Next Article in Journal
Structural Stabilization of Mullite Films Exposed to Oxygen Potential Gradients at High Temperatures
Next Article in Special Issue
Ethylene Glycol Functionalized Gadolinium Oxide Nanoparticles as a Potential Electrochemical Sensing Platform for Hydrazine and p-Nitrophenol
Previous Article in Journal
Thermal Boundaries in Cone Calorimetry Testing
Previous Article in Special Issue
Multi-Layered Mesoporous TiO2 Thin Films: Photoelectrodes with Improved Activity and Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Rapid Synthesis of Mesoporous Mn2O3 Nanoparticles for Supercapacitor Applications

1
School of Semiconductor and Chemical Engineering & Solar Energy Research Center, Chonbuk National University, Jeonju 54896, Korea
2
New and Renewable Energy Materials Development Center (NewREC), Chonbuk National University, Jeonbuk 54896, Korea
*
Authors to whom correspondence should be addressed.
Coatings 2019, 9(10), 631; https://doi.org/10.3390/coatings9100631
Submission received: 20 August 2019 / Revised: 20 September 2019 / Accepted: 24 September 2019 / Published: 30 September 2019
(This article belongs to the Special Issue Mesoporous Metal Oxide Films)

Abstract

:
Mn2O3 nanomaterials have been recently composing a variety of electrochemical systems like fuel cells, supercapacitors, etc., due to their high specific capacitance, low cost, abundance and environmentally benign nature. In this work, mesoporous Mn2O3 nanoparticles (NPs) were synthesized by manganese acetate, citric acid and sodium hydroxide through a hydrothermal process at 150 °C for 3 h. The synthesized mesoporous Mn2O3 NPs were thoroughly characterized in terms of their morphology, surfaces, as well as their crystalline, electrochemical and electrochemical properties. For supercapacitor applications, the synthesized mesoporous Mn2O3 NP-based electrode accomplished an excellent specific capacitance (Csp) of 460 F·g−1 at 10 mV·s−1 with a good electrocatalytic activity by observing good electrochemical properties in a 6 M KOH electrolyte. The excellent Csp might be explained by the improvement of the surface area, porous surface and uniformity, which might favor the generation of large active sites and a fast ionic transport over the good electrocatalytic surface of the Mn2O3 electrode. The fabricated supercapacitors exhibited a good cycling stability after 5000 cycles by maintaining ~83% of Csp.

1. Introduction

A popular electrochemical energy-storage system, the supercapacitor has been a well-explored device as a heartening energy storage because of its high-power density, marvelous cycling and fast charge-discharge mechanisms [1,2,3,4]. Supercapacitors, on the basis of their charge-storage process, are classified into two types: i) double layer electrochemical capacitors (EDLCs), built from the electrode and electrolyte interface for the ions adsorption-desorption, and ii) the electrode materials-governed faradaic reaction-based pseudocapacitors [5,6,7]. In comparison with the EDLCs, the pseudocapacitors have exhibited a fast and desirable reversible redox reaction that promotes an excellent charging and discharging process, resulting in the enhancement of the charge storage capacities. In addition, pseudocapacitors display a highly competent storage device for rechargeable batteries owing to their fast energy harvesting, high energy and high-power delivery [8,9,10]. In supercapacitors, the active electrodes are normally prepared with carbon-based materials, conducting polymers, and a variety of transition metal oxide materials [11,12,13,14]. In recent years, transition metal oxides, such as RuO2, MnO2, Mn3O4, NiO, Nb2O5, V2O5, CoOx, MoO3, and TiO2, have been frequently applied in preparing an effective electrode for supercapacitors due to their abundance in nature, inexpensive nature and extraordinary redox activity [13,14,15,16,17,18,19,20,21].
Apart from other transition metal oxides, the manganese-based oxides (MnOx) and their derivatives, like MnO2, Mn2O3, and Mn3O4, are popularly used as electrode materials in supercapacitors because they have a non-toxic nature, good structural flexibility and excellent chemical and physical stability in various electrolytes [22]. In particular, Mn2O3 materials as anode materials in lithium ion batteries have shown a high capacity and demonstrate a theoretically high specific capacity of ~1018 mAh/g, while also exhibiting a high specific capacitance [23]. Until now, the Mn2O3 materials-based electrodes in supercapacitors have been less explored and could be expected to have a high capacity and storage properties as they show an excellent environment compatibility and a resistance in acidic/alkaline electrolytes. Numerous efforts have been made to improve the performance of Mn2O3-based electrodes by adopting various modifications, such as chemical modifications [24], the incorporation of high surface-area conductive materials [25,26] and nanostructure fabrication [24,27]. Therefore, the synthesis of Mn2O3 materials with a controlled size, morphology and density through a cost-effective, simple and environment friendly method could be potentially sound [28,29]. It is expected that mesoporous Mn2O3 nanomaterials might show a surface that is fruitful for fast ion transportation in electrochemical supercapacitors [30].
In this work, a rapid and low temperature hydrothermal process was used to synthesize well-crystalline mesoporous Mn2O3 materials that were successfully applied as electro-active materials for a pseudosupercapacitor. The prepared mesoporous Mn2O3 materials-based electrode exhibits a high specific capacitance of 460 F·g−1 at a scan rate of 10 mV·s−1, with a good cycling stability after 5000 cycles.

2. Materials and Methods

In a typical synthesis, a mixture of 1 g of manganese acetate (Mn(CH3CO2)2, Sigma-Aldrich, Saint Louis, MO, USA) and 0.5 g citric acid (Samchun Chemicals, Seoul, Korea) was dissolved in 100 mL of deionized (DI) water. Using an aqueous 5 M NaOH solution, the pH of the reaction mixture was adjusted to pH ~10 as the solution color changed from bright brown to dark brown. The hydrothermal process was carried out by transferring the reaction mixture into a Teflon beaker that was tightly sealed. Finally, a temperature of 150 °C was maintained for 3 h. After cooling down the autoclave, the obtained precipitates were collected by filtration, washed several times with DI water and with ethanol. The collected precipitates were dried in an oven overnight at 80 °C, and finally the obtained black powder was calcined at 500 °C in 1 h to remove other impurities.
For the electrochemical supercapacitor application, the synthesized mesoporous Mn2O3 electrode was prepared by mixing 85 wt.% Mn2O3 powders, 10 wt.% carbon black (Super P, Venatech, Seoul, South Korea), 3 wt.% carboxyl methyl cellulose (CMC, Sigma-Aldrich) and 2 wt.% polytetrafluoroethylene (PTFE, TCI chemical, Tokyo, Japan) in DI water to obtain a paste that was spread over the Ni foam using a glass rod via the rolling method. Afterward, the mesoporous Mn2O3-coated Ni foam electrodes were dried in the oven at 80 °C for 20 min to remove the solvent. For the electrochemical measurement, a three-electrode system, comprised of Mn2O3-coated Ni foam as the working electrode, Pt wire as the counter electrode and Ag/AgCl as the reference electrode, was used, and a cyclicvoltametric measurement (VersaSTAT4, AMETEK, Inc., Berwyn, PA, USA) was performed in an aqueous 6 M KOH electrolyte. All cyclicvoltametry (CV) measurements were observed at different scan rates ranging from 10 to 500 mV·s−1 in the voltage range of 0 to 1.0 V. A potentiostat/galvanostat (VersaSTAT4, AMETEK, Inc.) was used to analyze the electrochemical impedance spectroscopy (EIS) of the fabricated supercapacitor, based on the mesoporous Mn2O3 electrode, with a frequency ranging from 0.1 Hz to 1 MHz. For the calculation of Csp, the mass loading of the mesoporous Mn2O3 on the electrode was ~0.001 g.

3. Results and Discussion

The morphological features of the synthesized mesoporous Mn2O3 materials were analyzed by field emission scanning electron microscopy (FESEM, Hitachi S-4800, Tokyo, Japan) and transmission electron microscopy (TEM, JEM-ARM200F, JEOL, Peabody, MA, USA) observations. Figure 1a,b shows the FESEM images of the synthesized mesoporous Mn2O3 materials at low and high magnifications. The spherical small particles, which are highly uniform, are visible in Figure 1a. From FESEM observations, it is difficult to identify the porosity of the synthesized materials. At a high magnification mode (Figure 1b), the obtained particles possess highly porous structures with average sizes of 10–30 nm. A similar observation for synthesized mesoporous Mn2O3 materials was detected in the TEM and high-resolution transmission electron microscopy (HRTEM) analyses, as shown in Figure 1c,d. As seen in Figure 1c, the synthesized Mn2O3 materials show a similar spherical shape, with a few semi-spherical particles having average particle sizes of ~10–30 nm. A close look at Figure 1c shows that the porosity of the synthesized Mn2O3 materials might be defined by the presence of visible voids over the particle surfaces. As reported earlier [31], these voids in materials may be a detrimental factor of the porosity of the materials. The HRTEM image is shown in Figure 1d; it expresses clear lattice fringes from when the measurement was focused on one particle from Figure 1c. From the HRTEM image, the interplanar distance between two lattice fringes is estimated to be ~0.27 nm, which is well-indexed to the (222) plane of Mn2O3 [32].
The crystalline behavior and crystal planes of the synthesized mesoporous Mn2O3 materials were determined via a wide-angle X-ray diffraction (XRD, PANalytical, Malvern, United Kingdom) measurement, as shown in Figure 2. The well-defined diffraction peaks at 18.9°, 23.1°, 33.1°, 38.2°, 45.2°, 47.3°, 49.5°, 55.1° and 65.8° are associated to (200), (211), (222), (400), (332), (422), (431), (440) and (622) planes, respectively. All of the obtained diffraction peaks are assigned perfectly to the Bixbyite crystal phase α-Mn2O3, with JCPDS no. 41-1442 and space group Ia3, lattice constants a = b = c = 9.4091 Å, α = β = γ = 90°. To estimate the crystallite sizes (CS) of the synthesized mesoporous Mn2O3 materials, the Debey–Scherrer equation was used [33]:
C S = 0.95 × λ β cos   ( θ )
where β is the breadth of the observed diffraction line at its half-intensity maximum, K is the so-called shape factor, which usually takes a value of about 0.9, and λ is the wavelength of the X-ray source used in the XRD. By taking the maximum diffraction peak of (222), the crystallite size of the mesoporous Mn2O3 materials is found to be 28 nm, which is very close to the FESEM and TEM results.
Figure 3 shows the infrared (IR, Nicolet, IR300, Thermo Fisher Scientific, Waltham, MA, USA) and Raman spectroscopic studies (Raman microscope, Renishaw, UK) that define the structural properties of the synthesized mesoporous Mn2O3 materials. As seen in Figure 3a, two sharp IR bands are observed at 610 and at 520 cm−1, assigning the stretching vibrations of Mn–O units and the asymmetric Mn–O–Mn stretching vibration, respectively [34]. Other IR bands at 1640 and 3343 cm−1 are detected, related to –OH and the water species from atmospheric moisture. It is believed that the observation of the IR bands at 520 and 610 cm−1 clearly reveals the formation of Mn2O3 without other impurities. Figure 3b depicts the Raman scattering spectroscopy of the synthesized mesoporous Mn2O3 materials. The mesoporous Mn2O3 NPs present a strong Raman band at 651 cm−1, including with two weak Raman bands at 268 and at 175.0 cm−1. The strong Raman band at 651 cm−1 represents the characteristic of the Mn2O3 along with the space group Ia3 structure [35], suggesting the typical symmetric stretching Mn–O–Mn bridge in Mn2O3. The main Raman band is well-matched with the reported literature of Mn2O3 [35]. Additionally, two weak Raman bands at ~268 and ~175 are assigned to the out-of-plane bending modes of Mn2O3 and the asymmetric stretching of the bridge oxygen species (Mn–O–Mn) [36], respectively.
To explain the thermal and structural properties, a thermal gravimetric analysis (TGA, Thermal analyzer, TA Instrument Ltd., New Castle, DE, USA) was performed for the synthesized mesoporous Mn2O3 materials, as displayed in Figure 4a,b. As seen in Figure 4b, four decomposition temperatures were visibly identified in the range of 25 to 800 °C. After 500 °C, the subsequent weight loss of ~1.5% that started from 500 to 600 °C is ascribed to the thermal decomposition of Mn2O3 to MnO, a decomposition that is usual in metal oxides. In the beginning, the synthesized mesoporous Mn2O3 materials were recorded as having a very small weight loss of ~0.5% up to 500 °C, which usually occurs via the removal of water/moisture from the sample. This suggests that the synthesized mesoporous Mn2O3 NPs exhibit a remarkably good stability with a high crystalline nature of Mn2O3.
The synthesized mesoporous Mn2O3 materials were further characterized in terms of their composition and the oxidation states of the elements using an X-ray photoelectron spectroscopy (XPS, AXISNOVA CJ109, Kratos Inc., Manchester, UK) analysis. Figure 5a shows the survey XPS spectrum of the synthesized mesoporous Mn2O3 materials, revealing Mn 2p, Mn 3s and O 1s with weak C 1s peaks. The high-resolution Mn 2p XPS spectra of the synthesized mesoporous Mn2O3 materials is shown in Figure 5b and demonstrates doublet binding energies at 641.0 (Mn 2p3/2) and 652.8 eV (Mn 2p1/2). It is noted that the doublet Mn 2p, with an estimated spin–orbit splitting value of 11.8 eV, is nearly the same as the values reported for Mn2O3 [37]. Additionally, Figure 5c displays the high-resolution Mn 3s with characteristic doublet binding energies at 88.8 and 83.5 eV for Mn2O3. From the Mn 3s XPS, the peak separation for the doublet binding energies are ~5.3 eV, which is very close to the peak separation of Mn 3s in Mn2O3 [38], which again implies the formation of Mn2O3. Figure 5d depicts the high-resolution O 1s XPS spectrum. The two binding energies at 528.9 and 530.7 eV are normally related to oxygen O2− in the lattice of Mn–O–Mn, indicating the formation of Mn2O3. Therefore, the XPS analysis implied the formation of a pure Mn2O3 form without any other oxide impurities [39].
The surface and porous behavior of the synthesized mesoporous Mn2O3 materials have been elucidated by analyzing the nitrogen (N2) adsorption-desorption isotherms and BET surface analyzer, as shown Figure 6. The N2 adsorption-desorption isotherms (Figure 6a) present a regular type IV isotherm with a small hysteresis in the relative pressure (p/p0) range of 0.75–0.9, which clearly imitated the mesoporous characteristic of materials. Figure 6b shows how the pore size distribution (d) of the mesoporous Mn2O3 materials was recorded in the range of 10–30 nm, which is the case for mesoporous materials. Using a Brunauer-Emmett-Teller (BET) surface analysis, the specific surface area (s) for the synthesized mesoporous Mn2O3 materials was determined to be 76.9 m2·g−1. Thus, the synthesized Mn2O3 materials are mesoporous in nature which could provide a larger surface area for the high ion diffusion of ions in the electrolyte for high-performance supercapacitors.
The synthesized mesoporous Mn2O3 materials were utilized as electroactive materials to evaluate the supercapacitor properties. The parameters for the supercapacitor with the synthesized mesoporous Mn2O3 electrode were determined by measuring the cyclicvoltametry (CV) at different scan rates in a 6 M KOH electrolyte. Figure 7a shows a series of CV curves of the synthesized mesoporous Mn2O3 electrode at different scan rates in a 6 M KOH electrolyte. Generally, in an electrochemical reaction, the conversion from Mn3+ to Mn4+ in an Mn2O3 electrode occurs via an oxidation reaction. This oxidation reaction normally accelerates the reaction kinetics of OH over the Mn2O3 cubic lattice through the chemisorbed and/or intercalated. The charge/discharge process can be explained by the following electrochemical reaction, which involves the chemisorption/intercalation of HO over the Mn2O3 surfaces:
Mn 2 O 3   +   HO Charge D i s c h a g e   Mn 4 +   [ Mn 3 + ] OH O 3 +   e ¯  
The prepared Mn2O3 electrode depicts the oxidation-reduction pair peaks with different curve shapes of the charge–discharge curves in relation to the variation in the scan rates. The observations principally suggest the origin of the faradaic pseudo-capacitance in an alkaline electrolyte over the surface of an Mn2O3 electrode. Moreover, the well-defined oxidation reduction pair within 0−1 V can also be ascribed to a faradaic redox reaction [40,41]. The plot of the capacitance versus the scan rate for a fabricated pseudosupercapacitor based on a mesoporous Mn2O3 electrode is displayed in Figure 7b. In general, the specific capacitance (Csp) is estimated from the CV curves using the following equation [42]:
C s p = 1 m ν ( V 2 V 1 ) V 1 V 2 I ( V ) d V
where Csp is the specific capacitance (F·g−1), I is the current (A), v is the scan rate, m is the mass of the active material (g) and ∆V is the potential range (V). As seen in Figure 7b, the fabricated pseudosupercapacitor based on the mesoporous Mn2O3 electrode exhibits a high Csp of ~460 F·g−1 at a scan rate of 10 mV·s−1. This might be explained by the fact that there is an improvement of the surface area, the porous surface and uniformity, which might favor the generation of large active sites [43,44] and fast ionic transport over the surface of the Mn2O3 electrode. Specifically, the porous structure could provide a large and accessible surface area to ion adsorption, improve the accessibility of cations and shorten the ion diffusion path. The stability of the mesoporous Mn2O3 electrode is explored by observing the multicycle CV measurements in a 6 M KOH electrolyte. From Figure 7c, the slight shift in the oxidation peaks expresses the good stability of the mesoporous Mn2O3 electrode in the alkaline electrolyte. It is also seen that the anodic current in the Mn2O3 electrode is positively shifted with the increase, while the cathodic current is negatively shifted due to the increment in the electrical polarization and the fast and irreversible reactions as the scan rates increase. Here, the fast, irreversible reactions after 25 cycles might result from the accumulation of Mn4+ ions in the electrochemical system. The stability of the fabricated pseudosupercapacitor based on a mesoporous Mn2O3 electrode is shown in Figure 7d by measuring the capacitance after 5000 cycles. From Figure 7d, after 5000 cycles, the electrochemical system shows a reasonably high stability by retaining 83% of the initial capacity. Furthermore, the inset of Figure 7d presents the FESEM image of the mesoporous Mn2O3 electrode after 5000 cycles. The morphology of the Mn2O3 materials is not altered, except for an aggregation of small particles after cycling, suggesting the stability of the Mn2O3 electrode in the alkaline electrolyte. Hence, the reproducibility of a pseudosupercapacitor based on a mesoporous Mn2O3 electrode in a KOH electrolyte is remarkable and implies the low dissolution of the electro-active materials in a strong alkaline electrolyte after an electrochemical process.
The electrochemical Impedance Spectroscopy (EIS) for the fabricated pseudosupercapacitor based on the synthesized mesoporous Mn2O3 electrode was conducted to understand the electrochemical and electrical properties. Figure 8 shows the EIS plot, which was measured in the frequency range of 0.1–105 Hz. In the illustration of the equivalent circuit, the starting point in the Nyquist plot at a high-frequency region and the starting point at a low-frequency region represent the series resistance (Rs) and the charge transfer resistance (Rct) of the electrode/electrolyte interface, respectively. Cdl, W and Cpseudo explain the double layer capacitance that arose by the parallel connection to Rct, the Warburg diffusion element and the Faradaic capacitance generated by the contribution of the Mn2O3 electrode. As shown in Figure 8, the pseudosupercapacitor based on the synthesized mesoporous Mn2O3 electrode features a large phase angle near the low-frequency region, indicating the faradic capacitance behavior of the electrode. Importantly, in the EIS plot, the straight line in the low-frequency region, and the absence of any small semicircle in the high-frequency region, are indicative of the good capacitive nature of the present electrochemical system based on a mesoporous Mn2O3 electrode. Likewise, the observed straight line at a low frequency might reduce the diffusion length and accelerate the ions transportation on the mesoporous surface of the Mn2O3 electrode, as evidenced by the CV results.

4. Conclusions

In summary, a facial hydrothermal process was adopted to synthesize well-crystalline mesoporous Mn2O3 materials for the fabrication of a pseudocapacitor. The crystalline and structural characterizations confirmed the formation of Mn2O3 materials without displaying any other oxide forms. The surface properties were analyzed, showing the mesoporous nature of Mn2O3 materials and an estimated high surface area of 76.9 m2·g−1 with a good pore distribution. The fabricated pseudo-capacitor based on a Mn2O3 mesoporous particle electrode shows a reasonably high specific capacitance of ~460 F·g−1 at 10 mV·s−1 in a 6 M KOH aqueous solution. The enhancement in the capacitive properties might be attributed to the high surface area, porous surface and uniformity of the unique mesoporous particles morphology, resulting in the large generation of active sites and a fast ionic transport over the surface of the Mn2O3 electrode.

Author Contributions

Conceptualization, Y.-H.S., P.T.M.B., M.S.A., H.-R.L., D.K.S. and O-B.Y.; methodology, Y.-H.S., P.T.M.B., M.S.A., H.-R.L. and D.K.S.; software, Y.-H.S. and P.T.M.B.; validation, P.T.M.B., M.S.A., H.-R.L. and D.K.S.; formal analysis, Y.-H.S. and P.T.M.B.; investigation, Y.-H.S. and P.T.M.B.; resources, Y.-H.S. and P.T.M.B.; data curation, Y.-H.S., P.T.M.B., H.-R.L. and D.K.S.; writing—original draft preparation, Y.-H.S., P.T.M.B., and M.S.A.; writing—review and editing, Y.-H.S., P.T.M.B., M.S.A., and O-B.Y.; supervision, M.S.A., and O-B.Y.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (NRF-2019R1F1A1063999).

Acknowledgments

This paper was also acknowledged the support of research funds from Chonbuk National University in 2018.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miller, J.R.; Simon, P. Electrochemical capacitors for energy management. Science 2008, 321, 651–652. [Google Scholar] [CrossRef] [PubMed]
  2. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [Green Version]
  3. Wei, D.; Scherer, M.R.; Bower, C.; Andrew, P.; Ryhanen, T.; Steiner, U. A nanostructured electrochromic supercapacitor. Nano Lett. 2012, 12, 1857–1862. [Google Scholar] [CrossRef] [PubMed]
  4. Yuan, C.Z.; Wu, H.B.; Xie, Y.; Lou, X.W. Mixed transition-metal oxides: Design, synthesis, and energy-related applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, H.; Li, C.; Ma, J. Polyaniline–MnO2 coaxial nanofiber with hierarchical structure for high-performance supercapacitors. J. Mater. Chem. 2012, 22, 2751. [Google Scholar] [CrossRef]
  6. Devaraj, S.; Munichandraiah, N. High capacitance of electrodeposited MnO2 by the effect of a surface-active agent. Electrochem. Solid State Lett. 2005, 8, 373–377. [Google Scholar] [CrossRef]
  7. Zhao, D.; Tian, J.S.; Ji, Q.Q.; Zhang, J.T.; Zhao, X.S.; Guo, P.Z. Mn2O3 nanomaterials: Facile synthesis and electrochemical properties. Chin. J. Inorg. Chem. 2010, 26, 832–838. [Google Scholar]
  8. Jang, G.; Ameen, S.; Akhtar, M.S.; Kim, E.-B.; Shin, H.S. Electrochemical Investigations of Hydrothermally Synthesized Porous Cobalt Oxide (Co3O4) Nanorods: Supercapacitor Application. ChemistrySelect 2017, 2, 8941–8949. [Google Scholar] [CrossRef]
  9. Li, Z.Y.; Bui, P.T.M.; Kwak, D.H.; Akhtar, M.S.; Yang, O.B. Enhanced electrochemical activity of low temperature solution process synthesized Co3O4 nanoparticles for pseudo-supercapacitors applications. Ceram. Int. 2016, 42, 1879–1885. [Google Scholar] [CrossRef]
  10. Li, Z.Y.; Akhtar, M.S.; Yang, O.B. Supercapacitors with ultrahigh energy density based on mesoporous carbon nanofibers: Enhanced double-layer electrochemical properties. J. Alloy. Compd. 2015, 653, 212–218. [Google Scholar] [CrossRef]
  11. Seo, M.K.; Saouab, A.; Park, S.J. Effect of annealing temperature on electrochemical characteristics of ruthenium oxide/multi-walled carbon nanotube composites. Mater. Sci. Eng. B 2010, 167, 65–69. [Google Scholar] [CrossRef]
  12. Shambharkar, B.H.; Umare, S.S. Production and characterization of polyaniline/Co3O4 nanocomposite as a cathode of Zn–polyaniline battery. Mater. Sci. Eng. B 2010, 175, 120–128. [Google Scholar] [CrossRef]
  13. Zheng, Z.; Huang, L.; Zhou, Y.; Hu, X.W.; Ni, X.M. Large-scale synthesis of mesoporous CoO-doped NiO hexagonal nanoplatelets with improved electrochemical performance. Solid State Sci. 2009, 11, 1439–1443. [Google Scholar] [CrossRef]
  14. Prasad, K.R.; Miura, N. Electrochemical synthesis and characterization of nanostructured tin oxide for electrochemical redox supercapacitors. Electrochem. Commun. 2004, 6, 849–852. [Google Scholar] [CrossRef]
  15. Jiang, H.; Zhao, T.; Ma, J.; Yan, C.; Li, C. Ultrafine manganese dioxide nanowire network for high-performance supercapacitors. Chem. Commun. 2011, 47, 1264. [Google Scholar] [CrossRef] [PubMed]
  16. Rakhi, R.B.; Chen, W.; Cha, D.; Alshareef, H.N. Substrate dependent self-organization of mesoporous cobalt oxide nanowires with remarkable pseudocapacitance. Nano Lett. 2012, 12, 2559. [Google Scholar] [CrossRef] [PubMed]
  17. Lu, Q.; Lattanzi, M.W.; Chen, Y.; Kou, X.; Li, W.; Fan, X.; Unruh, K.M.; Chen, J.G.; Xiao, J.Q. Supercapacitor electrodes with high-energy and power densities prepared from monolithic NiO/Ni nanocomposites. Angew. Chem. 2011, 50, 6847–6850. [Google Scholar] [CrossRef]
  18. Benson, J.; Boukhalfa, S.; Magasinski, A.; Kvit, A.; Yushin, G. Chemical vapor deposition of aluminum nanowires on metal substrates for electrical energy storage applications. ACS Nano 2012, 6, 118–125. [Google Scholar] [CrossRef]
  19. Xie, K.; Li, J.; Lai, Y.; Lu, W.; Zhang, Z.; Liu, Y.; Zhou, L.; Huang, H. Highly ordered iron oxide nanotube arrays as electrodes for electrochemical energy storage. Electrochem. Commun. 2011, 13, 657–660. [Google Scholar] [CrossRef]
  20. Ghosh, S.; Polaki, S.R.; Sahoob, G.; Jin, E.-M.; Kamruddin, M.; Cho, J.S.; Jeong, S.M. Designing metal oxide-vertical graphene nanosheets structures for 2.6 V aqueous asymmetric electrochemical capacitor. J. Ind. Eng. Chem. 2019, 72, 107–116. [Google Scholar] [CrossRef]
  21. Nagamuthu, S.; Ryu, K.-S. MOF-derived microstructural interconnected network porous Mn2O3/C as negative electrode material for asymmetric supercapacitor device. CrystEngComm 2019, 21, 1442–1451. [Google Scholar] [CrossRef]
  22. Sui, N.; Duan, Y.; Jiao, X.; Chen, D. Large-scale preparation and catalytic properties of one-dimensional α/β-MnO2 nanostructures. J. Phys. Chem. C 2009, 113, 8560–8565. [Google Scholar] [CrossRef]
  23. Kolathodi, M.S.; Rao, S.N.H.; Natarajana, T.S.; Singh, G. Beaded manganese oxide (Mn2O3) nanofibers: Preparation and application for capacitive energy storage. J. Mater. Chem. A 2016, 4, 7883–7891. [Google Scholar] [CrossRef]
  24. Nakayama, M.; Tanaka, A.; Sato, Y.; Tonosaki, T.; Ogura, K. Electrodeposition of manganese and molybdenum mixed oxide thin films and their charge storage properties. Langmuir 2005, 21, 5907. [Google Scholar] [CrossRef]
  25. Wu, Z.-S.; Ren, W.; Wang, D.-W.; Li, F.; Liu, B. High-energy MnO2 nanowire/graphene and graphene asymmetric electrochemical capacitors. ACS Nano 2010, 4, 5835–5842. [Google Scholar] [CrossRef]
  26. 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]
  27. Ghodbane, O.; Pascal, J.-L.; Favier, F. Microstructural effects on charge-storage properties in MnO2-based electrochemical supercapacitors. ACS Appl. Mater. Interfaces 2009, 1, 1130–1139. [Google Scholar] [CrossRef]
  28. Zhi, M.; Xiang, C.; Li, J.; Li, M.; Wu, N. Nanostructured carbon–metal oxide composite electrodes for supercapacitors: A review. Nanoscale 2013, 5, 72–88. [Google Scholar] [CrossRef]
  29. Jiang, J.; Li, Y.; Liu, J.; Huang, X.; Yuan, C.; Lou, X.W. Recent advances in metal oxide-based electrode architecture design for electrochemical energy storage. Adv. Mater. 2012, 24, 5166–5180. [Google Scholar] [CrossRef]
  30. Han, Z.J.; Seo, D.H.; Yick, S.; Chen, J.H.; Ostrikov, K. MnOx/carbon nanotube/reduced graphene oxide nanohybrids as high-performance supercapacitor electrodes. NPG Asia Mater. 2014, 6, e140. [Google Scholar] [CrossRef]
  31. Demir, M.; Ashourirad, B.; Mugumya, J.H.; Saraswata, S.K.; El-Kaderi, H.M.; Gupta, R.B. Nitrogen and oxygen dual-doped porous carbons prepared from pea protein as electrode materials for high performance supercapacitors. Int. J. Hydrog. Energy 2018, 43, 18549–18558. [Google Scholar] [CrossRef]
  32. Cheng, F.; Shen, J.; Ji, W.; Tao, Z.; Chen, J. Selective synthesis of manganese oxide nanostructures for electrocatalytic oxygen reduction. ACS Appl. Mater. Interfaces 2009, 1, 460–466. [Google Scholar] [CrossRef] [PubMed]
  33. Jenkins, R. Introduction to X-Ray Powder Diffractometry; John Wiley & Sons, Inc.: New York, NY, USA, 1996; p. 90. [Google Scholar]
  34. Gillot, B.; El Guendouzi, M.; Laarj, M. Particle size effects on the oxidation–reduction behavior of Mn3O4 hausmannite. Mater. Chem. Phys. 2001, 70, 54–60. [Google Scholar] [CrossRef]
  35. Chen, Z.W.; Lai, J.K.L.; Shek, C.H. Influence of grain size on the vibrational properties in Mn2O3 nanocrystals. J. Non Cryst. Solids 2006, 352, 3285–3289. [Google Scholar] [CrossRef]
  36. Han, Y.-F.; Ramesh, K.; Chen, L.; Widjaja, E.; Chilukoti, S.; Chen, F. Controlled synthesis, characterization, and catalytic properties of Mn2O3 and Mn3O4 nanoparticles supported on mesoporous silica SBA-15. J. Phys. Chem. B 2007, 111, 2830–2833. [Google Scholar]
  37. Zhang, Y.; Yan, Y.; Wang, X.; Li, G.; Deng, D.; Jiang, L.; Shu, C.; Wang, C. Facile synthesis of porous Mn2O3 nanoplates and their electrochemical behavior as anode materials for lithium ion batteries. Chem. A Eur. J. 2014, 20, 6126–6130. [Google Scholar] [CrossRef]
  38. Jahan, M.; Tominaka, S.; Henzie, J. Phase pure α-Mn2O3 prisms and their bifunctional electrocatalytic activity in oxygen evolution and reduction reactions. Dalton Trans. 2016, 45, 18494–18501. [Google Scholar] [CrossRef]
  39. Ilton, E.S.; Post, J.E.; Heaney, P.J.; Ling, F.T.; Kerisit, S.N. XPS determination of Mn oxidation states in Mn (hydr)oxides. Appl. Surf. Sci. 2016, 366, 475–485. [Google Scholar] [CrossRef] [Green Version]
  40. Toupin, M.; Brousse, T.; Bélanger, D. Influence of microstucture on the charge storage properties of chemically synthesized manganese dioxide. Chem. Mater. 2002, 14, 3946–3952. [Google Scholar] [CrossRef]
  41. Toupin, M.; Brousse, T.; Bélanger, D. Charge storage mechanism of MnO2 electrode used in aqueous electrochemical capacitor. Chem. Mater. 2004, 16, 3184–3190. [Google Scholar] [CrossRef]
  42. Rajkumar, M.; Hsu, C.-T.; Wu, T.-H.; Chen, M.-G.; Hu, C.-C. Advanced materials for aqueous supercapacitors in the asymmetric design. Prog. Nat. Sci. Mater. Int. 2015, 25, 527. [Google Scholar] [CrossRef]
  43. Gopalakrishnan, I.K.; Bagkar, N.; Ganguly, R.; Kulshreshtha, S.K. Synthesis of superparamagnetic Mn3O4 nanocrystallites by ultrasonic irradiation. J. Cryst. Growth 2005, 280, 436. [Google Scholar] [CrossRef]
  44. Li, Z.-Y.; Akhtar, M.S.; Bui, P.T.M.; Yang, O.-B. Predominance of two dimensional (2D) Mn2O3 nanowalls thin film for high performance electrochemical supercapacitors. Chem. Eng. J. 2017, 330, 1240–1247. [Google Scholar] [CrossRef]
Figure 1. The (a,b) FESEM, (c) TEM and (d) HRTEM images of the synthesized mesoporous Mn2O3 materials.
Figure 1. The (a,b) FESEM, (c) TEM and (d) HRTEM images of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g001
Figure 2. The XRD patterns of the synthesized mesoporous Mn2O3 materials.
Figure 2. The XRD patterns of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g002
Figure 3. The (a) Fourier-transform infrared spectroscopy (FTIR) and (b) Raman spectrum of the synthesized mesoporous Mn2O3 materials.
Figure 3. The (a) Fourier-transform infrared spectroscopy (FTIR) and (b) Raman spectrum of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g003
Figure 4. (a,b) TGA plot of the synthesized mesoporous Mn2O3 materials.
Figure 4. (a,b) TGA plot of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g004
Figure 5. (a) Survey, (b) high resolution Mn 2p, (c) Mn 3s, and (d) O 1s XPS of the synthesized mesoporous Mn2O3 materials.
Figure 5. (a) Survey, (b) high resolution Mn 2p, (c) Mn 3s, and (d) O 1s XPS of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g005
Figure 6. The (a) N2 adsorption-desorption isotherm and (b) pore size distribution plot of the synthesized mesoporous Mn2O3 materials.
Figure 6. The (a) N2 adsorption-desorption isotherm and (b) pore size distribution plot of the synthesized mesoporous Mn2O3 materials.
Coatings 09 00631 g006
Figure 7. The (a) CV curves and (b) specific capacitance of the synthesized mesoporous Mn2O3 electrode at different scan rates ranging from 10 to 500 mV·s−1, (c) multicycles CV curves and (d) variation in the specific capacitance of the synthesized mesoporous Mn2O3 electrode after 5000 cycles. The inset shows the FESEM image of the synthesized mesoporous Mn2O3 electrode after 5000 cycles.
Figure 7. The (a) CV curves and (b) specific capacitance of the synthesized mesoporous Mn2O3 electrode at different scan rates ranging from 10 to 500 mV·s−1, (c) multicycles CV curves and (d) variation in the specific capacitance of the synthesized mesoporous Mn2O3 electrode after 5000 cycles. The inset shows the FESEM image of the synthesized mesoporous Mn2O3 electrode after 5000 cycles.
Coatings 09 00631 g007
Figure 8. The EIS plot of the synthesized mesoporous Mn2O3 electrode at the Zim versus the Zre mode. The inset shows its corresponding equivalent circuit.
Figure 8. The EIS plot of the synthesized mesoporous Mn2O3 electrode at the Zim versus the Zre mode. The inset shows its corresponding equivalent circuit.
Coatings 09 00631 g008

Share and Cite

MDPI and ACS Style

Son, Y.-H.; Bui, P.T.M.; Lee, H.-R.; Akhtar, M.S.; Shah, D.K.; Yang, O.-B. A Rapid Synthesis of Mesoporous Mn2O3 Nanoparticles for Supercapacitor Applications. Coatings 2019, 9, 631. https://doi.org/10.3390/coatings9100631

AMA Style

Son Y-H, Bui PTM, Lee H-R, Akhtar MS, Shah DK, Yang O-B. A Rapid Synthesis of Mesoporous Mn2O3 Nanoparticles for Supercapacitor Applications. Coatings. 2019; 9(10):631. https://doi.org/10.3390/coatings9100631

Chicago/Turabian Style

Son, You-Hyun, Phuong T. M. Bui, Ha-Ryeon Lee, Mohammad Shaheer Akhtar, Deb Kumar Shah, and O-Bong Yang. 2019. "A Rapid Synthesis of Mesoporous Mn2O3 Nanoparticles for Supercapacitor Applications" Coatings 9, no. 10: 631. https://doi.org/10.3390/coatings9100631

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