Next Article in Journal
Optimal Polyethyleneimine Molecular Weight and Arrangement for Modification of γ-Cyclodextrin Metal Organic Frameworks (γ-CD-MOFs) for Post-Combustion CO2 Capture
Previous Article in Journal
Layered Hybrid iron Fluorides
Previous Article in Special Issue
Conductive Covalent Organic Frameworks Meet Micro-Electrical Energy Storage: Mechanism, Synthesis and Applications—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New High-Current Electrochemical Capacitor Using MnO2-Coated Vapor-Grown Carbon Fibers

1
Institute of New Energy on Chemical Storage and Power Sources, School of Chemical and Environmental Engineering, Yancheng Teachers University, Yancheng 224000, China
2
School of Chemistry and Materials Science, Nanjing University of Information Science and Technology, Nanjing 210044, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(10), 1444; https://doi.org/10.3390/cryst12101444
Submission received: 7 September 2022 / Revised: 10 October 2022 / Accepted: 11 October 2022 / Published: 13 October 2022
(This article belongs to the Special Issue Supercapacitors with High Energy Density)

Abstract

:
Composites of MnO2 nanosheet arrays vertically grown on the surface of vapor-grown carbon fibers (VGCFs) are fabricated by a low-temperature redox reaction between KMnO4 and the VGCFs. An assembled AC/0.5 M K2SO4/MnO2@VGCF electrochemical capacitor exhibits a higher specific capacitance, as well as a better rate capability, at a fast-current density compared to the capacitor built on hydrothermally prepared, standalone MnO2. Electrochemical tests revealed that VGCFs act as a conducting matrix, which effectively improves the conductivity of MnO2 nanosheets during cycling.

1. Introduction

Electrochemical devices that can store energy have created an interesting world in recent decades. These devices, including fuel cells, batteries and supercapacitors (SCs), possess many advantages, such as high-energy-utilization efficiency, a low output of carbon dioxide and the promotion of renewable energy use [1]. Among them, SCs can compensate the shortage of fuel cells and batteries in power [2]. Therefore, supercapacitor technologies are deemed as a better way to store electricity [3].
So far, a lot of transition metal oxides (TMOs) (e.g., RuO2 [4], MnO2 [5], MoO3 [6], V2O5 [7] and Co3O4 [8]) have been studied in SCs. By comparing them, MnO2 possesses the merits of resource abundance, low cost and toxicity, which has been researched widely. Moreover, it has a relatively broad electrochemical window (e.g., 0–1 V vs. SCE) in water-based electrolytes [9]. Generally, composites combing conductive frameworks with a large specific surface area and nanostructured TMOs are considered to be potential electrode materials in SCs. The most used conductive frameworks are templated mesoporous carbon [10], carbon nanotubes (CNTs) [11], carbon nanofoams [12], carbon fiber paper [13] and graphene [14]. Although the MnO2/carbon composites have been investigated extensively, this work is still distinctly different from previously reported works [15,16,17,18]. For example, the morphology of the MnO2/VGCF composites presented in this work is completely different when compared to those in the review article [15], which results in a very high sweep rate of 500 mV s−1. As a result, vapor-grown carbon fibers (VGCFs) are deemed a promising candidate because of their abundance, low cost and good electrical conductivity [19]. Additionally, they are a type of multiwalled carbon nanotubes, which possess much more excellent properties (e.g., thermal stability, chemical stability, mechanical strength [20] and a high aspect ratio of approximately 250–2000 (100 μm in length and 50–200 nm in diameter [21])). In terms of synthesis, various methods are used to prepare MnO2/carbon composites, e.g., redox reaction [22], electro-deposition [23], ball milling [24], thermal decomposition [25], hydrothermal methods [13], chemical co-precipitation [26], sonochemical synthesis [27], microwave irradiation [28] and physical mixing [29]. To the best of our knowledge, the permanganate can be reduced by surface carbon from conductive frameworks to generate MnO2 nanosheets, which simultaneously deposit on the surface of conductive frameworks. This is the most appealing strategy because of its inherently self-supporting feature [30] and the existence of synergistic effects in physical and chemical phenomena [31]. Li and co-workers also adopted the same strategy by reducing KMnO4 with CNTs [16]. However, CNTs can be partly cleaved in the process of synthesis because of its smaller diameter, thereby destroying the integrity of the material’s structure. On the contrary, the VGCFs used in this work can maintain their structure integrity, thereby leading to a very large sweep rate. Wang and co-workers synthesized MnO2/CNTs composites using Mn2+ ions to reduce KMnO4 [17]. It is obvious that the interaction between MnO2 and CNTs only involves physical adsorption instead of synergistic effects. Furthermore, Wu and co-workers used a similar method to prepare MnO2/CNTs composites with the assistance of surfactants [18], which made the fabrication process more complicated.
In this work, a co-axial electrode material with a flower-like structure is introduced for the electrochemical capacitor by assembling MnO2 nanosheet arrays onto the VGCFs’ surface without surfactant assistance, which is depicted in Scheme 1. The axial materials include VGCFs, which possess a large surface area and serve as a conductive agent, thereby reducing the mass ratio of inactive materials in electrodes. Additionally, the VGCFs are treated in an acidic solution before use, which makes the surface more hydrophilic due to the presence of –COOH and –OH groups. Moreover, some impurities can also be removed thoroughly. The structural characterization and electrochemical performances of the MnO2@VGCF composites are discussed and compared to the MnO2 materials synthesized in a similar condition, but without the use of the VGCFs, known as standalone MnO2.

2. Experimental Section

2.1. Chemicals and Materials

Potassium permanganate (KMnO4, 99%), potassium sulfate (K2SO4, 99%), manganese sulfate (MnSO4·H2O, 99%), ammonium persulfate ((NH4)2S2O8, 98%) and ammonium sulfate ((NH4)2SO4, 99%) were all purchased from Sigma-Aldrich, Shanghai, China. Acetylene black (Kappa 100, 99.9%) and poly (tetrafluoroethylene) (PTFE, 60%) solution were purchased from Canrd, Guangdong, China. Vapor-grown carbon fibers (VGCFs, diameters 120~150 nm) were purchased from Showa Denko, Japan.

2.2. Synthesis of the MnO2 Nanorods and the MnO2@VGCFs

6 M HNO3 was used to treat commercial VGCFs and refluxed under 60 °C for two hours, thereby promoting the formation of –COOH and –OH hydrophilic groups on the VGCF surface and disposing of impurities. After being washed, the VGCFs were filtered and collected. The MnO2 nanosheets were vertically deposited on the VGCF surface via redox reaction, following a modified method from the literature [32]. A schematic illustration is shown in Scheme 1. Briefly, 80 mg of VGCFs was placed in 100 mL of water containing 5 mmol of H2SO4 in a 0.25 L flask. Subsequently, 1 mmol of KMnO4 was gradually added. After 20 min of stirring, the above mixture was kept under 85 °C for 3 h. The as-synthesized powder was filtrated and collected.
For comparison, the MnO2 nanorods were synthesized via a hydrothermal method. In detail, the reaction solution was mixed with MnSO4, (NH4)2S2O8 and (NH4)2SO4 in a 1:1:4 molar ratio. Then, a 35 mL mixture was poured into a 0.05 L autoclave. Afterwards, the autoclave was kept under 140 °C for 1 h. Finally, the obtained powder was collected, washed and dried overnight at 90 °C.

2.3. Fabrication of Cathodes

The cathodes were fabricated by mixing the sample (MnO2@VGCF or standalone MnO2), poly (tetrafluoroethylene) (PTFE) and acetylene black. Their mass ratio was set as 75:10:15. The resulting flakes were beaten into small discs, each being about 2 mg in mass, 0.15 mm in thickness and 0.64 cm2 in area. At last, these discs were kept at 120 °C for 12 h. The activated carbon (AC) anode was fabricated in an identical way. The aqueous electrolyte was the 0.5 M K2SO4 solution.

2.4. Characterization Techniques

SEM was performed on a FEI Nova NanoSEM NPE207 (America), and the acceleration voltage was 20 kV. Bruker D8 Discover (Germany) was applied to measure XRD data, the parameter of wavelength was set as 1.54 Å and the Cu X-ray tube was used in a X-ray diffractometer. An STA209 PC (Germany) analyzer was conducted on oxygen flow to obtain thermogravimetry (TG) data. An electrochemical workstation (CHI604 C, Chenhua Ltd. Co., Shanghai, China) was used to test electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV). In a three-electrode system, galvanostatic charge/discharge (GCD) was carried out from 0 to 1 V versus SCE. A two-electrode supercapacitor was used for the investigation, consisting of the cathode (MnO2@VGCFs or MnO2) and the AC anode, filled with 0.5 M of K2SO4 aqueous electrolyte.

2.5. Calculation of Energy and Power Densities

The discharge capacitance C (F g−1), power density P (W kg−1) and energy density E (Wh kg−1) were calculated from CV curves and GCD profiles as seen below:
C = I d U 2 v m Δ U
E = I t U a m
P = I U a m
where I refers to the applied current, ΔU stands for the electrochemical window, m represents the total masses of the active materials in both the cathode (e.g., MnO2@VGCF or MnO2) and the anode (e.g., activated carbon), v typifies the sweep rate, 2 is a factor because both negative and positive scans are included, t stands for the time of one discharge cycle and Ua is the average voltage, which equates to (U1 + U2)/2; herein, U1 is the initial voltage and U2 is the cut-off voltage of the electrochemical window.

3. Results and Discussion

Usually, the following stoichiometric reaction formula is obeyed to synthesize MnO2, which can deposit different carbonaceous substrates by redox reaction in situ [33].
4 KMnO 4 + 3 C + H 2 O = 4 MnO 2 + 2 KHCO 3 + K 2 CO 3
In this work, the VGCFs were utilized as both a reducing agent and as a substrate for MnO2 growth. Figure 1 exhibits the scanning electron microscopy (SEM) of the MnO2@VGCF composites and the standalone MnO2 nanorods. This microscopy reveals that the as-prepared MnO2@VGCFs (Figure 1a,b) existed in the shape of a 1D co-axial microstructure, along with a width of about 600 nm and a length of 5–10 μm. This microstructure was assembled from the MnO2 nanosheet arrays vertically grown on the VGCF surface. This heterostructure makes the MnO2 more electrochemically active because of its inherent connection with the VGCFs while leaving a straightforward ingress/egress of the electrolyte ions, thereby enhancing the surface pseudo-capacitance and electric double-layer capacitance (EDLC) formation. As for the standalone MnO2 (Figure 1c), it is apparent that these MnO2 nanorods are about 300–500 nm and 20 nm in length and width, respectively.
Figure 2 displays X-ray diffraction (XRD) patterns related to the MnO2@VGCF composites and the standalone MnO2. All diffraction characteristic peaks in Figure 2a were referred to as the hexagonal birnessite MnO2, which is in good agreement with the standard card (JCPDS 18-0802) [34]. The intensity of the (002) plane from the composites is much stronger compared to other diffraction peaks, signifying the presence of graphite from the VGCFs at about 2θ = 26° [35]. Figure 2b exhibits the characteristic of δ-MnO2 (potassium birnessite) that is consistent with the standard card (JCPDS 15-0604) [36]. Additionally, an impurity broad peak was detected at 2θ = 22°, which was ascribed to the (101) plane of γ-MnO2 (ramsdellite, JCPDS 44-0142) [37].
Figure 3 exhibits the TG curves of the MnO2@VGCF composites. Free water and crystal-water loss could be read at under 400 °C. Subsequently, the combustion of the VGCFs occurred from 400 to 550 °C. The mass loss related to the reduction in MnO2→1/2Mn2O3 + 1/2O2 was observed between 550–650 °C [38,39], which is similar to that of the standalone MnO2. From the TGA results, the mass percentage of the MnO2 in the MnO2/VGCF composites was calculated to be about 65%.
Figure 4a displays the cyclic voltammetry (CV) of the MnO2@VGCF composites in the 0.5 M K2SO4 electrolyte between 0–1 V (vs. SCE). It is obvious that the MnO2@VGCFs composites showed an ideal symmetric rectangular shape: that is to say that this is typical pseudo-capacitive behavior. Interestingly, the rectangular shape was still symmetric at a high scan rate of 500 mV s−1, signifying that the composites possess good rate capabilities [40]. Figure 4b–d demonstrate the CV curves of the MnO2 in 0.5 M K2SO4 electrolytes at 10, 100 and 500 mV s−1, respectively. At the slow sweep rate of 10 and 100 mV s−1, the MnO2 nanorods showed small but distinct redox peaks, which is possible because of the reversible ingress/egress of the potassium ions into the MnO2 [41,42]. However, in the MnO2@VGCF composites, there were no apparent redox peaks. This may be explained by the heterostructure, which makes the deposited MnO2 nanosheets more electrochemically active and makes the intercalation of the electrolyte ions easier, thereby promoting pseudo-capacitance formation. This result is unanimous with what has been discussed in the literature [35,43].
At the faster sweep rate of 500 mV s−1 (Figure 4d), these redox peaks in the MnO2 nanorods diminished. In addition, the CV curves demonstrated an obvious polarization that deviated from pure capacitive behavior, which might be because the MnO2 nanorods cannot afford a very high current [5].
The influences of sweep rates on the capacitance for standalone MnO2 nanorods and MnO2@VGCF composites are demonstrated in Figure 4e. By integrating the areas of the CV curves, the specific capacitances can be found. In regions of slow scan rates, the capacitance values of the standalone MnO2 nanorods were greater than those of the MnO2@VGCF composites. This finding can be explained as follows. At slow scan rates, the standalone MnO2 nanorods produced an additional battery-type capacity contribution besides the absorption/desorption reaction, resulting in a higher capacitance than that of the MnO2@VGCF composites. After increasing the scan rate, the effective contribution of the redox reaction was controlled to some extent, owing to poor conductivity, causing the falloff of the capacitance of the standalone MnO2 nanorods. However, the capacitance of the MnO2@VGCF composites only slightly declined. This can be accredited to the VGCFs’ conductive additive, which can load large currents during discharging/charging processes and improve the rate capability of deposited MnO2 nanosheets. As a consequence, the MnO2@VGCF composites showed a larger capacitance compared to the MnO2 nanorods at fast sweep rates.
For further understanding, the impedance of the MnO2 nanorods and the MnO2@VGCF composites after 20 cycles were measured from 105 to 10−1 Hz at their open circuit potentials. An estimation of the impedance data was carried out with ZView software version 2.0. The tested impedance data were analyzed using a complex nonlinear least-squares fitting (NLSF) method [28,44] based on the equivalent circuit (Figure 4f). The experimental and fitting data are displayed in Figure 5a,b, and the values of the elements in the equivalent circuit are listed in Table 1. The intercept at the real part (Z’) axis at the highest frequency stands for the solution resistance (Rs). The Rs of the two electrodes were almost identical. Rct represents the charge-transfer resistance caused by Faradaic reactions of the active materials. The MnO2 electrode presented an Rct value of 3.49 Ω, whereas the MnO2@VGCF electrode showed a lower value of Rct of 1.44 Ω, thanks to the high electronic conductivity of the VGCFs. This is also a reason for the good CV performances of the MnO2@VGCF composites at a large sweep rate of 500 mV s−1. Inductance is the property of an electric circuit element that demonstrates opposition to the change of current flowing in it. From the viewpoint of common terminology in electric circuits, an inductor is defined as a two-terminal element, consisting of a winding of many (mostly denoted as N) turns to introduce inductance into an electric circuit [45]. In this work, an inductor (L) was introduced in series with a constant phase element (QCPE), and then paralleled with the charge-transfer resistor (Rct) in the equivalent circuit, which refers to the negative values of the imaginary part in the high frequency. The close fitting of the EIS data (Figure 5a,b) on the basis of the equivalent circuit (Figure 4f) implies that this equivalent circuit model can successfully represent the electrochemical processes of the active materials. The values of the inductance for the MnO2 and the MnO2@VGCF electrodes were 2 × 10−6 and 1 × 10−7 H, respectively. These values suggest that the as-prepared materials possessed metallic properties [46]. In the region of low frequency, the linear part was attributed to the frequency dependence of K+ ions transport/diffusion in the host, which is normally called Warburg diffusion [45]. Therefore, the Warburg impedance [47] is:
W o = Δ V Δ I = R 1 l ( D i ω ) P tan h [ l · ( i ω D ) P ]
R = k B T l ( n e ) 2 A D C o
where ω is angular frequency, i is equal to 1 ,   kB is Boltzmann constant, T is time, l is diffusion length, n is number of transfer electron, A is area of electrode, D is diffusion coefficient, Co is ion concentration.
The constant phase element (QCPE) represents the non-uniform charge distribution at the grain interfaces, and describes the deformed nature of the semicircles in Nyquist plots, which is written below [48]:
Q C P E = ( Y o ( j ω ) P ) 1
where Yo is determined by a combination of some properties of both electro-active parts and surfaces. The values of the parameter Yo can be obtained from the fitting results, which equate to those of the CPE-T and Wo-T for the CPE element and Warburg element, respectively. The P is frequency power, and when P = 0, 1 and −1, this refers to a resistor, capacitor and inductor, respectively; when P = 0.5, it represents Warburg diffusion.
Figure 6a manifests the constant current charge/discharge profiles of an assembled supercapacitor, which consists of the MnO2@VGCFs’ cathode and the AC anode using a 0.5 M K2SO4 electrolyte. The mass ratios of AC to MnO2 and MnO2@VGCF were fixed at 1:1.25. The electrochemical window for the supercapacitor was between 0 and 1.8 V. The GCD profiles are, on the whole, symmetric, which also indirectly proves ideal pseudo-capacitive properties. The Ragone plots of two aqueous SCs are shown in Figure 6b. The total masses of the two electrodes were adopted for calculations. The energy density of the MnO2 nanorods was 16 Wh kg−1 at 985 W kg−1, diminishing to 7 Wh kg−1 at 8600 W kg−1. On the contrary, the energy density of the MnO2@VGCF composites was 22.6 Wh kg−1 at 1000 W kg−1. Moreover, it retained a superior rate capability, accompanied with 10.7 Wh kg−1 even at 19,200 W kg−1. These data are much better than that of the MnO2 nanorods. Distinctly, this good rate capability is credited to the VGCFs’ co-axial structure, which is well matched with the CV results. Additionally, the rate property of the composites is better than that of previously published results of MnO2 [49,50], which is possible because the mobility of hydrated potassium ions is higher than that of hydrated sodium ions in aqueous solution [5,51].
The cycling performance of the assembled AC/0.5 M K2SO4/MnO2@VGCF SCs at 1000 mA g−1 is displayed in Figure 7. The discharge capacitance of the MnO2@VGCF composites was 50 F/g, which is larger when compared with the standalone MnO2 nanorods. After 10,000 cycles, there was no apparent capacitance fading. In contrast, the MnO2 nanorods exhibited a slight capacitance fading in the K2SO4 aqueous solution. The loss percentage was about 4% after 10,000 cycles. Consequently, the VGCFs’ co-axial structure for the MnO2@VGCFs composites presented good cycling and high specific capacitance at a high current density, which is compossible with the beforehand tests of rate behavior.

4. Conclusions

In this work, we prepared a composite of MnO2 nanosheet arrays vertically grown on the VGCF surface via a low-temperature redox reaction between KMnO4 and the VGCFs. The as-synthesized MnO2@VGCF composites demonstrated a larger capacitance, superior rate behavior and cycle performance compared with the standalone MnO2 synthesized by the hydrothermal method, which suggests that VGCFs work well as a conductive additive for aqueous electrochemical capacitors.

Author Contributions

Conceptualization, Y.L. and R.X.; data curation, Y.L. and K.S.; methodology, Y.X., Y.C. and Z.Z.; validation, Y.S. and J.W.; writing—original draft, Y.L.; writing—review and editing, F.Y. and R.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Jiangsu Natural Science Foundation Youth Fund Project (BK20220700) and the Project of Natural Science Research in Colleges and Universities of Jiangsu Province (21KJD150004). X.R. also thanks the Jiangsu Province Education Department Major Project (19KJA140003), Jiangsu Agriculture Science and Technology Innovation Fund (JASTIF) (CX(20)3169), Innovation and entrepreneurship training program for College Students (202010324017Z). Opening Project of Jiangsu Province Engineering Research Center of Agricultural Breeding Pollution Control and Resource (No. 2021ABPCR015), Open Fund of Key Laboratory for Advanced Technology in Environmental Protection of Jiangsu Province (No. JBGS023), and the Jiangsu Provincial Key Laboratory of Coastal Wetland Bioresources (No. JKLBS2020004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data needed to evaluate the conclusions in the paper are presented in the paper. Additional data related to this paper may be requested from the authors.

Acknowledgments

We thank F. Aghamiri and T.K.A. Hoang for their kind help in language polishing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, F.X.; Xiao, S.Y.; Zhu, Y.S.; Chang, Z.; Hu, C.L.; Wu, Y.P.; Holze, R. Spinel LiMn2O4 nanohybrid as high capacitance positive electrode material for supercapacitors. J. Power Sources 2014, 246, 19–23. [Google Scholar] [CrossRef]
  2. Liu, Y.; Zhang, B.H.; Wang, F.X.; Wen, Z.B.; Wu, Y.P. Nanostructured intercalation compounds as cathode materials for supercapacitors. Pure Appl. Chem. 2014, 86, 593–609. [Google Scholar] [CrossRef]
  3. Naoi, K.; Naoi, W.; Aoyagi, S.; Miyamoto, J.I.; Kamino, T. New generation “nanohybrid supercapacitor”. Acc. Chem. Res. 2013, 46, 1075–1083. [Google Scholar] [CrossRef] [PubMed]
  4. Algharaibeh, Z.; Liu, X.R.; Pickup, P.G. An asymmetric anthraquinone-modified carbon/ruthenium oxide supercapacitor. J. Power Sources 2009, 187, 640–643. [Google Scholar] [CrossRef]
  5. Qu, Q.T.; Zhang, P.; Wang, B.; Chen, Y.H.; Tian, S.; Wu, P.Y.; Holze, R. Electrochemical performance of MnO2 nanorods in neutral aqueous electrolytes as a cathode for asymmetric supercapacitors. J. Phys. Chem. C 2009, 113, 14020–14027. [Google Scholar] [CrossRef]
  6. Liu, Y.; Zhang, B.H.; Yang, Y.Q.; Chang, Z.; Wen, Z.B.; Wu, P.Y. Polypyrrole-coated α-MoO 3 nanobelts with good electrochemical performance as anode materials for aqueous supercapacitors. J. Mater. Chem. A 2013, 1, 13582–13587. [Google Scholar] [CrossRef]
  7. Tang, W.; Gao, X.W.; Zhu, Y.S.; Yue, Y.B.; Shi, Y.; Wu, Y.P.; Zhu, K. A hybrid of V2O5 nanowires and MWCNTs coated with polypyrrole as an anode material for aqueous rechargeable lithium batteries with excellent cycling performance. J. Mater. Chem. 2012, 22, 20143–20145. [Google Scholar] [CrossRef]
  8. Zhang, Y.Q.; Li, L.; Shi, S.J.; Xiong, Q.Q.; Zhao, X.Y.; Wang, X.L.; Gu, C.D.; Tu, J.P. Synthesis of porous Co3O4 nanoflake array and its temperature behavior as pseudo-capacitor electrode. J. Power Sources 2014, 256, 200–205. [Google Scholar] [CrossRef]
  9. Kundu, M.; Liu, L.F. Direct growth of mesoporous MnO2 nanosheet arrays on nickel foam current collectors for high-performance pseudocapacitors. J. Power Sources 2013, 243, 676–681. [Google Scholar] [CrossRef]
  10. 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]
  11. Hu, L.B.; Chen, W.; Xie, X.; Liu, N.A.; Yang, Y.; Wu, H.; Yao, Y.; Pasta, M.; Alshareef, H.N.; Cui, Y. Symmetrical MnO2–carbon nanotube–textile nanostructures for wearable pseudocapacitors with high mass loading. ACS Nano 2011, 5, 8904–8913. [Google Scholar] [CrossRef] [PubMed]
  12. Raymundo-Piñero, E.; Khomenko, V.; Frackowiak, E.; B’eguin, F. Performance of manganese oxide/CNTs composites as electrode materials for electrochemical capacitors. J. Electrochem. Soc. 2005, 152, A229–A235. [Google Scholar] [CrossRef]
  13. Huang, H.J.; Wang, X. Graphene nanoplate-MnO2 composites for supercapacitors: A controllable oxidation approach. Nanoscale 2011, 3, 3185–3191. [Google Scholar] [CrossRef]
  14. Wang, H.L.; Casalongue, H.S.; Liang, Y.Y.; Dai, H.J. Ni(OH)2 nanoplates grown on graphene as advanced electrochemical pseudocapacitor materials. J. Am. Chem. Soc. 2010, 132, 7472–7477. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, Q.; Zhang, D.; Miao, Z.; Zhang, X.; Chou, S. Research Progress in MnO2–Carbon Based Supercapacitor Electrode Materials. Small 2018, 14, 1702883. [Google Scholar] [CrossRef]
  16. Li, L.; Hu, Z.A.; An, N.; Yang, Y.Y.; Li, Z.M.; Wu, H.Y. Facile synthesis of MnO2/CNTs composite for supercapacitor electrodes with long cycle stability. J. Phys. Chem. C 2014, 118, 22865–22872. [Google Scholar] [CrossRef]
  17. Wang, K.; Gao, S.; Du, Z.; Yuan, A.; Lu, W.; Chen, L. MnO2-Carbon nanotube composite for high-areal-density supercapacitors with high rate performance. J. Power Sources 2016, 305, 30–36. [Google Scholar] [CrossRef]
  18. Wu, P.; Cheng, S.; Yang, L.; Lin, Z.; Gui, X.; Ou, X.; Zhou, J.; Yao, M.; Wang, M.; Zhu, Y.; et al. Synthesis and characterization of self-standing and highly flexible δ-MnO2@CNTs/CNTs composite films for direct use of supercapacitor electrodes. ACS Appl. Mater. Interfaces 2016, 8, 23721–23728. [Google Scholar] [CrossRef] [PubMed]
  19. Sekhar, S.C.; Nagaraju, G.; Cha, S.M.; Yu, J.S. Birnessite-type MnO2 nanosheet arrays with interwoven arrangements on vapor grown carbon fibers as hybrid nanocomposites for pseudocapacitors. Dalton Trans. 2016, 45, 19322–19328. [Google Scholar] [CrossRef] [Green Version]
  20. Nakamura, K.; Sato, Y.; Takase, T. Surface oxidation and/or corrosion behavior of vapor-grown carbon fibers (VGCFs) in nitric acid, hydrogen peroxide, and sulfuric acid solutions. Diam. Relat. Mater. 2017, 76, 108–114. [Google Scholar] [CrossRef]
  21. Singh, A.K.; Parhi, A.; Panda, B.P.; Mohanty, S.; Nayak, S.K.; Gupta, M.K. Aligned multi-walled carbon nanotubes (MWCNT) and vapor grown carbon fibers (VGCF) reinforced epoxy adhesive for thermal conductivity applications. J. Mater. Sci. Mater. Electron. 2017, 28, 17655–17674. [Google Scholar] [CrossRef]
  22. Dong, X.P.; Shen, W.H.; Gu, J.L.; Xiong, L.M.; Zhu, Y.F.; Li, H.; Shi, J.L. MnO2-embedded-in-mesoporous-carbon-wall structure for use as electrochemical capacitors. J. Phys. Chem. B 2006, 110, 6015–6019. [Google Scholar] [CrossRef] [PubMed]
  23. Lee, C.Y.; Tsai, H.M.; Chuang, H.J.; Li, S.Y.; Lin, P.; Tseng, T.Y. Characteristics and electrochemical performance of supercapacitors with manganese oxide-carbon nanotube nanocomposite electrodes. J. Electrochem. Soc. 2005, 152, A716–A720. [Google Scholar] [CrossRef] [Green Version]
  24. Zhou, Y.K.; He, B.L.; Zhang, F.B.; Li, H.L. Hydrous manganese oxide/carbon nanotube composite electrodes for electrochemical capacitors. J. Solid State Electrochem. 2004, 8, 482–487. [Google Scholar] [CrossRef]
  25. Fan, Z.; Chen, J.H.; Wang, M.Y.; Cui, K.Z.; Zhou, H.H.; Kuang, Y.F. Preparation and characterization of manganese oxide/CNT composites as supercapacitive materials. Diam. Relat. Mater. 2006, 15, 1478–1483. [Google Scholar] [CrossRef]
  26. Wang, H.L.; Yang, Y.; Liang, Y.Y.; Robinson, J.T.; Li, Y.G.; Jackson, A.; Cui, Y.; Dai, H.J. Graphene-wrapped sulfur particles as a rechargeable lithium–sulfur battery cathode material with high capacity and cycling stability. Nano Lett. 2011, 11, 2644–2647. [Google Scholar] [CrossRef] [Green Version]
  27. Kawaoka, H.; Hibino, M.; Zhou, H.S.; Honma, I. Sonochemical synthesis of amorphous manganese oxide coated on carbon and application to high power battery. J. Power Sources 2004, 125, 85–89. [Google Scholar] [CrossRef]
  28. Yan, J.; Fan, Z.J.; Wei, T.; Qian, W.Z.; Zhang, M.L.; Wei, F. Fast and reversible surface redox reaction of graphene–MnO2 composites as supercapacitor electrodes. Carbon 2010, 48, 3825–3833. [Google Scholar] [CrossRef]
  29. Fischer, A.E.; Pettigrew, K.A.; Rolison, D.R.; Stroud, R.M.; Long, J.W. Incorporation of homogeneous, nanoscale MnO2 within ultraporous carbon structures via self-limiting electroless deposition: Implications for electrochemical capacitors. Nano Lett. 2007, 7, 281–286. [Google Scholar] [CrossRef]
  30. Zhang, X.X.; Ran, F.; Fan, H.L.; Tan, Y.T.; Zhao, L.; Li, X.M.; Kong, L.B.; Kang, L. A dandelion-like carbon microsphere/MnO2 nanosheets composite for supercapacitors. J. Energy Chem. 2014, 23, 82–90. [Google Scholar] [CrossRef]
  31. Terasawa, N.; Asaka, K. Superior performance of PEDOT: Poly (4-styrenesulfonate)/vapor-grown carbon fibre/ionic liquid actuators exhibiting synergistic effects. Sens. Actuators B Chem. 2017, 248, 273–279. [Google Scholar] [CrossRef]
  32. Tang, W.; Hou, Y.Y.; Wang, F.X.; Liu, L.L.; Wu, Y.P.; Zhu, K. LiMn2O4 nanotube as cathode material of second-level charge capability for aqueous rechargeable batteries. Nano Lett. 2013, 13, 2036–2040. [Google Scholar] [CrossRef] [PubMed]
  33. Jin, X.; Zhou, W.; Zhang, S.; Chen, G.Z. Nanoscale microelectrochemical cells on carbon nanotubes. Small 2007, 3, 1513–1517. [Google Scholar] [CrossRef]
  34. Miao, J.; Lin, H.; Mao, Z.; He, S.J.; Xu, M.; Li, Q. Electrochemical performance of Sn-doped δ-MnO2 hollow nanoparticles for supercapacitors. J. Mater. Sci. Mater. Electron. 2018, 29, 2689–2697. [Google Scholar] [CrossRef]
  35. Lei, Z.B.; Shi, F.H.; Lu, L. Incorporation of MnO2-coated carbon nanotubes between graphene sheets as supercapacitor electrode. ACS Appl. Mater. Interfaces 2012, 4, 1058–1064. [Google Scholar] [CrossRef] [PubMed]
  36. Takashima, T.; Hashimoto, K.; Nakamura, R. Inhibition of charge disproportionation of MnO2 electrocatalysts for efficient water oxidation under neutral conditions. J. Am. Chem. Soc. 2012, 134, 1519–1527. [Google Scholar] [CrossRef]
  37. Gao, F.Y.; Tang, X.L.; Yi, H.H.; Chu, C.C.; Li, N.; Li, J.; Zhao, S.Z. In-situ DRIFTS for the mechanistic studies of NO oxidation over α-MnO2, β-MnO2 and γ-MnO2 catalysts. Chem. Eng. J. 2017, 322, 525–537. [Google Scholar] [CrossRef]
  38. Ruetschi, P.; Giovanoli, R. Cation vacancies in MnO2 and their influence on electrochemical reactivity. J. Electrochem. Soc. 1988, 135, 2663–2669. [Google Scholar] [CrossRef]
  39. Ma, R.Z.; Bando, Y.; Zhang, L.Q.; Sasaki, T. Layered MnO2 nanobelts: Hydrothermal synthesis and electrochemical measurements. Adv. Mater. 2004, 16, 918–922. [Google Scholar] [CrossRef]
  40. Zhu, D.D.; Wang, Y.D.; Yuan, G.L.; Xia, H. High-performance supercapacitor electrodes based on hierarchical Ti@MnO2 nanowire arrays. Chem. Commun. 2014, 50, 2876–2878. [Google Scholar] [CrossRef]
  41. Ma, Z.P.; Shao, G.J.; Fan, Y.Q.; Wang, G.L.; Song, J.J.; Shen, D.J. Construction of hierarchical α-MnO2 nanowires@ultrathin δ-MnO2 nanosheets core–shell nanostructure with excellent cycling stability for high-power asymmetric supercapacitor electrodes. ACS Appl. Mater. Interfaces 2016, 8, 9050–9058. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, Y.; Yuan, C.; Ye, K.; Jiang, X.; Yin, J.L.; Wang, G.; Cao, D.X. An aqueous capacitor battery hybrid device based on Na-ion insertion-deinsertion in λ-MnO2 positive electrode. Electrochim. Acta 2014, 148, 237–243. [Google Scholar] [CrossRef]
  43. Chen, Q.; Meng, Y.N.; Hu, C.G.; Zhao, Y.; Shao, H.B.; Chen, N.; Qu, L.T. MnO2-modified hierarchical graphene fiber electrochemical supercapacitor. J. Power Sources 2014, 247, 32–39. [Google Scholar] [CrossRef]
  44. Sugimoto, W.; Iwata, H.; Yokoshima, K.; Murakami, Y.; Takasu, Y. Proton and electron conductivity in hydrous ruthenium oxides evaluated by electrochemical impedance spectroscopy: The origin of large capacitance. J. Phys. Chem. B 2005, 109, 7330–7338. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, W.; Wen, T.; Hu, C.; Gopalan, A. Identification of inductive behavior for polyaniline via electrochemical impedance spectroscopy. Electrochim. Acta 2002, 47, 1305–1315. [Google Scholar] [CrossRef]
  46. Schneider, K.; Dziubaniuk, M.; Wyrwa, J. Impedance Spectroscopy of Vanadium Pentoxide Thin Films. J. Electron. Mater. 2019, 48, 4085–4091. [Google Scholar] [CrossRef] [Green Version]
  47. Lim, C.H., IV. Transport Phenomena: Lecture 20: Warburg Impedance. Available online: https://ocw.mit.edu/courses/chemical-engin-eering/10-626-electrochemical-energy-systems-spring-2014/lecture-notes/MIT10_626S14_S11lec20.pdf (accessed on 26 July 2022).
  48. Burashnikova, M.M.; Kazarinov, I.A.; Zotova, I.V. Nature of contact corrosion layers on lead alloys: A study by impedance spectroscopy. J. Power Sources 2012, 207, 19–29. [Google Scholar] [CrossRef]
  49. Wu, S.S.; Chen, W.F.; Yan, L.F. Fabrication of a 3D MnO2/graphene hydrogel for high-performance asymmetric supercapacitors. J. Mater. Chem. A 2014, 2, 2765–2772. [Google Scholar] [CrossRef]
  50. Deng, S.; Sun, D.; Wu, C.H.; Wang, H.; Liu, J.V.; Sun, Y.X.; Yan, H. Synthesis and electrochemical properties of MnO2 nanorods/graphene composites for supercapacitor applications. Electrochim. Acta 2013, 111, 707–712. [Google Scholar] [CrossRef]
  51. Qu, Q.T.; Wang, B.; Yang, L.C.; Shi, Y.; Tian, S.; Wu, Y.P. Study on electrochemical performance of activated carbon in aqueous Li2SO4, Na2SO4 and K2SO4 electrolytes. Electrochem. Commun. 2008, 10, 1652–1655. [Google Scholar] [CrossRef]
Scheme 1. Synthesis process of the MnO2@VGCF composites.
Scheme 1. Synthesis process of the MnO2@VGCF composites.
Crystals 12 01444 sch001
Figure 1. (a,b) SEM of the MnO2@VGCFs composites, (c) SEM micrograph of the MnO2 nanorods.
Figure 1. (a,b) SEM of the MnO2@VGCFs composites, (c) SEM micrograph of the MnO2 nanorods.
Crystals 12 01444 g001
Figure 2. XRD patterns of the MnO2@VGCF composites (a) and the MnO2 nanorods (b).
Figure 2. XRD patterns of the MnO2@VGCF composites (a) and the MnO2 nanorods (b).
Crystals 12 01444 g002
Figure 3. Thermogravimetric analysis of the MnO2 and the MnO2@VGCF composites in air.
Figure 3. Thermogravimetric analysis of the MnO2 and the MnO2@VGCF composites in air.
Crystals 12 01444 g003
Figure 4. (a) CV curves of MnO2@VGCFs; (bd) CV curves of MnO2 at 10, 100, 500 mV s−1 in 0.5 M K2SO4 aqueous electrolyte, respectively; (e) changes in capacitance of MnO2 and MnO2@VGCF at different scan rates; (f) equivalent circuit used for fitting impedance spectra of the MnO2 and the MnO2@VGCF.
Figure 4. (a) CV curves of MnO2@VGCFs; (bd) CV curves of MnO2 at 10, 100, 500 mV s−1 in 0.5 M K2SO4 aqueous electrolyte, respectively; (e) changes in capacitance of MnO2 and MnO2@VGCF at different scan rates; (f) equivalent circuit used for fitting impedance spectra of the MnO2 and the MnO2@VGCF.
Crystals 12 01444 g004
Figure 5. Nyquist plots of the MnO2 (a) and the MnO2@VGCFs (b) in 0.5 M K2SO4 aqueous electrolyte.
Figure 5. Nyquist plots of the MnO2 (a) and the MnO2@VGCFs (b) in 0.5 M K2SO4 aqueous electrolyte.
Crystals 12 01444 g005
Figure 6. (a) Potential−time profiles of the AC//MnO2@VGCF electrochemical capacitor in 0.5 M K2SO4 aqueous electrolyte and (b) Ragone plots of the AC//MnO2 and AC//MnO2@VGCF supercapacitors using 0.5 M K2SO4 electrolyte.
Figure 6. (a) Potential−time profiles of the AC//MnO2@VGCF electrochemical capacitor in 0.5 M K2SO4 aqueous electrolyte and (b) Ragone plots of the AC//MnO2 and AC//MnO2@VGCF supercapacitors using 0.5 M K2SO4 electrolyte.
Crystals 12 01444 g006
Figure 7. Cycling behavior of the AC//MnO2 (Blue dots) and AC//MnO2@VGCFs (red dots) supercapacitors at 1000 mA g−1 (black dots refer to Coulombic efficiency).
Figure 7. Cycling behavior of the AC//MnO2 (Blue dots) and AC//MnO2@VGCFs (red dots) supercapacitors at 1000 mA g−1 (black dots refer to Coulombic efficiency).
Crystals 12 01444 g007
Table 1. Fitting results of the electrochemical impedance.
Table 1. Fitting results of the electrochemical impedance.
TypesMnO2MnO2@VGCFs
Rs1.781.87
CPE-T5.43 × 10−34.46 × 10−3
CPE-P0.740.78
Rct3.491.44
L2.0 × 10−61.0 × 10−7
Wo-R8.16 × 10−38.34 × 10−4
Wo-T7.09 × 10−41.85 × 10−4
Wo-P0.450.44
Notes: Rs (solution resistance, Unit: Ω); CPE-T (constant phase element, Unit: S); CPE-P (exponent, no unit); Rct (charge-transfer resistance, Unit: Ω); L (inductance, Unit: H); Wo-R (Warburg resistance, Unit: Ω); Wo-T (Warburg conductance, Unit: S); Wo-P (exponent, no unit).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Xu, Y.; Chang, Y.; Sun, Y.; Zhao, Z.; Song, K.; Wang, J.; Yu, F.; Xing, R. A New High-Current Electrochemical Capacitor Using MnO2-Coated Vapor-Grown Carbon Fibers. Crystals 2022, 12, 1444. https://doi.org/10.3390/cryst12101444

AMA Style

Liu Y, Xu Y, Chang Y, Sun Y, Zhao Z, Song K, Wang J, Yu F, Xing R. A New High-Current Electrochemical Capacitor Using MnO2-Coated Vapor-Grown Carbon Fibers. Crystals. 2022; 12(10):1444. https://doi.org/10.3390/cryst12101444

Chicago/Turabian Style

Liu, Yu, Yu Xu, Yingna Chang, Yuzhen Sun, Zhiyuan Zhao, Kefan Song, Jindi Wang, Feng Yu, and Rong Xing. 2022. "A New High-Current Electrochemical Capacitor Using MnO2-Coated Vapor-Grown Carbon Fibers" Crystals 12, no. 10: 1444. https://doi.org/10.3390/cryst12101444

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