Next Article in Journal
Preparation of Cellulose Nanocrystal-Reinforced Physical Hydrogels for Actuator Application
Previous Article in Journal
Service Life Modeling of Concrete with SCMs Using Effective Diffusion Coefficient and a New Binding Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of Co(OH)xF2−x Nanosheets for Acetone Gas Sensor Applications: Material Characterization and Sensor Performance Evaluation

1
International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology of Ministry of Education, Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen 518060, China
2
School of Advanced Materials Science and Engineering, Sungkyunkwan University, Suwon 16419, Korea
3
SKKU Advanced Institute of Nano Technology, Sungkyunkwan University, Suwon 16419, Korea
4
School of Chemical Engineering, Sungkyunkwan University, Suwon 16419, Korea
5
Shenzhen Institute of Advanced Electronic Materials—Shenzhen Institutes of Advanced Technology, Chinese Academy of Sciences, Shenzhen 518055, China
6
Smart Sensor Research Center, Korea Electronics Technology Institute (KETI), Seongnam 13509, Korea
*
Authors to whom correspondence should be addressed.
These 1st authors equally contributed to this work.
Current address: Department of Smart Fab. Technology, Sungkyunkwan University, Suwon 16419, Korea.
Crystals 2020, 10(11), 968; https://doi.org/10.3390/cryst10110968
Submission received: 9 October 2020 / Revised: 15 October 2020 / Accepted: 23 October 2020 / Published: 26 October 2020
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
This study reports the employment of Co(OH)xF2−x nanosheets, a new material in the sensor field, for gas sensor applications. We synthesize Co(OH)xF2−x nanosheets via a hydrothermal route using SiO2 sphere templates. Our material characterization confirms that the material is a densely clustered Co(OH)xF2−x nanosheet with an amorphous microstructure with some short-range ordering. Sensors based on the nanosheets demonstrate a high response of 269% toward 4.5 ppm of acetone gas at an operation temperature of 200 °C and a very low minimum detection limit of 40 ppb. It functions effectively up to a temperature below 300 °C, above which F is found to start to evaporate. Our discussion suggests that an excellent sensor performance arises from the high catalytic function of F incorporated in a high concentration in the material as well as the high specific surface area due to the morphology of densely clustered nanosheets.

Graphical Abstract

1. Introduction

There is a growing demand for the sensitive detection of acetone due to its possible usage as a breath marker for the diagnosis of diseases such as diabetes [1]. For breath marker applications, a very high sensitivity (sub-ppm level or lower) is necessary since the exhaled acetone concentration range is very low—from 0.3 to 1.8 ppm [2,3]. Due to the low cost and simplicity of fabrication/operation, metal oxide semiconductors, such as ZnO, SnO2, and Co3O4, have been actively explored as acetone sensor materials [4,5,6,7]. Among them, Co3O4, as a p-type material, is a promising candidate for acetone sensor materials due to its low and stable initial resistance and the high catalytic reactivity of Co2+ [8,9,10,11]. In a study conducted by Vladimirova et al., for instance, the tolerance of Co3O4 toward high humidity levels is discussed [12]. In experiments by Zhang et al., a Co3O4-based sensor with an optimized morphology of crossed nanosheets showed a response of 36.5 toward acetone gas of 1000 ppm. Zhang et al. also reported Co3O4 nanowires being hybridized with hollow carbon spheres exhibiting a response of 3.71 toward acetone gas of 200 ppm [13]. Xu et al. synthesized metal-organic framework (MOF)-derived Co3O4/Fe2O3 p-n hollow cubes to enhance acetone sensing properties (response of 3.27 to 100 ppm) [14]. However, the detection of acetone at a low concentration (sub-ppm or lower) still remains a challenge.
Doping metal oxide materials with fluorine, known to promote catalytic activity, has been considered as one of the effective routes to enhance sensing ability [15,16]. Barreca et al. employed Co(hfa)2·TMEDA (hfa = 1,1,1,5,5,5-hexafluoro-2,4-pentanedionate; TMEDA = N,N,N′,N′-tetramethylethylenediamine) as a precursor in chemical vapor deposition to produce F-doped Co3O4 nanomaterials [15]. Nonetheless, such a doping method has a limitation in obtaining a fluorine concentration high enough for a significant sensitivity improvement. Recently, cobalt fluoride hydroxide (Co(OH)xF2−x), a material including F anions as one of its main elements, came into the spotlight for being applied to catalysts for water oxidation, hydrogen production, and overall water splitting, taking full advantage of the high catalytic activity of F anions in the materials [9,17,18,19]. Nevertheless, a challenge to employ this material for gas sensor applications is its poor thermal stability: it could easily transform into Co3O4 via fluorine evaporation during a high temperature process such as the aging process (200~450 °C) which is usually required for sensor fabrication [20].
In this study, we report, for the first time, Co(OH)xF2−x nanosheets employed as a sensing material for acetone gas sensors. We synthesized Co(OH)xF2−x nanosheets via a hydrothermal route on the template of SiO2 spheres. Our synthesis method using the SiO2 spheres as template helped achieve a high fluorine density and suppress fluorine evaporation to a considerable extent upon annealing at elevated temperatures. Sensors fabricated using the Co(OH)xF2−x materials showcased a high response at a low acetone concentration (269% at 4.5 ppm) and ppb-level detection limit (~40 ppb) at an operation temperature of 200 °C. We also discussed the effects of F anions on the sensing properties of Co(OH)xF2−x nanosheets.

2. Materials and Methods

2.1. Fabrication of Co(OH)xF2−x Nanosheets

All the chemicals were used without any purification. Figure 1 shows the schematic diagram of the synthesis procedure used for the fabrication of Co(OH)xF2−x nanosheets. SiO2 spheres were synthesized using tetraethyl orthosilicate (TEOS) as described in another report [21]. Co(OH)xF2−x nanosheets were synthesized as follows. CoCl2 (0.3 mM), NH4F (0.3 mM), and urea (0.1 mM) were added into de-ionized (DI) water (30 mL). After stirring for 30 min, the SiO2 spheres (60 mg) were added into the solution, followed by sonication for 30 min. Finally, the solution was transferred to a Teflon-lined steel autoclave with a capacity of 50 mL and heated at 120 °C for 12 h under autogenous pressure. The final solution was then cooled to room temperature. Powder, precipitated from the solution, was collected by filtering and was washed with water and ethanol. The powder was then dried at 80 °C overnight. It went through a subsequent aging process at 120 °C for 10 h for removal of the SiO2 sacrificial template [22]. Chemical analysis using electron dispersive spectroscopy (EDS; EDS attached to the JEOL ARM 200F), nevertheless, disclosed that some SiO2 had remained in the sample after the removal process (see Supplementary Materials Figure S1). For comparison, some materials were synthesized without templates via the same process.

2.2. Characterization

Scanning electron microscopy (SEM; JSM-7600F, JEOL, Tokyo, Japan) and transmission electron microscopy (TEM; JEM-ARM 200F, JEOL, Tokyo, Japan) were used to investigate the morphology of the SiO2 spheres and Co(OH)xF2−x nanosheets. The infrared spectra of materials were recorded using Fourier transform infrared spectroscopy (FTIR; IFS-66/S, Bruker, Billerica, MA, USA). For characterization of the phases of the materials, X-ray diffraction (XRD; D8 Advance diffractometer, Bruker, Billerica, MA, USA) was utilized. We used X-ray photoelectron spectroscopy (XPS; K-alpha, Thermo Scientific, Waltham, MA, USA) for the chemical analysis of the materials. The contour of the sensor device with Co(OH)xF2−x powders was scanned by a 3D profiler (Contour GT-X, Bruker, Billerica, MA, USA). The Brunauer–Emmett–Teller (BET; Micromeritics, Norcross, GA, USA) surface area of the Co(OH)xF2−x nanosheets was determined from nitrogen adsorption-desorption isotherm. The SEM equipment used in this study is located at the MEMS·Sensor Platform Center of SungKyunKwan University (SKKU), Suwon, Korea.

2.3. Fabrication of Gas Sensor and Gas-Sensing Measurements

The detailed fabrication process of interdigitated (IDT) electrodes and details of gas-sensing system are reported in our previous work [23]. Co(OH)xF2−x nanosheet powders were dissolved in DI water (0.5 wt. %) and drop-coated on the IDT electrodes on a hot plate (100 °C). After full evaporation of the solvent, these devices were used for gas-sensing measurements (see Figure S2). Besides, some of the sensors were annealed at 300, 350, 400, and 450 °C for 2 or 4 h to investigate the effects of post-annealing conditions on the materials. For additional platinum (Pt) doping on the materials, a 1 nm-thick Pt film was loaded on the Co(OH)xF2−x nanosheets of the devices using a magnetron sputtering system. A home-made gas-sensing system was described in our previous study. Gas-sensing characteristics were investigated for acetone (CH3COCH3), toluene (C6H5CH3), carbon monoxide (CO), hydrogen (H2), and nitrogen dioxide (NO2), which were diluted with dry air (2 L/min) using mass flow controllers (MFCs). The relative humidity (RH) of the laboratory varied from 30 to 63% and that of the sensing chamber was kept at 10% for all sensing tests. The operating temperature of gas sensors was controlled from 25 to 350 °C by Joule heating using a ceramic heater connected to a power supplier. The response of the test is defined as [(Rg − Ra)/Ra]*100 or [(Ra − Rg)/Rg]*100 for reactions with reducing gases and oxidizing gases, respectively, where Ra and Rg represent resistances in ambient and an analyte gas.

3. Results

3.1. Synthesis and Characterization of Co(OH)xF2−x Nanosheets

The morphology of the templates of SiO2 spheres and the material hydrothermally grown on the templates were characterized using SEM, as shown in Figure 2. Figure 2a,b indicates that the SiO2 spheres were fabricated with a good dispersity, clean surface, and uniform size. The average diameter of the SiO2 spheres is around 160 nm. Figure 2c,d shows the morphology of the material grown on the templates after removal of the templates. It consists of very thin, very curly sheets, which appear clustered around what seem to have been occupied by the SiO2 spheres before the removal.
We employed XPS to investigate the chemical state of the synthesized material. Figure 3a–c present XPS analysis results on the peaks of F 1s, Co 2p, and O 1s. The binding energy of the F 1s (~684.3 eV) peak is close with that reported in a previous study on Co(OH)F [9]. Co 2p subpeaks corresponding to inorganic Co2+ species with unpaired 3d electrons and mainly appear at 781.1 ± 0.05 and 783.3 ± 0.05 (Co 2p1/2), and 796.9 ± 0.05 and 798.0 ± 0.05 eV (Co 2p3/2), which are somewhat higher than the binding energies of Co3+ species [17,18,24]. Other subpeaks at 785.9 ± 0.05, and 803.2 ± 0.05 eV are ascribed to the Co2+ shake-up satellite [18]. The O 1s peak is deconvoluted into two subpeaks—a large one at 531.6 ± 0.05 eV corresponding to metal hydroxides and a small one at 533.5 ± 0.05 eV from organic contaminants [9]. A quantitative analysis of the XPS data discloses that the fluorine content in the sample is substantial (peak intensity ratio between F and Co peaks (F/Co): ~25%), much higher than that obtained via doping (F/Co: ~10%) [15]. Figure 3d shows an XRD pattern of the synthesized material, which exhibits very small peaks at positions close to those of the orthorhombic Co(OH)F phase. It is interesting to note the asymmetric bundle of peaks at 32~45°, an unusual feature reported also in other studies on Co(OH)2 materials [25,26]. These data of XRD and XPS suggest that the synthesized material could be amorphous or nanocrystalline Co(OH)xF2−x with a broad range of fluorine compositions.
In the TEM images shown in Figure 4a,b, the sheets appear highly electron-transparent, confirming that they are very thin, probably only a few atomic layers thick. The morphology of the nanosheets suggests that they have been densely grown around SiO2 spheres, the places of which now appear empty because of the removal of the spheres via aging (for example, see the box in Figure 4a). The BET analysis confirmed that the specific surface area of the material was around 193.38 m2 g−1, a value larger than or comparable with those of 2D materials such as graphene and MoS2 (see Figure S3 for the BET data) [27,28,29].
Figure 4c shows elemental mapping acquired from the EDS analysis for the area shown in Figure 4b, exhibiting that Co, O, and F are uniformly distributed throughout the sheets. Figure 4d,e shows high-resolution TEM (HRTEM) images of the Co(OH)xF2−x nanosheets, showing that the material is mainly amorphous, without any noticeable long-range order. Nevertheless, a close look into the image discloses that there are many areas (some marked with yellow circles) showing some short-range ordering. Analysis on such short-range ordering areas indicates that the plane spacings are close to those of Co(OH)F, as shown in Figure 4e [19]. These results confirm that the synthesized nanosheets are amorphous Co(OH)xF2−x with some short-range ordering. In addition, it should be also noted from the EDS analysis that although the SiO2 spheres were supposedly removed as a whole through the removal process, the analysis indicates that there are some Si atoms uniformly distributed over the Co(OH)xF2−x nanosheets, implying that the removal process did not fully remove SiO2 but rather left some redistributed over the nanosheets.

3.2. Post-Annealing Effects on Co(OH)xF2−x Nanosheets

For sensor applications, it is important to ensure the thermal stability of sensing materials since sensors are usually operated at elevated temperatures. Figure 5 exhibits SEM images of Co(OH)xF2−x nanosheets on a sensor device annealed at 350 °C for 2 h, showing that the nanosheet cluster structure remains stable. Figure 6a shows the fluorine content (F/Co*100) acquired from an XPS analysis for the samples of Co(OH)xF2−x nanosheets synthesized with SiO2 templates and without templates after annealing at various temperatures for 2 h. It is clear that the F ratio in the nanosheets decreases with the post-annealing temperature increasing, suggesting that F evaporates at high temperatures. It should be also noted that the sample grown on the template has retained fluorine more effectively at elevated temperatures than the sample without a template, implying that the SiO2 template has possibly played some role in retaining F.
Our FTIR analysis sheds some light on the possible influence of the SiO2 template on retaining fluorine, as shown in Figure 6b. In the FTIR data, the sample synthesized with SiO2 shows a strong peak at 1020 cm−1, corresponding to bonding between Si and F, which does not decline upon annealing at elevated temperatures [30]. In contrast, the no-template sample does not show any peak at 1020 cm−1 but a large one at 650 cm−1, which could be attributed to Co–O bonding [31]. These data indicate that the remnants of the SiO2 templates have possibly helped hold fluorine atoms (prone to evaporation) effectively by providing a strong bond with Si.
The TEM investigation further indicated no difference in the morphology of nanosheets before and after post-annealing at 350, 400, and 450 °C (see Figure S4). Figure 7 displays the high-resolution TEM images of the Co(OH)xF2−x nanosheets annealed at 350, 400, and 450 °C. A careful observation of the images of Figure 7a,d, and g reveals that the nanosheets have become nanocrystalline. Some of the nanocrystals are Co(OH)F while others belong to Co3O4 according to d-spacing analysis (see the images of the center and right-side columns). These results indicate that a significant part of the material has been transformed into Co3O4 upon post-annealing.

3.3. Gas-Sensing Characteristics of Co(OH)xF2−x Nanosheet Sensors

We evaluated the gas-sensing ability of the sensor based on the Co(OH)xF2−x nanosheets by examining the operating temperature, acetone gas concentration, gas species, and time elapse effects. Figure 8a shows gas responses at operating temperatures of 25, 43, 72, 150, 200, and 250 °C. The response graph exhibits a typical volcano shape with an optimized operating temperature of 200 °C. The response at 200 °C is around 269% toward acetone gas at a concentration of 4.5 ppm. For such a low acetone concentration, a response of 269% is quite high, compared with those of other reports on various cobalt-based materials for acetone gas sensor applications, as shown in Table 1.
Figure 8b shows dynamic sensing characteristics of the sensor upon the input of acetone gas at various concentrations at 200 °C. Calculated responses at each acetone concentration were also plotted (see Figure S5). The limit of detection of the sensor toward acetone gas is 0.04 ppm with a response of 17%. The sensor also shows a good selectivity toward other gases such as CH3COCH3 (2.2 ppm), C6H5CH3 (3.6 ppm), CO (0.8 ppm), H2 (0.8 ppm), and NO2 (1.1 ppm) at 200 °C, as shown in Figure 8c. Figure 8d shows responses to 4.5-ppm acetone gas as a function of time elapse for 20 days. The response of the Co(OH)xF2−x nanosheet-based sensor degrades from 269 at the first measurement to 106% after 6 days and stayed at a similar level during following measurements.

3.4. Gas-Sensing Characteristics of Annealed Co(OH)xF2−x Nanosheet Sensors and Pt-Doping Effect

Next, we examined if the sensor functioned effectively at high temperatures (≤450 °C). We investigated dynamic sensing characteristics by changing the operating temperature between 200 and 300 °C toward acetone gas (4.5 ppm), as shown in Figure 9a,b. In the first cycle, the sensor shows a stable response and recovery at 200 °C, whereas it displays an unstable behavior at 250 and 300 °C [14]. We also noticed some color change of the material (from pink to dark brown) on the device during measurements at elevated temperatures (see the photographs in Figure 9a), an indication of the deterioration of the material, possibly due to fluorine evaporation [20].
We tested the sensors after post-annealing at 300 to 450 °C to further investigate the temperature effects on sensor performance. Figure 9c displays the gas responses of sensors annealed at 300, 350, 400, and 450 °C for 2 or 4 h toward 4.5 ppm acetone at 300 °C, which show much lower responses than those measured before post-annealing. It should be noted that the trend of the response, decreasing with the post-annealing temperature increasing, is similar to that of the fluorine concentration with temperature (see Figure 6a), hinting that the loss of fluorine possibly influenced the response to acetone gas. Increasing the post-annealing time from 2 to 4 h also deteriorates the gas response to 28 and 14% at 300 and 350 °C, respectively, as shown in Figure 9c.
Using the sample annealed at 350 °C for 2 h, we further performed a dynamic test, finding that although the response is low, the sensor works reliably with a full recovery for three successive gas on/off cycles at 300 °C, as shown in Figure 9d. Furthermore, additional annealing at 300 °C for 2 h does not decrease the response, as shown in Figure 9e, suggesting that once the sensing material is annealed at 350 °C, further annealing at a lower temperature would not affect the response much. The result of a long-term stability test shown in Figure 9d also exhibits stable operations with a similar response level for around 3 months. Thus, the annealed sensor at 350 °C shows stable responses although the response level is significantly lower, as compared to the performance of the non-annealed sensor.
It is well known that noble metal dopants such as Pt, Pd, and Au could enhance gas sensor responses because of their strong catalytic properties [43]. In an effort to recover the deteriorated gas response of the 350 °C sample, we loaded Pt (1 nm) on the sensor and conducted measurements for performance (Figure 10). Figure 10a clearly shows improvements in responses at all working temperatures after loading Pt to the device. The sensor also demonstrated a good selectivity toward 4.5 ppm of acetone gas, compared to 4.5 ppm of other gases such as C6H5CH3, CO, H2, and NO2, as shown in Figure 10b. Figure 10c presents a good sensitivity of the sensor toward acetone at a concentration of 1.1, 2.2, 3.4, and 4.5 ppm. Dynamic sensing characteristics in Figure 10d displays responses of stepwise increase upon the injection of acetone with the concentration increasing from 1.1 to 4.5 ppm, showing a sensitive response at each step and full recovery upon stopping the injection of acetone gas. These data confirm that the Pt loading on the sensor device has enhanced the sensing response, enabling a sensor with a high sensitivity and good stability.

3.5. Gas-Sensing Mechanisms

In this study, we synthesized Co(OH)xF2−x nanosheets, a material successfully incorporating a high concentration of fluorine and keeping it to some extent through annealing at elevated temperatures (owing to the support from the bond with Si of the remnant SiO2). It is clear from our sensor data that the presence of fluorine of a high concentration helped achieve a high gas response. In comparison, a sensor based on a material fabricated without adding NH4F in the synthesis process showed a low response (92%) (see Figure S6), highlighting the effects of F.
It should be noted that incorporated F, with a high electronegativity, could enhance the Lewis acidity of the material, which could facilitate the adsorption of gas molecules [15,16,19]. For the present study, the main reactions occurring at the surface of the Co(OH)xF2−x nanosheets can be expressed as shown below [1,44,45]. The high Lewis acidity of the nanosheets with a high concentration of F has possibly helped the material accept electrons easily and promoted the reaction of Equation (4).
O2(gas) → O2(ads)
O2(ads) + e → O2(ads) (T < 100 °C)
O2(ads) + e → 2O(ads) (100 °C < T < 300 °C)
CH3COCH3(gas) + 8O(ads) → 3CO2 + 3H2O + 8e
e + h+ → Null
In addition, the morphology of the Co(OH)xF2−x nanosheets with a high specific surface area (193.38 m2/g), brought about by the assistance of the SiO2 template, has possibly contributed to the high sensing performance. For example, the high response of the Co(OH)xF2−x nanosheets (269%) toward 4.5 ppm of acetone gas at 200 °C contrasts with the low response (147%) of the sensor based on Co(OH)xF2−x fabricated without SiO2 templates, which has a bulk-type morphology (see Figure S7). Other studies also show the strong influence of morphology on the sensitivity of sensors, attributing to the increase in oxygen-adsorption sites promoting the reaction of Equation (1) [46,47]. Thus, the adoption of an SiO2 template in the synthesis process played a beneficial role in two ways, maintaining a high F concentration for a high catalytic property and producing a structure with a very high specific surface area.

4. Conclusions

We synthesized Co(OH)xF2−x nanosheets by adopting SiO2 sphere templates and fabricated sensor devices based on the materials. Our material characterization using such analytical tools as XPS, FTIR, and TEM confirms that the material is made up of Co(OH)xF2−x nanosheets containing a high concentration of F with an amorphous microstructure with some short-range ordering. The sensor based on the Co(OH)xF2−x nanosheets exhibited a very high response of 269% toward 4.5 ppm of acetone gas at an operation temperature of 200 °C, and a minimum detection limit of 40 ppb was obtained. Although the sensor showed some indications of degradation (possibly due to F evaporation) at a temperature higher than 300 °C, the Pt doping was found to help recover gas response to some extent (77% toward 4.5 ppm of acetone). We discussed the strong catalytic property of F in a high concentration as well as the high specific surface area due to the morphology of densely clustered nanosheets being responsible for the excellent sensor performance. This study highlights the possibility of further applications of a new material group of metal fluoride hydroxides to gas sensors owing to their distinct advantages compared to conventional metal oxide materials.

Supplementary Materials

The Supplementary Materials are available online at https://www.mdpi.com/2073-4352/10/11/968/s1. Figure S1. EDS spectrum of Co(OH)xF2−x nanosheets with atomic percentages of each element. Figure S2. Profile image of the interdigitated sensor device with coated Co(OH)xF2−x nanosheets powders. Figure S3. N2 adsorption-desorption isotherm with calculated BET surface area (193.38 m2/g). Figure S4. TEM bright-field images of the Co(OH)xF2−x nanosheets (a) before and after annealing at (b) 350, (c) 400, and (d) 450 °C for 2 h. Figure S5. Responses of Co(OH)xF2−x nanosheets-based sensor toward acetone gas of various concentrations at 200 °C. Figure S6. Responses of a material synthesized without NH4F (other synthesis procedures were same with the Co(OH)xF2−x nanosheets) at various temperatures toward 4.5 ppm of acetone (Responses of this material at temperatures lower than 150 °C were immeasurable because of high resistance). Figure S7. Responses of Co(OH)xF2−x without SiO2 templates at various temperatures toward 4.5 ppm of acetone.

Author Contributions

Conceptualization, Y.Y. and T.-y.E.; methodology, Y.Y. and T.-y.E.; software, S.X. and P.J.Y.; validation, J.-S.P. and H.-J.L.; formal analysis, T.-y.E.; investigation, T.-y.E.; resources, J.-S.P. and H.-J.L.; data curation, T.-y.E., J.-S.P. and H.-J.L.; writing—original draft preparation, T.-y.E.; writing—review and editing, Y.Y., C.Y., J.-S.P. and H.-J.L.; visualization, S.X. and P.J.Y.; supervision, J.-S.P. and H.-J.L.; project administration, J.-S.P. and H.-J.L.; funding acquisition, J.-S.P. and H.-J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Basic Science Institute (KBSI) National Research Facilities and Equipment Center (NFEC) grant funded by the Korea government (Ministry of Education) (No. 2019R1A6C1030010). This work was also financially supported by National Natural Science Foundation of China (NSFC, 61904190). Additionally, J.-S. Park was partially supported by Project No. P0006858 of International Collaboration Program by KIAT and MOTIE in Rep. Korea.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Z.; Zhu, L.; Wen, Z.; Ye, Z. Controllable synthesis of Co3O4 crossed nanosheet arrays toward an acetone gas sensor. Sens. Actuators B Chem. 2017, 238, 1052–1059. [Google Scholar] [CrossRef]
  2. Kao, K.-W.; Hsu, M.-C.; Chang, Y.-H.; Gwo, S.; Yeh, J.A. A sub-ppm acetone gas sensor for diabetes detection using 10 nm thick ultrathin InN FETs. Sensors 2012, 12, 7157–7168. [Google Scholar] [CrossRef] [Green Version]
  3. Xie, Y.; Xing, R.; Li, Q.; Xu, L.; Song, H. Three-dimensional ordered ZnO–CuO inverse opals toward low concentration acetone detection for exhaled breath sensing. Sens. Actuators B Chem. 2015, 211, 255–262. [Google Scholar] [CrossRef]
  4. Barsan, N.; Koziej, D.; Weimar, U. Metal oxide-based gas sensor research: How to? Sens. Actuators B Chem. 2007, 121, 18–35. [Google Scholar] [CrossRef]
  5. Zeng, Y.; Zhang, T.; Yuan, M.; Kang, M.; Lu, G.; Wang, R.; Fan, H.; He, Y.; Yang, H. Growth and selective acetone detection based on ZnO nanorod arrays. Sens. Actuators B Chem. 2009, 143, 93–98. [Google Scholar] [CrossRef]
  6. Kou, X.; Xie, N.; Chen, F.; Wang, T.; Guo, L.; Wang, C.; Wang, Q.; Ma, J.; Sun, Y.; Zhang, H. Superior acetone gas sensor based on electrospun SnO2 nanofibers by Rh doping. Sens. Actuators B Chem. 2018, 256, 861–869. [Google Scholar] [CrossRef]
  7. Koo, W.-T.; Yu, S.; Choi, S.-J.; Jang, J.-S.; Cheong, J.Y.; Kim, I.-D. Nanoscale PdO catalyst functionalized Co3O4 hollow nanocages using MOF templates for selective detection of acetone molecules in exhaled breath. ACS Appl. Mater. Interfaces 2017, 9, 8201–8210. [Google Scholar] [CrossRef]
  8. Li, W.-Y.; Xu, L.-N.; Chen, J. Co3O4 nanomaterials in lithium-ion batteries and gas sensors. Adv. Funct. Mater. 2005, 15, 851–857. [Google Scholar] [CrossRef]
  9. Wan, S.; Qi, J.; Zhang, W.; Wang, W.; Zhang, S.; Liu, K.; Zheng, H.; Sun, J.; Wang, S.; Cao, R. Hierarchical Co(OH)F Superstructure Built by Low-Dimensional Substructures for Electrocatalytic Water Oxidation. Adv. Mater. 2017, 29, 1700286. [Google Scholar] [CrossRef]
  10. Zhou, T.; Zhang, T.; Deng, J.; Zhang, R.; Lou, Z.; Wang, L. P-type Co3O4 nanomaterials-based gas sensor: Preparation and acetone sensing performance. Sens. Actuators B Chem. 2017, 242, 369–377. [Google Scholar] [CrossRef]
  11. Kang, J.-S.; Park, J.-S.; An, B.-S.; Yang, C.-W.; Lee, H.-J. Nickel Doping on Cobalt Oxide Thin Film Using by Sputtering Process-a Route for Surface Modification for p-type Metal Oxide Gas Sensors. J. Korean Phys. Soc. 2018, 73, 1867–1872. [Google Scholar] [CrossRef]
  12. Vladimirova, S.; Krivetskiy, V.; Rumyantseva, M.; Gaskov, A.; Mordvinova, N.; Lebedev, O.; Martyshov, M.; Forsh, P. Co3O4 as p-type material for CO sensing in humid air. Sensors 2017, 17, 2216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhang, C.; Li, L.; Hou, L.; Chen, W. Fabrication of Co3O4 nanowires assembled on the surface of hollow carbon spheres for acetone gas sensing. Sens. Actuators B Chem. 2019, 291, 130–140. [Google Scholar] [CrossRef]
  14. Xu, K.; Zhao, W.; Yu, X.; Duan, S.; Zeng, W. MOF-derived Co3O4/Fe2O3 pn hollow cubes for improved acetone sensing characteristics. Phys. E Low Dimens. Syst. Nanostruct. 2020, 118, 113869. [Google Scholar] [CrossRef]
  15. Barreca, D.; Bekermann, D.; Comini, E.; Devi, A.; Fischer, R.A.; Gasparotto, A.; Gavagnin, M.; Maccato, C.; Sada, C.; Sberveglieri, G. Plasma enhanced-CVD of undoped and fluorine-doped Co3O4 nanosystems for novel gas sensors. Sens. Actuators B Chem. 2011, 160, 79–86. [Google Scholar] [CrossRef]
  16. Zhao, Y.; Du, X.; Wang, X.; He, J.; Yu, Y.; He, H. Effects of F doping on TiO2 acidic sites and their application in QCM based gas sensors. Sens. Actuators B Chem. 2010, 151, 205–211. [Google Scholar] [CrossRef]
  17. Song, M.; Zhang, Z.; Li, Q.; Jin, W.; Wu, Z.; Fu, G.; Liu, X. Ni-foam supported Co(OH)F and Co–P nanoarrays for energy-efficient hydrogen production via urea electrolysis. J. Mater. Chem. A 2019, 7, 3697–3703. [Google Scholar] [CrossRef]
  18. Wang, Z.; Liu, Z.; Du, G.; Asiri, A.M.; Wang, L.; Li, X.; Wang, H.; Sun, X.; Chen, L.; Zhang, Q. Ultrafine PtO2 nanoparticles coupled with a Co(OH)F nanowire array for enhanced hydrogen evolution. Chem. Commun. 2018, 54, 810–813. [Google Scholar] [CrossRef] [PubMed]
  19. Lv, J.; Yang, X.; Zang, H.-Y.; Wang, Y.-H.; Li, Y.-G. Ultralong needle-like N-doped Co(OH)F on carbon fiber paper with abundant oxygen vacancies as an efficient oxygen evolution reaction catalyst. Mater. Chem. Front. 2018, 2, 2045–2053. [Google Scholar] [CrossRef]
  20. Wen, Z.; Zhu, L.; Mei, W.; Hu, L.; Li, Y.; Sun, L.; Cai, H.; Ye, Z. Rhombus-shaped Co3O4 nanorod arrays for high-performance gas sensor. Sens. Actuators B Chem. 2013, 186, 172–179. [Google Scholar] [CrossRef]
  21. Wang, W.; Gu, B.; Liang, L.; Hamilton, W.A. Fabrication of near-infrared photonic crystals using highly-monodispersed submicrometer SiO2 spheres. J. Phys. Chem. B 2003, 107, 12113–12117. [Google Scholar] [CrossRef]
  22. Simmonov, Y.A.; Kritskii, A.A.; Rychkov, V.N.; Tomashov, V.A. Investigation of the treatment process of associated dioxides of zirconium and silicon by an aqueous solution of ammonium fluoride. Rus. J. Non Ferr. Met. 2010, 51, 320–323. [Google Scholar] [CrossRef]
  23. Nundy, S.; Eom, T.-Y.; Song, K.-Y.; Park, J.-S.; Lee, H.-J. Hydrothermal synthesis of mesoporous ZnO microspheres as NOX gas sensor materials—Calcination effects on microstructure and sensing performance. Ceram. Int. 2020, 46, 19354–19364. [Google Scholar] [CrossRef]
  24. Yang, J.; Liu, H.; Martens, W.N.; Frost, R.L. Synthesis and characterization of cobalt hydroxide, cobalt oxyhydroxide, and cobalt oxide nanodiscs. J. Phys. Chem. C 2010, 114, 111–119. [Google Scholar] [CrossRef]
  25. Kong, L.-B.; Liu, M.-C.; Lang, J.-W.; Liu, M.; Luo, Y.-C.; Kang, L. Porous cobalt hydroxide film electrodeposited on nickel foam with excellent electrochemical capacitive behavior. J. Solid State Electrochem. 2011, 15, 571–577. [Google Scholar] [CrossRef]
  26. Zhou, W.-J.; Zhang, J.; Xue, T.; Zhao, D.-D.; Li, H.-L. Electrodeposition of ordered mesoporous cobalt hydroxide film from lyotropic liquid crystal media for electrochemical capacitors. J. Mater. Chem. 2008, 18, 905–910. [Google Scholar] [CrossRef]
  27. Niu, Y.; Wang, R.; Jiao, W.; Ding, G.; Hao, L.; Yang, F.; He, X. MoS2 graphene fiber based gas sensing devices. Carbon 2015, 95, 34–41. [Google Scholar] [CrossRef]
  28. Zhang, D.; Liu, A.; Chang, H.; Xia, B. Room-temperature high-performance acetone gas sensor based on hydrothermal synthesized SnO2-reduced graphene oxide hybrid composite. RSC Adv. 2015, 5, 3016–3022. [Google Scholar] [CrossRef]
  29. Yang, C.; Xiao, F.; Wang, J.; Su, X. 3D flower-and 2D sheet-like CuO nanostructures: Microwave-assisted synthesis and application in gas sensors. Sens. Actuators B Chem. 2015, 207, 177–185. [Google Scholar] [CrossRef]
  30. Liptak, R.; Yang, J.; Kramer, N.; Kortshagen, U.; Campbell, S. Environmental photostability of SF6-etched silicon nanocrystals. Nanotechnology 2012, 23, 395205. [Google Scholar] [CrossRef]
  31. Adam, F.; Thiam-Seng, C. The Production of Hydrogen by Catalytic Oxidation of Hydrocarbons using a Rice Husk Based Cobalt Phyllosilicate Catalyst. J. Phys. Sci. 2013, 24, 61. [Google Scholar]
  32. Xu, J.; Zhang, J.; Wang, B.; Liu, F. Shape-regulated synthesis of cobalt oxide and its gas-sensing property. J. Alloys Compd. 2015, 619, 361–367. [Google Scholar] [CrossRef]
  33. Nguyen, H.; El-Safty, S.A. Meso-and macroporous Co3O4 nanorods for effective VOC gas sensors. J. Phys. Chem. C 2011, 115, 8466–8474. [Google Scholar] [CrossRef]
  34. Zhang, R.; Zhou, T.; Wang, L.; Zhang, T. Metal–organic frameworks-derived hierarchical Co3O4 structures as efficient sensing materials for acetone detection. ACS Appl. Mater. Interfaces 2018, 10, 9765–9773. [Google Scholar] [CrossRef]
  35. Wang, S.; Cao, J.; Cui, W.; Fan, L.; Li, X.; Li, D. Facile synthesis of bamboo raft-like Co3O4 with enhanced acetone gas sensing performances. J. Alloys Compd. 2018, 758, 45–53. [Google Scholar] [CrossRef]
  36. Zhang, Z.; Wen, Z.; Ye, Z.; Zhu, L. Gas sensors based on ultrathin porous Co3O4 nanosheets to detect acetone at low temperature. RSC Adv. 2015, 5, 59976–59982. [Google Scholar] [CrossRef]
  37. Lin, Y.; Ji, H.; Shen, Z.; Jia, Q.; Wang, D. Enhanced acetone sensing properties of Co3O4 nanosheets with highly exposed (111) planes. J. Mater. Sci. Mater. Electron. 2016, 27, 2086–2095. [Google Scholar] [CrossRef]
  38. Khandekar, M.; Tarwal, N.; Mulla, I.; Suryavanshi, S. Nanocrystalline Ce doped CoFe2O4 as an acetone gas sensor. Ceram. Int. 2014, 40, 447–452. [Google Scholar] [CrossRef]
  39. Zhang, D.; Yang, Z.; Wu, Z.; Dong, G. Metal-organic frameworks-derived hollow zinc oxide/cobalt oxide nanoheterostructure for highly sensitive acetone sensing. Sens. Actuators B Chem. 2019, 283, 42–51. [Google Scholar] [CrossRef]
  40. Xiong, Y.; Zhu, Z.; Ding, D.; Lu, W.; Xue, Q. Multi-shelled ZnCo2O4 yolk-shell spheres for high-performance acetone gas sensor. Appl. Surf. Sci. 2018, 443, 114–121. [Google Scholar] [CrossRef]
  41. Alali, K.T.; Liu, J.; Liu, Q.; Li, R.; Zhang, H.; Aljebawi, K.; Liu, P.; Wang, J. Enhanced acetone gas sensing response of ZnO/ZnCo2O4 nanotubes synthesized by single capillary electrospinning technology. Sens. Actuators B Chem. 2017, 252, 511–522. [Google Scholar] [CrossRef]
  42. Qu, F.; Thomas, T.; Zhang, B.; Zhou, X.; Zhang, S.; Ruan, S.; Yang, M. Self-sacrificing templated formation of Co3O4/ZnCo2O4 composite hollow nanostructures for highly sensitive detecting acetone vapor. Sens. Actuators B Chem. 2018, 273, 1202–1210. [Google Scholar] [CrossRef]
  43. Wang, Y.-D.; Wu, X.-H.; Su, Q.; Li, Y.-F.; Zhou, Z.-L. Ammonia-sensing characteristics of Pt and SiO2 doped SnO2 materials. Solid State Electron. 2001, 45, 347–350. [Google Scholar] [CrossRef]
  44. Lim, S.K.; Hwang, S.-H.; Kim, S.; Park, H. Preparation of ZnO nanorods by microemulsion synthesis and their application as a CO gas sensor. Sens. Actuators B Chem. 2011, 160, 94–98. [Google Scholar] [CrossRef]
  45. Ge, Y.; Kan, K.; Yang, Y.; Zhou, L.; Jing, L.; Shen, P.; Li, L.; Shi, K. Highly mesoporous hierarchical nickel and cobalt double hydroxide composite: Fabrication, characterization and ultrafast NOx gas sensors at room temperature. J. Mater. Chem. A 2014, 2, 4961–4969. [Google Scholar] [CrossRef]
  46. Gao, T.; Wang, T. Synthesis and properties of multipod-shaped ZnO nanorods for gas-sensor applications. Appl. Phys. A 2005, 80, 1451–1454. [Google Scholar] [CrossRef]
  47. Bie, L.-J.; Yan, X.-N.; Yin, J.; Duan, Y.-Q.; Yuan, Z.-H. Nanopillar ZnO gas sensor for hydrogen and ethanol. Sens. Actuators B Chem. 2007, 126, 604–608. [Google Scholar] [CrossRef]
Figure 1. Illustration of the synthesis process of Co(OH)xF2−x nanosheets.
Figure 1. Illustration of the synthesis process of Co(OH)xF2−x nanosheets.
Crystals 10 00968 g001
Figure 2. SEM images of (a,b) SiO2 spheres and (c,d) Co(OH)xF2−x nanosheets.
Figure 2. SEM images of (a,b) SiO2 spheres and (c,d) Co(OH)xF2−x nanosheets.
Crystals 10 00968 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of the Co(OH)xF2−x nanosheets ((a) F 1s, (b) Co 2p, and (c) O 1s), and X-ray diffraction (XRD) pattern of the Co(OH)xF2−x nanosheets (d).
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of the Co(OH)xF2−x nanosheets ((a) F 1s, (b) Co 2p, and (c) O 1s), and X-ray diffraction (XRD) pattern of the Co(OH)xF2−x nanosheets (d).
Crystals 10 00968 g003
Figure 4. Transmission Electron Microscopy (TEM) analysis of the Co(OH)xF2−x nanosheets: (a,b) bright-field images, (c) electron dispersive spectroscopy (EDS) mapping, and (d,e) high-resolution images. An inset image in (d) is a FFT (Fast Fourier Transform) pattern of the high-resolution image.
Figure 4. Transmission Electron Microscopy (TEM) analysis of the Co(OH)xF2−x nanosheets: (a,b) bright-field images, (c) electron dispersive spectroscopy (EDS) mapping, and (d,e) high-resolution images. An inset image in (d) is a FFT (Fast Fourier Transform) pattern of the high-resolution image.
Crystals 10 00968 g004
Figure 5. (a) High- and (b) low-resolution SEM images of the Co(OH)xF2−x nanosheets annealed at 350 °C for 2 h on device.
Figure 5. (a) High- and (b) low-resolution SEM images of the Co(OH)xF2−x nanosheets annealed at 350 °C for 2 h on device.
Crystals 10 00968 g005
Figure 6. (a) Annealing temperature effects on the contents of fluorine (F/Co*100), acquired from XPS, in the nanosheets synthesized with and without SiO2 template, and (b) comparison of the Fourier transform infrared spectroscopy (FTIR) spectra of the samples.
Figure 6. (a) Annealing temperature effects on the contents of fluorine (F/Co*100), acquired from XPS, in the nanosheets synthesized with and without SiO2 template, and (b) comparison of the Fourier transform infrared spectroscopy (FTIR) spectra of the samples.
Crystals 10 00968 g006
Figure 7. High-resolution TEM images of the Co(OH)xF2−x nanosheets annealed at (ac) 350, (df) 400, and (gj) 450 °C for 2 h. (b,e,h) and (c,f,i) are the magnified images of the blue and red squares in (a,d,g), respectively.
Figure 7. High-resolution TEM images of the Co(OH)xF2−x nanosheets annealed at (ac) 350, (df) 400, and (gj) 450 °C for 2 h. (b,e,h) and (c,f,i) are the magnified images of the blue and red squares in (a,d,g), respectively.
Crystals 10 00968 g007
Figure 8. Sensing characteristics of a sensor based on Co(OH)xF2−x nanosheets: (a) responses at various operating temperatures toward 4.5 ppm of acetone gas, (b) resistance changes upon input of acetone gas of a various concentration at 200 °C, (c) selectivity toward CH3COCH3 (2.2 ppm), C6H5CH3 (3.6 ppm), CO (0.8 ppm), H2 (0.8 ppm), and NO2 (1.1 ppm) at 200 °C, and (d) stability (for 20 days) toward 4.5 ppm of acetone at 200 °C.
Figure 8. Sensing characteristics of a sensor based on Co(OH)xF2−x nanosheets: (a) responses at various operating temperatures toward 4.5 ppm of acetone gas, (b) resistance changes upon input of acetone gas of a various concentration at 200 °C, (c) selectivity toward CH3COCH3 (2.2 ppm), C6H5CH3 (3.6 ppm), CO (0.8 ppm), H2 (0.8 ppm), and NO2 (1.1 ppm) at 200 °C, and (d) stability (for 20 days) toward 4.5 ppm of acetone at 200 °C.
Crystals 10 00968 g008
Figure 9. (a,b) Dynamic sensing characteristics upon changes of applied temperature toward 4.5 ppm of acetone gas. (c) Responses of annealed Co(OH)xF2−x nanosheet sensors with various post-annealing conditions. Dynamic sensing characteristics of the sample after post-annealing at 350 °C for 2 h (d) and after additional annealing at 300 °C for 2 h (e). (f) Long-term stability test of the 350 °C sample for 118 days toward 4.5 ppm of acetone gas at 300 °C.
Figure 9. (a,b) Dynamic sensing characteristics upon changes of applied temperature toward 4.5 ppm of acetone gas. (c) Responses of annealed Co(OH)xF2−x nanosheet sensors with various post-annealing conditions. Dynamic sensing characteristics of the sample after post-annealing at 350 °C for 2 h (d) and after additional annealing at 300 °C for 2 h (e). (f) Long-term stability test of the 350 °C sample for 118 days toward 4.5 ppm of acetone gas at 300 °C.
Crystals 10 00968 g009
Figure 10. Sensing characteristics of Pt-doped Co(OH)xF2−x nanosheet sensors annealed at 350 °C for 2 h. Responses at various operating temperatures compared with a pristine sample toward 4.5 ppm of acetone gas (a), toward various gases of 4.5 ppm (CH3COCH3, C6H5CH3, CO, H2, and NO2) at 300 °C (b), and toward various acetone concentrations at 300 °C (c). (d) Dynamic sensing characteristics of the sensor toward an increasing acetone gas concentration at 350 °C.
Figure 10. Sensing characteristics of Pt-doped Co(OH)xF2−x nanosheet sensors annealed at 350 °C for 2 h. Responses at various operating temperatures compared with a pristine sample toward 4.5 ppm of acetone gas (a), toward various gases of 4.5 ppm (CH3COCH3, C6H5CH3, CO, H2, and NO2) at 300 °C (b), and toward various acetone concentrations at 300 °C (c). (d) Dynamic sensing characteristics of the sensor toward an increasing acetone gas concentration at 350 °C.
Crystals 10 00968 g010
Table 1. Comparison of the performances of acetone sensors using Co-based materials reported in the literature.
Table 1. Comparison of the performances of acetone sensors using Co-based materials reported in the literature.
MaterialOperating Temp. (°C)Concentration (ppm)Response (Rg/Ra)Reference
Hierarchical and hexagonal Co3O424021.32[32]
Co3O4 crossed nanosheet arrays111206.8[1]
Meso- and macroporous Co3O4 nanorods30074,57018.5[33]
Rhombus-shaped Co3O4 nanorod arrays16050020.1[20]
Co3O4 core-shell19020013[34]
Bamboo raft-like Co3O418020010.5[35]
Ultrathin porous Co3O4 nanosheets15010011.4[36]
Co3O4 nanosheets1601006.1[37]
Ce-doped CoFe2O422550001.77[38]
Hollow ZnO/Co3O430011.5[39]
Multi-shelled ZnCo2O42000.51.36[40]
PdO Catalyst functionalized Co3O4 hollow nanocages35052.51[7]
ZnO/ZnCo2O4 nanotubes1755015[41]
Co3O4/ZnCo2O425510016.3[42]
Co(OH)xF2−x nanosheets2004.53.69 (269%)Our study
1.12.17 (117%)
0.041.17 (17%)
Pt-doped Co(OH)xF2−x nanosheets-annealed at 300 °C3004.51.77 (77%)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, Y.; Eom, T.-y.; Xu, S.; Yoo, P.J.; Yan, C.; Park, J.-S.; Lee, H.-J. Development of Co(OH)xF2−x Nanosheets for Acetone Gas Sensor Applications: Material Characterization and Sensor Performance Evaluation. Crystals 2020, 10, 968. https://doi.org/10.3390/cryst10110968

AMA Style

Yan Y, Eom T-y, Xu S, Yoo PJ, Yan C, Park J-S, Lee H-J. Development of Co(OH)xF2−x Nanosheets for Acetone Gas Sensor Applications: Material Characterization and Sensor Performance Evaluation. Crystals. 2020; 10(11):968. https://doi.org/10.3390/cryst10110968

Chicago/Turabian Style

Yan, Yaping, Tae-yil Eom, Shiyu Xu, Pil J. Yoo, Changzeng Yan, Joon-Shik Park, and Hoo-Jeong Lee. 2020. "Development of Co(OH)xF2−x Nanosheets for Acetone Gas Sensor Applications: Material Characterization and Sensor Performance Evaluation" Crystals 10, no. 11: 968. https://doi.org/10.3390/cryst10110968

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