Next Article in Journal
Influence of Upstream Disturbances on the Vortex Structure of Francis Turbine Based on the Criteria of Identification of Various Vortexes
Next Article in Special Issue
Effect of Synthesis Conditions on Capacitive Properties of Porous Carbon Derived from Hemp Bast Fiber
Previous Article in Journal
Fundamental Responsiveness in European Electricity Prices
Previous Article in Special Issue
Digitally Patterned Mesoporous Carbon Nanostructures of Colorless Polyimide for Transparent and Flexible Micro-Supercapacitor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of the Hydrophilic Surface of Nanofiber Support on the Performance of Hybrid Supercapacitors

1
Department of Chemical Engineering, College of Engineering, Wonkwang University, Iksan 54538, Korea
2
Department of Automotive science, Facility of Integrated Frontier Sciences, Kyushu University, Fukuoka 813-8503, Japan
3
Nanoscale Sciences and Technology Institute, Wonkwang University, Iksan 54538, Korea
4
Department of Mechanical Systems Engineering, Sookmyung Women’s University, Seoul 04310, Korea
5
Institute of Advanced Materials and Systems, Sookmyung Women’s University, Seoul 04310, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Energies 2021, 14(22), 7621; https://doi.org/10.3390/en14227621
Submission received: 18 October 2021 / Revised: 3 November 2021 / Accepted: 12 November 2021 / Published: 15 November 2021
(This article belongs to the Special Issue Novel Materials and Technologies for Supercapacitor Applications)

Abstract

:
Concerns associated with global warming and the depleting reserves of fossil fuels have highlighted the importance of high−performance energy storage systems (ESSs) for efficient energy usage. ESSs such as supercapacitors can contribute to improved power quality of an energy generation system, which is characterized by a slow load response. Composite materials are primarily used as supercapacitor electrodes because they can compensate for the disadvantages of carbon or metal oxide electrode materials. In this study, a composite of oxide nanoparticles loaded on a carbon nanofiber support was used as an electrode material for a hybrid supercapacitor. The addition of a small amount of hydrophilic FeN@GnP (Fe− and N−doped graphene nanoplates) modified the surface properties of carbon nanofibers prepared by electrospinning. Accordingly, the effects of the hydrophobic/hydrophilic surface properties of the nanofiber support on the morphology of Co3O4 nanoparticles loaded on the nanofiber, as well as the performance of the supercapacitor, were systematically investigated.

1. Introduction

Recently, renewable energy forms such as solar energy and wind power have been attracting much interest due to fossil fuel depletion and environmental pollution concerns associated with global warming and climate change. With the increasing use of renewable energy, it is necessary to develop an energy storage system (ESS) to efficiently use the produced electrical energy. Recently, secondary batteries, such as Li−ion batteries, have been primarily used as ESSs with renewable energy. Supercapacitors, as electrical capacitors, are attractive ESSs that offer the advantages of a high power density and stable cycle life owing to their charge−storage mechanism, compared to secondary Li−ion batteries and conventional capacitors [1,2,3,4]. In addition, Arunachalam et al. reported that the need for research on devices to store clean energy sources that are considered alternative energy sources. Moreover, supercapacitors are attracting attention as important energy storage and other applications; hence, it is necessary to develop a composite electrode material of carbon material and metal oxide, which controls morphology and size by various synthesis methods for improved energy storage [5].
Supercapacitors [6,7,8,9] are classified into three types according to the mechanism of energy storage: electrical double−layer capacitors (EDLCs), pseudocapacitors (PCs), and hybrid supercapacitors (HCs). EDLCs store energy by charge separation at the electrode–electrolyte interface, while PCs store energy through the Faradaic process. [10] HCs can provide high pseudocapacitance, in addition to the high capacitance provided by the electrical double layer at the supporting electrode surface [11]. EDLCs employ various carbon materials such as activated carbon [12], carbon nanotubes [13], carbon nanofibers (CNFs) [14], and graphene [15] as electrodes because their large specific surface area provides a high capacity, and their electrochemical stability provides a long service life. However, EDLCs employing carbon materials present the disadvantage of a low energy density. PCs use metal oxides such as RuO2 [16], NiO [17,18], Co3O4 [19,20], MnO2 [21,22], and conducting polymers [23] as electrode materials. Among the electrode materials, Co3O4 has been considered promising because of its high theoretical specific capacitance (Cs = 3560 F/g) [24]. However, Co3O4 exhibits a low charge/discharge rate, low capacity due to its low electroconductivity, and variations in the crystalline structure during the redox process [25].
To overcome the drawbacks, studies have focused on hybrid supercapacitors using various metal oxide−loaded materials. [26,27,28]. Metal oxides were selected as the loading materials owing to their high electroconductivity compared to conducting polymers. Among the methods for fabricating metal oxide−loaded carbon materials, hydrothermal synthesis enables the synthesis of metal oxides at a low temperature and has been used in various applications [27,28,29,30,31]. Metal oxides prepared on carbon materials may be coated entirely on the carbon surface [29] or exist in the form of agglomerates on the carbon surface [30]. Hydrothermal synthesis generates crystals by raising the temperature and pressure of a homogeneous aqueous solution or precursor suspension. [31] To obtain a high specific capacitance through the addition of metal oxides, it is necessary to evenly distribute and grow metal oxides on the surface of carbon materials. The surface properties of the carbon material in the precursor suspension contribute crucially to the morphology of the metal oxide particles in the electrode. Furthermore, the surface characteristics of carbon materials can control the loading conditions of the metal oxide particles and thus affect the performance of supercapacitors (e.g., if the particles are well−loaded).
Therefore, in this study, the relationship between the growth of metal oxide nanoparticles and the hydrophilic/hydrophobic characteristics of a CNF support was investigated.

2. Materials and Methods

A CNF support was prepared by electrospinning using a 10 wt.% polyacrylonitrile (PAN, MW ≈ 150,000 g·mol−1, Sigma−Aldrich Co., St. Louis, MO, USA) solution. The prepared PAN solution was filled in a 10 mL syringe with a capillary tip (d = 0.5 mm). Electrospinning was performed with a variable high−voltage power supply (SHV30R, ConverTech. Co., Gyeongbuk, Korea); the anode of the power supply was clamped to a syringe needle tip, while the cathode was connected to an aluminum foil collector (diameter = 90 mm). The applied voltage was 20 kV, the distance between the nozzle and collector was 18 cm, and the supply rate of the solution was 0.8 mL/h. The collected electrospun fibers were stabilized at 523 K for 1 h, carbonized at 1073 K for 2 h in N2, and then activated at 1073 K for 2 h in CO2. CNFs with Fe− and N−doped graphene nanoplates (CNF−FeN@GnP) were also prepared by electrospinning. For the Fe− and N−doped graphene nanoplates (FeN@GnP)−dispersed solution, the 0.009 g FeN@GnPs (Graphene Edge, Inc. Seoul, Korea) and 9 g DMF mixture was sonicated by a homogenizer for 30 min to completely disperse them. Then 1 g PAN powder was solved into the mixture solution. Nanofibers prepared using 10 wt.% PAN solution with 1 g of PAN became carbon fibers of 0.18 g after stabilization, carbonization, and activation. In our previous work, the carbon fiber−containing 5 wt.% FeN@GnP exhibited the highest active surface area [32] Therefore, in this work, the same ratio of FeN@GnPs was used to prepare CNF−FeN@GnP. The experimental protocol thereafter was the same as that used for preparing the CNFs.
Co3O4 nanoparticles were loaded by hydrothermal synthesis using the prepared nanofiber support. The hydrothermal reaction solution was prepared by adding cobalt(II) nitrate hexahydrate (Co(NO3)2·6H2O, 97%, Samchun Pure Chemical Co., Ltd., Pyeongtaek, Korea) and urea (NH2CONH2, 98%, Sigma Aldrich Co.) to 80 mL of distilled water. The precursors and urea were mixed in a molar ratio of 3:0.5, followed by continuous stirring for 10 min. When the solute was completely dissolved, 0.1 g of the prepared nanofibers was added and stirred for 10 min. The solution was transferred to a Teflon−lined autoclave and heated at 393 K for 12 h, following which the autoclave was cooled to room temperature. Co3O4/CNF and Co3O4/CNF−FeN@GnP were obtained after repeated washing with distilled water and drying at 353 K in a vacuum oven.
Morphological observations and analysis of the elemental compositions of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP were performed by field−emission scanning electron microscopy (FE−SEM, S−4800, Hitachi, Ibarakiken, Japan). Structural characterization of the prepared sample was carried out via X−ray diffraction (XRD) using a MiniFlex 600 (Rigaku, Tokyo, Japan) diffractometer with Cu−Kα radiation (λ = 1.5406 Å) in the 2θ range of 20°–80° at 4°/min. The specific surface areas and pore size distributions were determined by applying the Brunauer−Emmett−Teller (BET) equation to the isotherms recorded at 77 K (BELSORP−mini II, Mictrotrac BEL, Osaka, Japan). The weight change and thermal stability of the prepared samples were estimated by thermogravimetric analysis (TGA, TGA8000, PerkinElmer, Waltham, MA, USA). Chemical structures, such as functional groups, were examined using Fourier−transform infrared (FT−IR) spectroscopy (Spectrum Two, PerkinElmer, USA). The hydrophobic/hydrophilic properties were identified by contact angle measurements (Smart Drop, Femtobiomed, Gyeonggi-do, Korea).
Electrochemical measurements were performed using a battery testing system (BCS−815, Bio−logic, Gottingen, Germany). The area of each electrode was approximately 1 cm2, and a 6 M KOH aqueous solution was used as the electrolyte. The electrode material was fabricated by mixing 80 wt.% of the active material, 10 wt.% of the binder (PVDF), and 10 wt.% of Super−P as a conductive carbon material, using N−methyl−2−pyrrolidone as the solvent, to yield a slurry. The slurry was pressed onto nickel foil as a current collector and dried at 373 K for 24 h. The GCD−specific capacitance was calculated using the following formula:
C= (I•∆t)/(m•∆V)
In the above formula [33,34,35], C (F•g−1) denotes the specific capacitance, ∆t (s) represents the discharge time, I (mA) is the discharge current, ∆V (V) refers to the voltage, and m (mg) indicates the mass of the electrode material.

3. Results and Discussion

Figure 1a,d show SEM images of the electrospun fibers after calcination at 1073 K in CO2. The surface morphologies of the CNF and CNF−FeN@GnP nanofibers were rough due to the partial burn−off caused by the reaction between carbon and CO2 during activation [36]. The surface of the CNF−FeN@GnP nanofiber was rougher than that of the CNFs because the FeN@GnP particles that were distributed on the nanofiber surface allowed less heat shrinkage during heat treatment [32]. The average diameters of the CNF and CNF−FeN@GnP nanofibers were approximately 333 nm and 350 nm, respectively. As an electrode material, a one−dimensional nanofiber is perceived to facilitate short pathways for electron transport in electrochemical reactions [37]. After hydrothermal synthesis in the cobalt solution, the morphology was maintained in the form of a fibrous framework, as shown in Figure 1b,e, which confirmed that cobalt oxide species nanoparticles were loaded on the nanofiber support. Loading cobalt oxide species such as Co3O4 can be effective in enhancing the performance of the supercapacitor because metal oxides are more redox−active than CNFs [38]. In addition, unlike in the CNF/FeN@GnP nanofiber support (Figure 1f) agglomeration of the Co3O4 nanoparticles on the CNF support (Figure 1c) was detected after hydrothermal synthesis. To confirm the reason for this, the hydrophobic/hydrophilic properties of the nanofiber support were analyzed by contact angle measurements, the results of which are shown in Figure 2. To further investigate the elemental composition of the nanoparticles, Energy−dispersive X−ray spectroscopy (EDX) was conducted on Cobalt oxide/CNF and Cobalt oxide /CNF−FeN@GnP, and the micrographs are shown in Supplementary Material Figure S1a,b, respectively. The nanoparticles were determined to contain Co and O, confirming that cobalt oxide species were loaded on the nanofiber support.
The contact angle results for CNF and the CNF/FeN@GnP nanofiber support powders are shown in Figure 2a,b. To determine whether the surface was hydrophobic (contact angle higher than 90°) or hydrophilic (contact angle lower than 90°), contact angle measurements were conducted using water. The CNF support showed hydrophobic properties with a contact angle of 143.4°, while CNF/FeN@GnP showed hydrophilic properties with a contact angle of 41.7°. The contact angle of FeN@GnP powder was also measured to determine the hydrophilicity of CNF/FeN@GnP (Figure 2c). As shown in Figure 2d, FeN@GnP powder had hydrophilic properties with a contact angle of 11.3°. Ju et al. reported that when carbon composite fibers were prepared by electrospinning a polymer solution containing graphene nanoplates, the nanoplates were distributed on the surface of the fibers by the centrifugal force generated during electrospinning [37]. Hence, it is expected that the properties of the CNF surface will be affected by the added FeN@GnP powder. Cheng et al. and Lu et al. reported that hydrophobic surface properties induce high agglomeration due to the strong hydrophobic interaction between particles [39,40]. Therefore, compared to the hydrophilic FeN@GnP nanofiber support, the hydrophobic CNF support can enhance the aggregation of Co3O4 nanoparticles loaded on the surface of the nanofiber support, consistent with the SEM results.
The crystallinity of CNF, CNF/FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP was analyzed by XRD, and the results are shown in Figure 3. The presence of the (002) peak in the 2θ range of 20°–30° in the XRD patterns of Figure 3a,b indicates the existence of amorphous carbon. After hydrothermal synthesis, the intensity of the peak corresponding to amorphous carbon decreased owing to the relatively high crystallinity of Co3O4 loaded on the nanofiber support. The results confirmed that the prepared Co3O4/CNF and Co3O4/CNF−FeN@GnP were well−developed into Co3O4 (JCPDS No. 42−1467) crystalline structures, as shown in Figure 3c,d, corresponding to Co3O4 phases with the (111), (220), (311), (222), (400), (422), (511), and (440) planes. In addition, Co3O4 did not exhibit any additional peak corresponding to any impurity or secondary phase. The presence of a Co3O4 peak in the XRD pattern confirmed that the nanoparticles loaded on the nanofiber support were those of Co3O4, as also corroborated by the SEM−EDX results. These results prove that Co3O4/CNF and Co3O4/CNF−FeN@GnP were successfully prepared after hydrothermal synthesis [24].
To further identify the hydrophobic/hydrophilic surface properties of CNF and the CNF−FeN@GnP nanofiber supports, the functional groups were determined by FT−IR spectroscopy. Figure 4 shows that CNF and CNF/FeN@GnP possessed functional groups such as −OH (3400 cm−1), C−H (2900 cm−1), C=C (1600 cm−1), and −CH3 (1380 cm−1). The FT−IR peak intensities of CNF and CNF/FeN@GnP were difficult to confirm in the FT−IR spectra because a small amount (0.009 g) of FeN@GnP powder was added to the CNFs. However, the peak intensities of the −CH and −CH3 groups of the CNFs were slightly higher than those of CNF/FeN@GnP. The integral values of the peak(−CH3)/(C=C) for CNFs and CNF/FeN@GnP were 0.044 and 0.042, respectively. Zhu et al. reported that −CH [41] and −CH3 groups [42] in the FT−IR spectra induced hydrophobic properties. Therefore, these results confirmed that the CNF support has hydrophobic properties, making the FT−IR results consistent with the contact angle measurements.
Figure 5 shows the nitrogen adsorption/desorption isotherms of CNF, CNF/FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP. The isotherms were typical type I, indicating the existence of well−developed micropores in the samples, as shown in Figure 5a. The specific surface areas of CNF, CNF/FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP were 598.25, 509.38, 342.98, and 391.88 m2/g, respectively. The specific surface areas of CNF and the CNF/FeN@GnP nanofibers were similar. CNF/FeN@GnP had a slightly lower specific surface area than the CNFs because FeN@GnP was partially embedded in the CNFs and provided a slightly less rough nanofiber surface of FeN@GnP. In addition, the pore volumes of CNF, CNF/FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP were 0.2757, 0.2499, 0.1883, and 0.1966 cm3/g, respectively, as shown in Figure 5b. The smaller pore volume of CNF/FeN@GnP compared to that of the CNFs can additionally confirm the effect of FeN@GnP powder on the specific surface area. In addition, the specific surface area decreased after loading Co3O4 nanoparticles onto CNF and the CNF/FeN@GnP nanofiber supports because the loaded nanoparticles induced a decrease in the pore volume of the nanofiber. These results were confirmed by a decrease in the pore volume intensity, as shown in Figure 5b.
Figure 6 shows the TGA curves of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP in air. Overall, the nanofibers show a weight−loss stage of approximately 10 % up to 273–373 K. This weight loss corresponded to the evaporation of water and the solvent. Significant weight loss occurred between 723 and 873 K. The second weight loss was attributed to the low thermal stability of the CNFs. The loaded content of Co3O4 was predicted from the residual content after burning out the weight lost by PAN and the FeN@GnP−embedded PAN fiber composite in the air. The residual contents of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP were 0.5, 1.5, 12.5, and 12.8, respectively. The higher weights of Co3O4/CNF and Co3O4/CNF−FeN@GnP compared to those of CNF and CNF−FeN@GnP in the TGA curve indicated that similar Co3O4 nanoparticles were well−loaded on CNFs and CNF/FeN@GnP. In comparison to CNFs, the CNF−FeN@GnP nanofibers were confirmed to have undergone a rapid weight loss. In addition, after loading Co3O4 nanoparticles on the nanofiber support, weight loss occurred faster at a lower temperature than that in the pristine nanofiber supports, as shown in the TGA curves (Figure 6). It is assumed that the interaction of the surface of the nanofibers may have been affected by the loaded Co3O4 nanoparticles or Fe in FeN@GnP. Oh et al. reported that the carbon around metal particles is easily decomposed by the chemical catalytic effect of the metal particles at high temperatures [43]. The faster weight loss originated from the chemical decomposition of carbon during burn−off by Fe in FeN@GnP or loaded Co3O4.
Figure 7a shows typical Nyquist plots of the CNF and CNF−FeN@GnP samples measured under open−circuit conditions. The polarization resistance of CNF/FeN@GnP was evaluated to be approximately 62.9 Ω, which is much lower than that of the CNFs (73.9 Ω). A lower polarization resistance leads to improved conductivity [44] and faster charge transport owing to the redox properties of FeN@GnP. Figure 7b shows the Nyquist plots of Co3O4/CNF (1.49 Ω) and Co3O4/CNF−FeN@GnP (0.2 Ω). Co3O4 loading on each nanofiber support resulted in a lower polarization resistance compared to those of other nanofiber supports. Co3O4 loading increased the electroconductivity through the redox reaction of metal oxides and can improve the performance of the supercapacitor [45]. At the lower frequency, the linear line was related to the Warburg diffusion resistance. Most of the samples had a vertical line tilt at an angle of 45 degree, and Co3O4/CNF−FeN@GnP samples with a slope close to vertical showed that the ions diffused rapidly [46]. In the GCD curve (Figure 7c), Co3O4/CNF−FeN@GnP showed an excellent specific capacitance of 668 F/g compared to those of the other samples (CNF, CNF−FeN@GnP, Co3O4/CNF had capacitances of 71, 102, and 291.1 F/g, respectively). The Co3O4/CNF and Co3O4/CNF−FeN@GnP nanofibers showed different results despite being loaded with Co3O4. This was because the nanoparticles were more evenly dispersed and grown on the surface of the Co3O4/CNF−FeN@GnP nanofibers due to the hydrophilic properties of the CNF/FeN@GnP nanofiber support. The loading amount of Co3O4 particles was confirmed to be similar in TGA, but it was obvious that the dispersion and agglomeration of Co3O4 particles affected the difference in the performance. The GCD curve of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP electrode measured in the potential range of 0 to 0.9 V at various current densities is shown in Supplementary Material Figure S2a–d. It can be seen that the specific capacitance of the electrode was decreased with an increase in current density. At low current density, the specific capacitance was high, which might be due to a slow Faradic, as well as non−Faradic reaction. At a high current density, the diffusion of ions occurred mainly implying a fast electrochemical reaction.
Figure 7d shows the cyclic voltammograms (CV) of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP electrodes with a 20 mV/s sweep rate. The CV tests were carried out over the voltage range from 0 to 0.9 V in a 6 M KOH aqueous solution. The voltammogram of the electrode made from CNF and CNF−FeN@GnP without Co3O4 was a rectangular type. It indicated that CNF and CNF−FeN@GnP electrodes had ideal electric double−layer capacitance. On the other hand, the Co3O4/CNF and Co3O4/CNF−FeN@GnP electrodes exhibited the current peaks from oxidation and reduction attributed to the changed valence state of cobalt. This means that the charge storage mechanism of Co3O4 particles coated samples was not only electric double−layer capacitive but also pseudocapacitive owing to Faradaic processes [47,48].
In order to evaluate the cycle stability of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP electrodes, a continuous charge−discharge process was conducted at a current density of 1 mA/cm2 for 700 cycles, as shown in Figure 7e. The Co3O4/CNF−FeN@GnP electrode showed 87.5 % capacitance retention after 700 cycles compared to other electrode materials (capacitance retention of CNF, CNF−FeN@GnP, and Co3O4/CNF was 86 %, 75.5 %, and 75.6 %, respectively.) The superior cycle stability was an important factor for high−performance supercapacitors. These results demonstrated that the Co3O4/CNF−FeN@GnP using the hydrophilic support improved the cycle stability due to its redox properties of well−dispersed cobalt oxide and superior cycle stability by the carbon−based support materials. Therefore, this study reveals the dependence of the dispersibility of the loaded nanoparticles on the hydrophobic/hydrophilic surface properties of nanofibers for advanced supercapacitor materials.

4. Conclusions

In this study, a composite electrode material was prepared using a two−step process. Two types of nanofiber samples, CNFs and CNF−FeN@GnP nanofibers, were prepared by electrospinning, following which Co3O4 nanoparticles were loaded onto the nanofiber support via hydrothermal synthesis. Depending on the hydrophilicity and hydrophobicity of the two types of nanofibers, there was a difference in the dispersibility when loading the nanoparticles onto the nanofibers. CNF−FeN@GnP with hydrophilic surface properties enabled the even loading of Co3O4 nanoparticles on the surface of the nanofiber support compared to CNFs. The prepared Co3O4/CNF−FeN@GnP nanofiber exhibited an excellent capacitance of 668 F/g owing to the high redox characteristics of evenly dispersed Co3O4 nanoparticles. Thus, the results proved that the dispersibility of the loaded nanoparticles and the dependence on the hydrophobic/hydrophilic surface properties of the nanofibers are crucial for determining the usefulness of the prepared materials as electrodes.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/en14227621/s1, Figure S1: SEM−EDX analysis of (a) Co3O4/CNF (b) Co3O4/CNF−FeN@GnP, Figure S2: GCD curves of (a) CNF, (b) CNF−FeN@GnP, (c) Co3O4/CNF, and (d) Co3O4/CNF−FeN@GnP at different current density (1, 2, 3, 4, 5, 7, 10 mA/cm2)”.

Author Contributions

Conceptualization, J.S. and Y.-W.J.; investigation, H.-Y.K. and S.-Y.L.; data curation, H.-Y.K. and S.-Y.L.; writing—original draft preparation, H.-Y.K.; writing—review and editing, J.S. and Y.-W.J.; visualization, H.-Y.K.; supervision, I.-Y.J., J.S. and Y.-W.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Mid−career Researcher Program (NRF− 2020R1A2C1007847) and the Basic Research Laboratory Program (No. NRF−2020R1A4A3079200) through the National Research Foundation of Korea, funded by the Korea government Ministry of Education.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Conway, B.E. Electrochemical Supercapacitors; Plenum Publishing Co.: New York, NY, USA, 1999; pp. 1–698. [Google Scholar]
  2. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. In Nanoscience and Technology: A Collection of Reviews from Nature Journals; Nature: London, UK, 2010; pp. 320–329. [Google Scholar]
  3. Yan, J.; Wang, Q.; Wei, T.; Fan, Z. Recent advances in design and fabrication of electrochemical supercapacitors with high energy densities. Adv. Energy Mater. 2014, 4, 1300816. [Google Scholar] [CrossRef]
  4. Shao, Y.; El−Kady, M.F.; Wang, L.J.; Zhang, Q.; Li, Y.; Wang, H.; Mousavi, M.F.; Kaner, R.B. Graphene−based materials for flexible supercapacitors. Chem. Soc. Rev. 2015, 44, 3639–3665. [Google Scholar] [CrossRef]
  5. Arunachalam, S.; Kirubasankar, B.; Pan, D.; Liu, H.; Yan, C.; Guo, Z.; Angaiah, S. Research progress in rare earths and their composites based electrode materials for supercapacitors. Green Energy Environ. 2020, 5, 259–273. [Google Scholar] [CrossRef]
  6. Arunachalam, S.; Kirubasankar, B.; Murugadoss, V.; Vellasamy, D.; Angaiah, S. Facile synthesis of electrostatically anchored Nd (OH)3 nanorods onto graphene nanosheets as a high capacitance electrode material for supercapacitors. New J. Chem. 2018, 42, 2923–2932. [Google Scholar] [CrossRef]
  7. Arunachalam, S.; Kirubasankar, B.; Rajagounder Nagarajan, E.; Vellasamy, D.; Angaiah, S. A facile chemical precipitation method for the synthesis of Nd (OH)3 and La (OH)3 nanopowders and their supercapacitor performances. Chem. Sel. 2018, 3, 12719–12724. [Google Scholar] [CrossRef]
  8. Balakrishnan, K.; Kumar, M.; Angaiah, S. Synthesis of Polythiophene and Its Carbonaceous Nanofibers as Electrode Materials for Asymmetric Supercapacitors. In Advanced Materials Research; Scientific: Basel, Switzerland, 2014; pp. 151–157. [Google Scholar]
  9. Kirubasankar, B.; Balan, B.; Yan, C.; Angaiah, S. Recent Progress in Graphene-Based Microsupercapacitors. Energy Technol. 2021, 9, 2000844. [Google Scholar] [CrossRef]
  10. Halper, M.S.; Ellenbogen, J.C. Supercapacitors: A Brief Overview; MITRE Corp.: Bedford, MA, USA, 2006; Volume 1, p. 34. [Google Scholar]
  11. Zhao, X.; Johnston, C.; Grant, P.S. A novel hybrid supercapacitor with a carbon nanotube cathode and an iron oxide/carbon nanotube composite anode. J. Mater. Chem. 2009, 19, 8755–8760. [Google Scholar] [CrossRef]
  12. Ishimoto, S.; Asakawa, Y.; Shinya, M.; Naoi, K. Degradation responses of activated−carbon−based EDLCs for higher voltage operation and their factors. J. Electrochem. Soc. 2009, 156, A563. [Google Scholar] [CrossRef]
  13. Signorelli, R.; Ku, D.C.; Kassakian, J.G.; Schindall, J.E. Electrochemical double−layer capacitors using carbon nanotube electrode structures. Proc. IEEE 2009, 97, 1837–1847. [Google Scholar] [CrossRef]
  14. Kim, B.-H.; Yang, K.S.; Ferraris, J.P. Highly conductive, mesoporous carbon nanofiber web as electrode material for high−performance supercapacitors. Electrochim. Acta 2012, 75, 325–331. [Google Scholar] [CrossRef]
  15. Miller, J.R.; Outlaw, R.; Holloway, B. Graphene electric double layer capacitor with ultra−high−power performance. Electrochim. Acta 2011, 56, 10443–10449. [Google Scholar] [CrossRef]
  16. Kakaei, K.; Esrafili, M.D.; Ehsani, A. Alcohol oxidation and hydrogen evolution. Interface Sci. Technol. 2019, 27, 253–301. [Google Scholar]
  17. Lee, J.Y.; Liang, K.; An, K.H.; Lee, Y.H. Nickel oxide/carbon nanotubes nanocomposite for electrochemical capacitance. Synth. Met. 2005, 150, 153–157. [Google Scholar] [CrossRef]
  18. Yuan, C.; Zhang, X.; Wu, Q.; Gao, B. Effect of temperature on the hybrid supercapacitor based on NiO and activated carbon with alkaline polymer gel electrolyte. Solid State Ion. 2006, 177, 1237–1242. [Google Scholar] [CrossRef]
  19. Meher, S.K.; Rao, G.R. Ultralayered Co3O4 for high−performance supercapacitor applications. J. Phys. Chem. C 2011, 115, 15646–15654. [Google Scholar] [CrossRef]
  20. Xia, X.-H.; Tu, J.-P.; Mai, Y.-J.; Wang, X.-L.; Gu, C.-D.; Zhao, X.-B. Self−supported hydrothermal synthesized hollow Co3O4 nanowire arrays with high supercapacitor capacitance. J. Mater. Chem. 2011, 21, 9319–9325. [Google Scholar] [CrossRef]
  21. Chang, J.-K.; Chen, Y.-L.; Tsai, W.-T. Effect of heat treatment on material characteristics and pseudo−capacitive properties of manganese oxide prepared by anodic deposition. J. Power Sources 2004, 135, 344–353. [Google Scholar] [CrossRef]
  22. Prasad, K.R.; Miura, N. Potentiodynamically deposited nanostructured manganese dioxide as electrode material for electrochemical redox supercapacitors. J. Power Sources 2004, 135, 354–360. [Google Scholar] [CrossRef]
  23. Hou, Y.; Chen, L.; Zhang, L.; Kang, J.; Fujita, T.; Jiang, J.; Chen, M. Ultrahigh capacitance of nanoporous metal enhanced conductive polymer pseudocapacitors. J. Power Sources 2013, 225, 304–310. [Google Scholar] [CrossRef]
  24. Yang, S.; Liu, Y.; Hao, Y.; Yang, X.; Goddard, W.A., III; Zhang, X.L.; Cao, B. Oxygen-vacancy abundant ultrafine Co3O4/graphene composites for high-rate supercapacitor electrodes. Adv. Sci. 2018, 5, 1700659. [Google Scholar] [CrossRef] [Green Version]
  25. Zhang, Y.; Hu, Y.; Wang, Z.; Lin, T.; Zhu, X.; Luo, B.; Hu, H.; Xing, W.; Yan, Z.; Wang, L. Lithiation-Induced Vacancy Engineering of Co3O4 with Improved Faradic Reactivity for High-Performance Supercapacitor. Adv. Funct. Mater. 2020, 30, 2004172. [Google Scholar] [CrossRef]
  26. Rakhi, R.; Alshareef, H.N. Enhancement of the energy storage properties of supercapacitors using graphene nanosheets dispersed with metal oxide−loaded carbon nanotubes. J. Power Sources 2011, 196, 8858–8865. [Google Scholar] [CrossRef]
  27. Zawadzki, M.; Wrzyszcz, J. Hydrothermal synthesis of nanoporous zinc aluminate with high surface area. Mater. Res. Bull. 2000, 35, 109–114. [Google Scholar] [CrossRef]
  28. Stein, A.; Keller, S.W.; Mallouk, T.E. Turning down the heat: Design and mechanism in solid−state synthesis. Science 1993, 259, 1558–1564. [Google Scholar] [CrossRef]
  29. Lu, F.; Wang, J.; Sun, X.; Chang, Z. 3D hierarchical carbon nanofibers/TiO2@ MoS2 core−shell heterostructures by electrospinning, hydrothermal and in−situ growth for flexible electrode materials. Mater. Des. 2020, 189, 108503. [Google Scholar] [CrossRef]
  30. Mao, K.; Wu, X.; Min, X.; Huang, Z.; Liu, Y.-G.; Fang, M. New efficient visible−light−driven photocatalyst of chitin−modified titanium dioxide/carbon fiber composites for wastewater. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef]
  31. Laudise, R.; Nielsen, J. Hydrothermal crystal growth. In Solid State Physics; Elsevier: Amsterdam, The Netherlands, 1961; Volume 12, pp. 149–222. [Google Scholar]
  32. Ju, Y.W.; Yoo, S.; Kim, C.; Kim, S.; Jeon, I.Y.; Shin, J.; Baek, J.B.; Kim, G. Fe@ N-Graphene Nanoplatelet-Embedded Carbon Nanofibers as Efficient Electrocatalysts for Oxygen Reduction Reaction. Adv. Sci. 2016, 3, 1500205. [Google Scholar] [CrossRef]
  33. Kirubasankar, B.; Narayanasamy, M.; Yang, J.; Han, M.; Zhu, W.; Su, Y.; Angaiah, S.; Yan, C. Construction of heterogeneous 2D layered MoS2/MXene nanohybrid anode material via interstratification process and its synergetic effect for asymmetric supercapacitors. Appl. Surf. Sci. 2020, 534, 147644. [Google Scholar] [CrossRef]
  34. Kirubasankar, B.; Palanisamy, P.; Arunachalam, S.; Murugadoss, V.; Angaiah, S. 2D MoSe2−Ni (OH)2 nanohybrid as an efficient electrode material with high rate capability for asymmetric supercapacitor applications. Chem. Eng. J. 2019, 355, 881–890. [Google Scholar] [CrossRef]
  35. Narayanasamy, M.; Kirubasankar, B.; Shi, M.; Velayutham, S.; Wang, B.; Angaiah, S.; Yan, C. Morphology restrained growth of V2O5 by the oxidation of V−MXenes as a fast diffusion controlled cathode material for aqueous zinc ion batteries. Chem. Commun. 2020, 56, 6412–6415. [Google Scholar] [CrossRef]
  36. Zhou, X.; Yi, H.; Tang, X.; Deng, H.; Liu, H. Thermodynamics for the adsorption of SO2, NO and CO2 from flue gas on activatecarbon fiber. Chem. Eng. J. 2012, 200, 399–404. [Google Scholar] [CrossRef]
  37. Ouyang, T.; Cheng, K.; Yang, F.; Zhou, L.; Zhu, K.; Ye, K.; Wang, G.; Cao, D. From biomass with irregular structures to 1D carbon nanobelts: A stripping and cutting strategy to fabricate high performance supercapacitor materials. J. Mater. Chem. A 2017, 5, 14551–14561. [Google Scholar] [CrossRef]
  38. Yang, P.; Chen, Y.; Yu, X.; Qiang, P.; Wang, K.; Cai, X.; Tan, S.; Liu, P.; Song, J.; Mai, W. Reciprocal alternate deposition strategy using metal oxide/carbon nanotube for positive and negative electrodes of high−performance supercapacitors. Nano Energy 2014, 10, 108–116. [Google Scholar] [CrossRef]
  39. Cheng, W.; Deng, Z.; Tong, X.; Lu, T. Hydrophobic agglomeration of fine pyrite particles induced by flotation reagents. Minerals 2020, 10, 801. [Google Scholar] [CrossRef]
  40. Lu, S.; Song, S. Hydrophobic interaction in flocculation and flotation 1. Hydrophobic flocculation of fine mineral particles in aqueous solution. Colloids Surf. 1991, 57, 49–60. [Google Scholar] [CrossRef]
  41. Zhu, Y.; Otsubo, M.; Honda, C. Degradation of polymeric materials exposed to corona discharges. Polym. Test. 2006, 25, 313–317. [Google Scholar] [CrossRef]
  42. Zhu, Y.; Otsubo, M.; Honda, C.; Tanaka, S. Loss and recovery in hydrophobicity of silicone rubber exposed to corona discharge. Polym. Degrad. Stab. 2006, 91, 1448–1454. [Google Scholar] [CrossRef]
  43. Oh, G.-Y.; Ju, Y.-W.; Jung, H.-R.; Lee, W.-J. Preparation of the novel manganese−embedded PAN−based activated carbon nanofibers by electrospinning and their toluene adsorption. J. Anal. Appl. Pyrolysis 2008, 81, 211–217. [Google Scholar] [CrossRef]
  44. Jiang, R.; Huang, T.; Tang, Y.; Liu, J.; Xue, L.; Zhuang, J.; Yu, A. Factors influencing MnO2/multi−walled carbon nanotubes composite’s electrochemical performance as supercapacitor electrode. Electrochim. Acta 2009, 54, 7173–7179. [Google Scholar] [CrossRef]
  45. Xiang, K.; Xu, Z.; Qu, T.; Tian, Z.; Zhang, Y.; Wang, Y.; Xie, M.; Guo, X.; Ding, W.; Guo, X. Two dimensional oxygen−vacancy−rich Co3O4 nanosheets with excellent supercapacitor performances. Chem. Commun. 2017, 53, 12410–12413. [Google Scholar] [CrossRef] [PubMed]
  46. Subasri, A.; Balakrishnan, K.; Nagarajan, E.; Devadoss, V.; Subramania, A. Development of 2D La (OH)3/graphene nanohybrid by a facile solvothermal reduction process for high−performance supercapacitors. Electrochim. Acta 2018, 281, 329–337. [Google Scholar] [CrossRef]
  47. Kirubasankar, B.; Murugadoss, V.; Lin, J.; Ding, T.; Dong, M.; Liu, H.; Zhang, J.; Li, T.; Wang, N.; Guo, Z. In situ grown nickel selenide on graphene nanohybrid electrodes for high energy density asymmetric supercapacitors. Nanoscale 2018, 10, 20414–20425. [Google Scholar] [CrossRef] [PubMed]
  48. Kirubasankar, B.; Vijayan, S.; Angaiah, S. Sonochemical synthesis of a 2D–2D MoSe2/graphene nanohybrid electrode material for asymmetric supercapacitors. Sustain. Energy Fuels 2019, 3, 467–477. [Google Scholar] [CrossRef]
Figure 1. FE−SEM image of (a) CNF (d) CNF−FeN@GnP, high−magnification of (b) Co3O4/CNF (e) Co3O4/CNF−FeN@GnP and low−magnification of (c) Co3O4/CNF (f) Co3O4/CNF−FeN@GnP.
Figure 1. FE−SEM image of (a) CNF (d) CNF−FeN@GnP, high−magnification of (b) Co3O4/CNF (e) Co3O4/CNF−FeN@GnP and low−magnification of (c) Co3O4/CNF (f) Co3O4/CNF−FeN@GnP.
Energies 14 07621 g001
Figure 2. Contact angle image of (a) CNF (b) CNF−FeN@GnP (c) FeN@GnP powder (d) comparison graph of contact angle.
Figure 2. Contact angle image of (a) CNF (b) CNF−FeN@GnP (c) FeN@GnP powder (d) comparison graph of contact angle.
Energies 14 07621 g002
Figure 3. X−ray diffraction (XRD) patterns of (a) CNF (b) CNF−FeN@GnP (c) Co3O4/CNF (d) Co3O4/CNF−FeN@GnP.
Figure 3. X−ray diffraction (XRD) patterns of (a) CNF (b) CNF−FeN@GnP (c) Co3O4/CNF (d) Co3O4/CNF−FeN@GnP.
Energies 14 07621 g003
Figure 4. Fourier transform infrared spectroscopy (FT−IR) spectra of CNF and CNF−FeN@GnP.
Figure 4. Fourier transform infrared spectroscopy (FT−IR) spectra of CNF and CNF−FeN@GnP.
Energies 14 07621 g004
Figure 5. (a) Nitrogen adsorption−desorption isotherms and (b) pore distribution of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Figure 5. (a) Nitrogen adsorption−desorption isotherms and (b) pore distribution of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Energies 14 07621 g005
Figure 6. Thermogravimetric analysis (TGA) curves of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Figure 6. Thermogravimetric analysis (TGA) curves of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Energies 14 07621 g006
Figure 7. Nyquist plot of (a) CNF and CNF−FeN@GnP, (b) Co3O4/CNF and Co3O4/CNF−FeN@GnP, (c) galvanostatic charge discharge (GCD) curve (d) cyclic voltammetry at 20mV/s, and (e) cyclic stability of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Figure 7. Nyquist plot of (a) CNF and CNF−FeN@GnP, (b) Co3O4/CNF and Co3O4/CNF−FeN@GnP, (c) galvanostatic charge discharge (GCD) curve (d) cyclic voltammetry at 20mV/s, and (e) cyclic stability of CNF, CNF−FeN@GnP, Co3O4/CNF, and Co3O4/CNF−FeN@GnP.
Energies 14 07621 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, H.-Y.; Lee, S.-Y.; Jeon, I.-Y.; Shin, J.; Ju, Y.-W. Influence of the Hydrophilic Surface of Nanofiber Support on the Performance of Hybrid Supercapacitors. Energies 2021, 14, 7621. https://doi.org/10.3390/en14227621

AMA Style

Kim H-Y, Lee S-Y, Jeon I-Y, Shin J, Ju Y-W. Influence of the Hydrophilic Surface of Nanofiber Support on the Performance of Hybrid Supercapacitors. Energies. 2021; 14(22):7621. https://doi.org/10.3390/en14227621

Chicago/Turabian Style

Kim, Hyo-Young, Seon-Yeong Lee, In-Yup Jeon, Jeeyoung Shin, and Young-Wan Ju. 2021. "Influence of the Hydrophilic Surface of Nanofiber Support on the Performance of Hybrid Supercapacitors" Energies 14, no. 22: 7621. https://doi.org/10.3390/en14227621

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