Electrochemistry
Online ISSN : 2186-2451
Print ISSN : 1344-3542
ISSN-L : 1344-3542
Articles
Study on Fabrications and Storage Capacity of Coal Tar Pitch Based V2O3@C Composite Materials
Jianke LIXincheng MIAOShenhao WANGShaobei CHENBeibei HANGuiying XUKun WANGBaigang ANDongying JUWeimin ZHOU
Author information
JOURNAL OPEN ACCESS FULL-TEXT HTML
J-STAGE Data

2022 Volume 90 Issue 7 Pages 077002

Details
Abstract

The coal tar pitch based V2O3@C composite materials were successfully fabricated by a hydrothermal method using the V2O5, H2C2O4·2H2O and coal tar pitches dissolved in water. It is observed that changing the dosages of coal tar pitches dissolved in water can lead to the various carbon contents in V2O3@C composite materials, resulting in the V2O3@C composite materials possessing different electrochemical performances. As a result, when the carbon content is adjusted to 42.7 %, the prepared V2O3@C-2.0 shows more excellent electrochemical performances than other prepared V2O3@C composite materials. For instance, the V2O3@C-2.0 exhibits the high Li+ storage capacity at 580.2 mAh g−1, after charge-discharge was carried out 100 cycles at 0.1 A g−1. Surprisingly, the V2O3@C-2.0 still shows the Li+ storage capacity at 234 mAh g−1, after 500 cycles at 5.0 A g−1. The prepared V2O3@C composite materials show an excellent industrial application perspective because coal tar pitches are the generally industrial crude materials.

1. Introduction

Accommodating the increase in demands for lithium-ion batteries (LIBs) having high electrochemical performances such as high storage capacity, excellent cycling performance, tremendous rate performance and so on, the issues that how to ameliorate the performances of LIBs becomes a popular topic more than ever before.1 Thus, a lot of positive and negative electrodes are extensively unfolded, with a view to enhancing the electrochemical performances of LIBs.24

It is acknowledged that graphite is extensively utilized as a negative electrode material in the fabrication of LIBs, however, its characteristic properties restrict it to satisfy the demands developing the LIBs with excellent performances.5 In recent years, a number of methods are investigated to replace traditional graphite. Among them, metal oxide material is an attracted material because they possess an exceedingly high storage capacity, compared to graphite.610 Nevertheless, the poor conductivity and volume expansion of metal oxides influence their storage capacity.11

In particular, in view of that the V2O3 possesses many advantages such as low toxicity, high theoretical capacity (1070 mAh g−1), cost-effectiveness and so on, which draws public attention to V2O3 used as an electrode material.1214 Especially, the V-O-V three-dimensional structures of V2O3 cause that electrons are capable to facilely transfer among the V-O-V structures, which leads to the V2O3 possessing metallic conductivity. Nevertheless, the collapse problem during the charge-discharge process also exists in V2O3 compounds.

Covering the carbons on the surface of V2O3 is an effective methods to address the aforementioned problem.15 On the basis of a view point that soft carbons have excellent conductivity, pitches based carbons show extensive popularity as a representative of soft carbons.16 In our studies, the water soluble coal tar pitches (WS-CTPs) are fabricated using the coal tar pitches (CTPs) by mixed acid methods.17 However, the agglomeration phenomenon of metal oxides could not be avoided very well, which influenced the enhancement of electrochemical performances of metal oxides in composite materials. It is observed that the carbon contents in V2O3@C composite materials can be changed by controlling the dosages of WS-CTPs in the reactive cases.

The electrochemical performances of prepared V2O3@C materials are evaluated in detail. As a result, after cycling the charge-discharge 100 cycles at 0.1 A g−1, the V2O3@C-2.0 with carbon contents (39.6 %) shows a high cycling performance of 580.2 mAh g−1, which is much higher than the V2O3 materials. Additionally, it is found that controlling the carbon contents on the surface of V2O3@C materials is an efficacious method to improve the long circulating performance of V2O3. The electrochemical measurements are indicative of that WS-CTPs are able to become an effective carbon source in the field for fabricating the metal oxide @C materials.

2. Experimental

2.1 Characteristics

The measurements of X-ray diffraction (XRD) used an X’pert Powder instrument from PANalytical. The X-ray photoelectron spectroscopy (XPS) measurements were carried out on a K-Alpha instrument from Thermo Fisher Scientific, USA. Nitrogen adsorption and desorption isotherms were measured by a Quadrasorb autosorb-iQ surface analyzer which was purchased from Quantachrome Instruments, USA. Specific surface areas were determined in detail, according to the Brunauer-Emmett-Teller (BET) method. The pore size distribution was assessed by a DFT model for slit pores. Morphologies were evaluated by scanning electron microscopy (SEM) using an instrument produced by Carl Zeiss AG, Germany. The FT-IR measurements were performed on an ALPHA II spectrophotometer (Bruker Co. Ltd., Karlsruhe, Badensko-wuertembersko, Germany). Electrochemical measurements were performed by the CHI660E electrochemical system (ChenHua, Shanghai, China). Thermogravimetric analysis (TGA) measurements were carried out by using a TG-DTA8122 (Rigaku, Japan).

2.2 Preparations of water-solving pitches (WS-CTPs)

The CTPs were firstly ground under 75 µm. After adding the ground coal tar pitches (3 g) into the beakers containing the acid mixtures which were obtained by mixing the nitric acid (concentration: 65 %) and sulfuric acid (concentration: 98 %) in mixing ratio of 3 : 7 (v/v), the obtained mixture was mixed uniformly. The same uniform mixture was placed in a reactor, and then the reactions were performed at 90 °C for 3 h. After filtration, the obtained solids were washed by deionized water until the pH value of the filtrate became 7. The obtained solids were added into the 1 M NaOH (200 mL), and the pH value was adjusted over 12. The obtained mixture was stirred (350 r/min) at 80 °C for 3 h, and then the mixture was filtered. Finally, the HCl solution (1 M) was slowly dropped in the filtration until the pH value became 2. At the same time, the precipitations appeared in the same mixture. After centrifugation, the obtained solids were washed by deionized water until pH value of solution became 5–6. The same solid was placed in the oven, and dried for 12 h at 80 °C.

2.3 Preparation of V2O3@C composite materials

The mixture constructed by V2O5 (0.36 g) and H2C2O4·2H2O (0.76 g) with a mole ratio of 1 : 3 was dissolved in a beaker containing the deionized water (15 mL). The beaker was placed in a water-bath, and the mixture was stirred for 0.5 h at 70 °C. The color of solution changed from green to dark blue, and the vanadyl oxalate produced in the same solution.

The WS-CTPs (0.1 g, 0.2 g and 0.4 g) were respectively dissolved in the ethylene glycol (60 mL), and the obtained mixtures were placed in the water batches respectively. Simultaneously, these solutions were stirred for 1 h at 70 °C. The obtained mixtures containing WS-CTPs and prepared vanadyl oxalate solutions were together added into a reactor, in which the reactions were carried out for 12 h at 180 °C.

After the reacting process, the reacted mixtures were filtered, and the obtained solids were washed by deionized water and anhydrous alcohol several times. Then the solids were placed in an oven, and dried for 12 h at 80 °C. Thereafter, the same solids were placed in the vacuum tube furnace, where mixture gas (N2/H2 = 9 : 1) was being followed continuously. The temperature of vacuum tube furnace was raised to 600 °C with a heating rate of 3 °C/min, and this temperature was maintained at 600 °C for 3 h. After cooling, the V2O3@C composite materials were prepared finally. According to the adding amount of WS-CTPs (0.1 g, 0.2 and 0.4 g), the prepared composite materials were named as V2O3@C-1, V2O3@C-2 and V2O3@C-4, respectively.

2.4 Electrochemical measurements

The electrochemical cells were prepared using the V2O3@C composite materials. Firstly, V2O3@C composite materials (0.08 g) were respectively mixed with acetylene black (0.01 g) and polyvinylidine fluoride (PVDF) binder (0.01 g) in a weight ratio of 80 : 10 : 10 in N-methyl-2-pyrrolidone (NMP) solution. The obtained slurry was coated on the Cu foil and dried in vacuum drying oven at 80 °C for 1 h to remove solution. Subsequently, the Cu foil with the active materials was dried at 120 °C for 12 h in the same vacuum drying oven and cut into round shape strips of φ 11 mm in size. The mass loading of the active materials was controlled at 1.20 mg/cm2. The two-electrode electrochemical cells (CR2032 coin-type) were assembled in a glove box filled with high-purity argon, in which cells were assembled using the lithium metal foil (φ 15.60 mm × 0.45 mm) as reference electrode, Celgard 2400 micro-porous membrane as separator, and 1 M LiPF6 in the mixture of EC, DMC, EMC (1 : 1 : 1, vol%) as electrolyte. Galvanostatic charge-discharge test was carried out by LAND (CT 2001A) battery test system in the voltage range of 0.01–3.00 V. The same electrochemical cells were also used to carry out measurements of cyclic voltammetry (CV). CV and electrochemical impedance spectroscopy (EIS) measurements were carried out using the CHI 660E. The CV curves were recorded in the voltage region of 0.01–3.00 V at a scan rate of 0.2 mV/s. The impedance spectra were recorded in a frequency range of 100 kHz–0.01 Hz. Specific capacity was calculated based on the mass of the active material V2O3@C.

3. Result and Discussions

As shown in Fig. S1, the WS-CTPs showed the more excellent solubility in water than the CTPs. To investigate a reason why the CTPs have the improved solubility, the FT-IR measurements were applied to determine the conversions of groups of materials (Fig. S2). Compared with the intensities of aromatic groups around 1573–1591 cm−1, the intensity of peak around 3298 cm−1 attributing to the –OH groups in WS-CTPs obviously increased, compared with that in coal tar pitch materials (Fig. S2). Additionally, the intensities of peaks (744 cm−1) assigned to the benzene groups in WS-CTPs remarkably decreased, compared to that in CTPs, which indicates that esterification and etherification reactions between the groups substituted on the polycyclic aromatic hydrocarbons (PAHs) probably existed in the fabrication processes of WS-CTPs.18 The peak (1180 cm−1) assigned to the -C-O groups were clearly observed in WS-CTPs, which was the effective evidence to indicate that esterification and etherification reactions existed among the PAHs in WS-CTPs.18 The formations of these groups facilely lead to the proceeding of polycondensations among the PAHs.18,19

After the same carbonization processes were performed at 600 °C for 3 h in mixture gas (N2/H2 = 9 : 1), the carbons obtained by different crude materials showed the different structures, which were verified by XRD measurements. As shown in Fig. S3, compared to coal tar pitch based carbons (CTPCs), the peak at 43.6° ascribed to the (001) disappeared in carbon materials (WS-CTPCs) obtained by carbonization of WS-CTPs crude materials, which revealed that the sizes of formed microcrystalline structure in WS-CTPCs were smaller than that in CTPCs. Meanwhile, the peak at 25.2° corresponding to the (002) peak in CTPCs became broad, and this peak shifted to the low angle (24.9°) in WS-CTPCs, indicating that layer spacing of WS-CTPCs was expanded.20,21 As a result, the structurally expending of WS-CTPCs led to the WS-CTPCs owning the more excellent storage capacity than the CTPCs.

Furthermore, the carbon contents were evaluated by thermogravimetric analysis (TGA). As presented in Fig. 1, the obvious loss of weight was observed in a temperature range of 300–400 °C, which is generally attributed to the burning of carbon.15 When increasing the temperature to 450 °C, the weight loss of V2O3@C had no distinct change. After calculations, the carbon contents in V2O3@C-1, V2O3@C-2 and V2O3@C-4 were exhibited at 10.3 %, 39.6 % and 53.5 %, respectively. It is considerable that the existences of polycondensations among the PAHs are the possible reason that causes that the relationships between the dosages and carbon contents in V2O3@C materials are not proportional.

Figure 1.

TGA results of V2O3@C materials.

Furthermore, the structures of V2O3@C were investigated by XRD in detail (Fig. 2). In comparison with the standard card (JCPDS card NO. 34-0187) of V2O3, it was observed that the characteristic peaks assigned to the 24.4°, 33.0°, 36.2°, 41.3°, 49.8° and 53.9° of V2O3 having hexagonal structures also distinctly existed in V2O3@C composite materials, implying that the characteristic structure of V2O3 did not change, even though the surface of V2O3 was covered by the carbon materials.

Figure 2.

XRD results of V2O3 and V2O3@C composite materials.

The SEM morphological features of V2O3 and V2O3@C materials were described as shown in Fig. 3. Figures 3a and 3b exhibited that there is a phenomenon that V2O3 nanocrystals strongly agglomerated. These agglomerated materials showed irregular shapes, and their sizes were 0.4–1.0 µm. With a increasing amount of carbon contents, it was found that the agglomeration phenomenon decreased remarkably (Figs. 3c and 3d), and the V2O3@C materials mainly existed as sphere-like morphologies (Figs. 3e and 3f). In addition, with continuously increasing the adding amount of carbon contents, the main morphologies of V2O3@C materials became the bulk form (Figs. 3g and 3h). These results suggested that adding carbon contents influenced the shapes of V2O3 in V2O3@C composite materials. As a consequence, it is indicative of that suitable carbon contents can improve the agglomeration phenomenon of V2O3, however, excessive amount of WS-CTPs as carbon resources is not beneficial to diminish the agglomeration of V2O3. Besides, SEM-EDS images indicate that V elements are distributed uniformly in the V2O3@C composite materials (Fig. S4).

Figure 3.

SEM images of V2O3 and V2O3@C composite materials. (a) and (b) are the images of V2O3. (c) and (d) are the images of V2O3@C-1. (e) and (f) are the images of V2O3@C-2. (g) and (h) are the images of V2O3@C-4.

The surface states of V2O3@C composite materials gave detailed descriptions in Fig. 4. First of all, the tree peaks attributing to the C, O and V elements were was observed at 284.6 eV, 530.8 eV and 517.1 eV, 524.0 eV, respectively. The C1s peak was able to be fitted to C-C (284.6 eV), C-O (286.2 eV) and O-C=O (288.6 eV), respectively, indicating that C-C, C-O and O-C=O groups exist in the V2O3@C composite materials. In accordance with a report from Li et al., the fitting of O 1s, the peaks of 530.2 eV, 531.3 eV and 533.3 eV were assigned to the V-O, C-O-V and H-O respectively.15 The existences of the C-O-V bond and V-O bond are suggestive of that conductive nets exist among V2O3 with carbon, which accelerates the enhancement of storage capacity through the improvement of conductivity.15 In the fitting peaks of V2p, the V 2p3/2 and V 2p1/2 were observed at 517.5 eV and 524.3 eV, respectively. Additionally, the peak attributing to the V4+ was observed at 515.9 eV, indicating that oxidation of V2O3 probably existed.13,15

Figure 4.

XPS spectra of the V2O3@C-2 composite material.

The structures of V2O3@C composite materials were further verified by Raman measurements. The structures of V2O3@C materials were verified by Raman measurements (Fig. 5). Similar to general carbon materials, the D peak and G peak were also observed at 1350 cm−1 and 1588 cm−1, respectively, corresponding to the sp2 hybrid covalent bonds. Generally, the ID/IG values are widely used to evaluate the structures of carbon materials. After calculations, the ID/IG values of V2O3@C-1, V2O3@C-2 and V2O3@C-4 are 1.18, 1.14 and 1.21, respectively, indicating that V2O3@C-2 possesses the relatively ordered structures. The order carbon structures of V2O3@C-2 suggest that the V2O3@C-2 possesses the comparatively excellent conductivity, compared to other V2O3@C materials.22

Figure 5.

Raman results of V2O3@C materials.

The specific surface area and porous structures were investigated in detail. As shown in Table S1, the V2O3@C-2 exhibited a bigger specific surface area than the V2O3@C-1, V2O3@C-4 and WSPCs. Associated with the analyses of TG measurements, a reasonable dosage of WS-CTPs in the reactive case is an important factor to converse the structures of carbons on the surface of V2O3, leading to the changes in specific surface areas. It is well known that a big specific surface area enables to accelerate the electrolyte infiltration and transformation of Li+ ions, resulting in a remarkable enhancement on storage capacity. Figure 6 exhibited that the V2O3@C-2 owned the more complex porous structures than other materials.

Figure 6.

Porous structures of WS-CTPCs and V2O3@C materials.

On the basis of general evaluation methods, the electrochemical performances of V2O3@C materials were evaluated comprehensively. The cycling performances were measured at 0.1 A g−1 in a potential rage of (0.05–3.0 V). As presented in Fig. 7a, the V2O3@C-2 showed the storage capacity at 580.2 mAh g−1 which was higher than the V2O3@C-1 (490.1 mAh g−1) and V2O3@C-4 (430.6 mAh g−1). By contrast, the WS-CTPCs and V2O3 respectively showed the storage capacity at 210.0 mAh g−1 and 300.2 mAh g−1, which were remarkably lower than the V2O3@C materials. These results reveal that covering carbon on the surface of V2O3 is one of the effective methods to improve the electrochemical performances of V2O3. Especially, the V2O3@C-2 showed tremendous stability at bigger current densities. For instance, when the current density was set at 5.0 A g−1, the V2O3@C-2 still showed the storage capacity at 234 mAh g−1, after 500 cycles (Fig. 7c). These aforementioned results indicate that V2O3@C-2 possesses fabulous cycle performances.

Figure 7.

Cycling performances of WS-CTPCs, V2O3 and V2O3@C materials (a). Rate performances of WS-CTPCs, V2O3 and V2O3@C materials (b). A cycling performance of V2O3@C-2 at a current density of 5 A g−1 (c).

The rate performances at different current densities (0.1 A g−1, 0.2 A g−1, 0.5 A g−1, 1.0 A g−1, 2.0 A g−1, 5.0 A g−1) were performed (Fig. 7b). Compared with the other materials, it was observed that V2O3@C-2 possessed more excellent rate performances than other materials because the V2O3@C-2 showed higher storage capacity than other materials at different current densities (Table S2). In charge-discharge measurement results, the V2O3, V2O3@C-1, V2O3@C-2 and V2O3@C-4 manifested the first coulomb efficiencies at 64.7 %, 66.6 %, 64.2 % and 69 %, respectively (Fig. S5). The low first coulomb efficiency was generally attributed to the formation of solid electrolyte interface (SEI). Compared with other materials, the relatively lower first coulomb efficiency of V2O3@C-2 is attributed to that the V2O3@C-2 possesses the bigger specific surface area than other materials (Table S1).

Electrochemical impedance spectroscopy (EIS) measurements were used to evaluate the conductive properties of V2O3, carbon and V2O3@C materials. As a result, the diameters of the semicircle of anode electrodes of V2O3@C-2 and V2O3@C-4 were much smaller than that of V2O3 and V2O3@C-1, which has been considerable that V2O3@C-2 and V2O3@C-4 electrodes possess lower charge-transfer impedances (Fig. 8a).2328 In accordance with the equivalent circuit, the R3 of V2O3, V2O3@C-1, V2O3@C-2 and V2O3@C-4 were calculated at 363.2 Ω, 332.1 Ω, 152.4 Ω and 120.2 Ω, respectively (Fig. 8c, Table S3). Although the carbons V2O3@C-2 have more order structures (Fig. 5) than V2O3@C-1 and V2O3@C-4, the fact that V2O3@C-4 possesses the more excellent conductivity is naturally attributed to the more carbon contents exist in V2O3@C composite materials (Fig. 1).

Figure 8.

Nyquist plot results (a), illustrations of relationships between Z′ and ω−1/2 in the low-frequency region (b) and the equivalent circuit model (c). Thereinto, R1 is the total resistance of the electrolyte, separator, and electrical contacts; R2 and R3 are the Li+ migration resistances through the SEI film and charge-transfer resistance, respectively; W1 is the Warburg impedance connected with the Li+ diffusion process; CPE1 and CPE2 represent the double layer resistance.

Besides, the kinetic differences between V2O3 and V2O3@C materials were further confirmed by σ value of Warburg coefficient in Fig. 8b. The σ values could be obtained by measurement of the Randles plot which is plotting of Z′ with ω−1/2 (ω = 2πf) for a low-frequency. The larger value of reflects the poor ion diffusion performance. As a result, the σ corresponding values of V2O3, V2O3@C-1, V2O3@C-2 and V2O3@C-4 were calculated at 115.2 Ω s−1/2, 92.2 Ω s−1/2, 60.8 Ω s−1/2 and 187.4 Ω s−1/2, respectively, indicating that the V2O3@C-2 possessed the remarkably higher Li+ transfer than other materials. These aforementioned results are strongly suggestive that V2O3@C-2 owned the comprehensively excellent electron/ionic conductivity than V2O3 and other V2O3@C materials.

To fully confirm the Li+ ion transfer of V2O3@C materials, the galvanostatic intermittent titration technique (GITT) calculations were performed. In particular, similar to a report from Ng et al., the concept of apparent diffusion coefficients was used in our evaluations.29 As a result, the V2O3@C-2 exhibited more tremendous Li+ ions transfer than other V2O3@C materials, no matter what they were in charge and discharge processes (Fig. 9). Associating with the BET results (Table S1, Fig. 6), the relatively large specific surface area and complex porous structures are the reasons why the V2O3@C-2 possesses a more excellent Li+ transfer property than other V2O3@C materials. Compared to the other reported V2O3/C composite materials, this V2O3@C-2 also manifests the relatively excellent storage capacity (Table S4).

Figure 9.

Relationships between voltages and apparent diffusivities during the discharge and charge process. (a) is the discharge process, and (b) is the charge process.

The electrochemical performances of V2O3, WS-CTPCs and V2O3@C materials were investigated by CV measurements (Fig. S6). The similar CV behaviors of V2O3 were also observed in our studies.3032 As shown in Fig. S6b, the oxidative peak was observed at 1.37 V, and this peak intensity diminished as increasing the cycling numbers, which indicates that the V2O3 owned the instable electrochemical property. In contrast, a similar peak of V2O3@C-2 was observed at 1.36 V, and the intensity of this peak did not change obviously, even though the cycling numbers were increased. Additionally, this oxidative potential (1.37 V) shifted to 1.58 V in V2O3@C-1 and 1.74 V in V2O3@C-4, respectively, which distinctly indicated the electrochemical property of V2O3 in V2O3@C-2 is more stable than that in V2O3@C-1 and V2O3@C-4 materials. Consequently, the CV measurement results are also suggestive of that suitable carbon content covered on the surface of V2O3 is the pivotal factor to perfectly develop the electrochemical performances of V2O3.

Finally, to understand the Li+ storage mechanism, the dynamic analyses with the CV curves at scan rates from 0.2 mV/s to 3.0 mV/s were performed systematically, similar to the reports of Dunn et al.33 According to the dynamic analysis using CV measurement results at a sweep rate of 3 mV/s, the V2O3@C-2 shows the relatively high capacitive contribution at (78.6 %) (Fig. 10). Furthermore, in spite of the fact that the sweep rate was decreased to 0.2 mV/s, the capacitive contributions still exhibit at 52.1 %. Therefore, the dynamic analysis results strongly suggest that capacitive contribution is the important factor to enhance the storage capacity of V2O3@C-2.

Figure 10.

Capacitive contributions for storage capacity of V2O3@C materials. (a) shows the CV measurement results of V2O3@C at various scan rates. (b) illustrates the bar charts showing the capacitive contributions of V2O3@C at different scan rates.

4. Conclusions

V2O3@C composite materials were successfully prepared by using the V2O5, H2C2O4·2H2O and water solving coal tar pitches (WS-CTPs). It is observed that controlling the dosages of WS-CTPs in preparation cases is a pivotal factor to improve the electrochemical performances of V2O3@C composite materials. For example, V2O3@C-2 (carbon content: 39.6 %) shows the high Li+ storage capacity at 580.2 mAh g−1, after charge-discharge was carried out 100 cycles at 0.1 A g−1. Furthermore, the V2O3@C-2.0 still shows Li+ storage capacity at 234 mAh g−1, after the charge-discharge was performed 500 cycles at 5.0 A g−1. Because of fact that the coal tar pitches are the general industrial products, the prepared V2O3@C materials should have a significant advantage on cost, leading us to consider V2O3@C composite materials prepared by our provided methods should manifest the worth expecting prospective in the usage of materials fabricating anode electrodes of LIBs.

Acknowledgments

We are grateful to the support of University of Science and Technology Liaoning (601009816-39) and 2017RC03. This work obtains the support by the Liaoning Province Education Department of China (Grant No. 601009887-16). This work is partly supported with the project supported by the National Natural Science Foundation of China (Grant No. 51672117 and 51672118).

Data Availability Statement

The data that support the findings of this study are openly available under the terms of the designated Creative Commons License in J-STAGE Data at https://doi.org/10.50892/data.electrochemistry.19972922.


CRediT Authorship Contribution Statement

Jianke Li: Formal analysis (Equal), Investigation (Equal), Writing – original draft (Equal)

Xincheng Miao: Conceptualization (Equal), Investigation (Equal)

Shenhao Wang: Data curation (Equal)

Shaobei Chen: Data curation (Equal), Investigation (Equal), Writing – original draft (Equal)

Beibei Han: Investigation (Equal), Methodology (Equal)

Guiying Xu: Conceptualization (Equal), Formal analysis (Equal)

Kun Wang: Data curation (Equal), Formal analysis (Equal), Methodology (Equal)

Baigang An: Project administration (Lead)

Dongying Ju: Resources (Equal), Supervision (Lead)

Weimin Zhou: Conceptualization (Lead), Methodology (Equal), Resources (Lead), Writing – review & editing (Lead)

Conflict of Interest

The authors declare no conflict of interest in the manuscript.

Funding

University of Science and Technology Liaoning: 601009816-39

University of Science and Technology Liaoning: 2017RC03

Department of Education of Liaoning Province: 601009887-16

National Natural Science Foundation of China: 51672117

National Natural Science Foundation of China: 51672118

Footnotes

W. Zhou: ECSJ Active Member

References
 
© The Author(s) 2022. Published by ECSJ.

This is an open access article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse of the work in any medium provided the original work is properly cited. [DOI: 10.5796/electrochemistry.22-00032].
http://creativecommons.org/licenses/by/4.0/
feedback
Top