4) elemental mapping of Zn, O, N, and C (CNZ-3).
Figure 4. FESEM images of g-C3N4 (a), pure ZnO (b), and CNZ-3 (c,d), (e,f) TEM and (g) HRTEM image of CNZ-3, (h1,h2,h3,h4) elemental mapping of Zn, O, N, and C (CNZ-3).
Nanomaterials 09 00724 g004
Figure 5. FTIR spectra of the pure ZnO and CNZ-3.
Figure 5. FTIR spectra of the pure ZnO and CNZ-3.
Nanomaterials 09 00724 g005
Figure 6. N2 adsorption-desorption isotherms.
Figure 6. N2 adsorption-desorption isotherms.
Nanomaterials 09 00724 g006
Figure 7. XPS spectra of the prepared sample: (a) full survey scan spectrum of CNZ-3, (b) high resolution of N 1s spectrum of CNZ-3, and O1s spectra of (c) pure ZnO and (d) CNZ-3.
Figure 7. XPS spectra of the prepared sample: (a) full survey scan spectrum of CNZ-3, (b) high resolution of N 1s spectrum of CNZ-3, and O1s spectra of (c) pure ZnO and (d) CNZ-3.
Nanomaterials 09 00724 g007
Figure 8. (a) The temperature-dependent response of the sensors to 1000 ppm CH4; (b) the resistances in air (Ra) of the pure ZnO and CNZ-3 sensors at different operating temperatures.
Figure 8. (a) The temperature-dependent response of the sensors to 1000 ppm CH4; (b) the resistances in air (Ra) of the pure ZnO and CNZ-3 sensors at different operating temperatures.
Nanomaterials 09 00724 g008
Figure 9. Dynamic electrical responses of the (a) ZnO and (c) CNZ-3 sensors with the CH4 concentration increasing from 500 ppm to 5000 ppm at 320 °C; the concentration-dependent response curves of (b) ZnO and (d) CNZ-3.
Figure 9. Dynamic electrical responses of the (a) ZnO and (c) CNZ-3 sensors with the CH4 concentration increasing from 500 ppm to 5000 ppm at 320 °C; the concentration-dependent response curves of (b) ZnO and (d) CNZ-3.
Nanomaterials 09 00724 g009
Figure 10. (a) Transient response-recover curves of the sensors towards 100 ppm CH4 at 320 °C; (b) dynamic response of the sensors to different concentrations of CH4 at 320 °C.
Figure 10. (a) Transient response-recover curves of the sensors towards 100 ppm CH4 at 320 °C; (b) dynamic response of the sensors to different concentrations of CH4 at 320 °C.
Nanomaterials 09 00724 g010
Figure 11. Long-term stability of the CNZ-3 sensor towards 500 ppm CH4 at 320 °C.
Figure 11. Long-term stability of the CNZ-3 sensor towards 500 ppm CH4 at 320 °C.
Nanomaterials 09 00724 g011
Figure 12. Schematic illustration of CH4 sensing mechanism of pure ZnO (a,b) and g-C3N4/ZnO (c,d).
Figure 12. Schematic illustration of CH4 sensing mechanism of pure ZnO (a,b) and g-C3N4/ZnO (c,d).
Nanomaterials 09 00724 g012
Table 1. Comparison of the CH4 sensing performance of different reports.
Table 1. Comparison of the CH4 sensing performance of different reports.
Sensitive MaterialsTemp. (°C)S (Ra/Rg)CH4 Con. (ppm)Tres/Trec (s)
SnO2/Au [36]4000.823000--
Fe-doped SnO2 [37]3500.671000--
Ca-doped Pt-catalyzed SnO2 [38]400175000--
Mesopores SnO2 [39]6000.64000--
Pd-Nanopores SnO2 [40]600206000--
MoO3 [41]50010500--
Pd–Al2O3/SnO2 [42]450104000--
Pd-doped SnO2/rGO [43]Rt10%*14,000300/420
WO3/SnO2 [44]1102.9500--
G-C3N4/ZnO This work3202.5610015/28
* (   Ra Rg Ra × 100   % = S   % ) .

Share and Cite

MDPI and ACS Style

Li, X.; Li, Y.; Sun, G.; Luo, N.; Zhang, B.; Zhang, Z. Synthesis of a Flower-Like g-C3N4/ZnO Hierarchical Structure with Improved CH4 Sensing Properties. Nanomaterials 2019, 9, 724. https://doi.org/10.3390/nano9050724

AMA Style

Li X, Li Y, Sun G, Luo N, Zhang B, Zhang Z. Synthesis of a Flower-Like g-C3N4/ZnO Hierarchical Structure with Improved CH4 Sensing Properties. Nanomaterials. 2019; 9(5):724. https://doi.org/10.3390/nano9050724

Chicago/Turabian Style

Li, Xiaojie, Yanwei Li, Guang Sun, Na Luo, Bo Zhang, and Zhanying Zhang. 2019. "Synthesis of a Flower-Like g-C3N4/ZnO Hierarchical Structure with Improved CH4 Sensing Properties" Nanomaterials 9, no. 5: 724. https://doi.org/10.3390/nano9050724

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