Next Article in Journal
Alginate-Based Zinc Oxide Nanoparticles Coating Extends Storage Life and Maintains Quality Parameters of Mango Fruits “cv. Kiett”
Next Article in Special Issue
Influence of Synthetic Conditions on the Crystal Structure, Optical and Magnetic Properties of o-EuFeO3 Nanoparticles
Previous Article in Journal
Calcium Phosphate Loaded Biopolymer Composites—A Comprehensive Review on the Most Recent Progress and Promising Trends
Previous Article in Special Issue
Thin Films on the Surface of GaAs, Obtained by Chemically Stimulated Thermal Oxidation, as Materials for Gas Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Novel Oxidation–Reduction Route for the Morphology-Controlled Synthesis of Manganese Oxide Nanocoating as Highly Effective Material for Pseudocapacitors

Ioffe Institute, 194021 Saint-Petersburg, Russia
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(2), 361; https://doi.org/10.3390/coatings13020361
Submission received: 19 December 2022 / Revised: 21 January 2023 / Accepted: 25 January 2023 / Published: 5 February 2023

Abstract

:
In recent years, pseudocapacitors have been receiving much attention as low-cost and safe energy storage technology for emerging applications in flexible and safe devices. However, creating high-energy-density electrode materials is now the main limit for high-performance pseudocapacitors. In this work, we propose a novel reduction route for the synthesis of uniform MnO2 nanocoating with porous morphology on nickel foam via the SILD method as electrode material for high-effective pseudocapacitors. The obtained nanocoatings were characterized by SEM, TEM, EDX, XRD, XPS, and electrochemical techniques. Comparisons of MnO2 coatings were conducted to obtain the reduction and oxidative routes of synthesis. The influence of the oxidation–reduction reaction type on the structures, morphologies, and capacity performance of manganese oxide was investigated. The results show that the nanocoatings synthesized via the reduction route were formed of amorphous uniform ultra-thick coating MnO2 with a porous morphology of “nanoflakes.” Due to the unique morphology and uniform coating of nanosized manganese oxide, electrodes based on this process have shown a high specific capacity (1490 F/g at 1 A/g) and excellent cycling stability (97% capacity retention after 1000 charge–discharge cycles).

1. Introduction

In recent years, the need for the development and creation of new efficient energy sources, which determine the prospects for the development of industrial plants, electric transport, space, aircraft, etc., has increased. Such power sources are subject to a number of requirements: high values of specific energy and capacity, stability of characteristics at multiple charge–discharge cycles, long service life, high operating voltage, and power. In addition, these devices must be compact, environmentally sustainable, environmentally friendly, and inexpensive to produce [1,2].
Currently, among traditional power sources, the most widely used are lithium-ion batteries whose high values of specific energy (80–200 W·h/kg) and power allow for widespread practical use [3]. However, lithium-ion batteries have many disadvantages, namely the high cost of lithium raw materials caused by the constant increase in the production rates of such power sources and the problem of ensuring their safe disposal, a relatively large mass, and a long charge time (which is critical for their use in electric vehicles and portable devices), as well as the problems of overheating and the risk of explosion [4].
Consequently, there has been a growing interest in the development of alternative types of power sources, such as supercapacitors, known for their widespread use and low cost, as well as their increased operational safety compared to lithium-ion batteries. Of particular interest are hybrid batteries, which combine two mechanisms of energy storage, namely a double electric layer at the electrode–electrolyte interface and charge transfer in the electrolyte. The pseudocapacitor energy storage devices based on capacity and supercapacitor electrodes offer a promising way to construct devices with the merits of both secondary batteries and supercapacitors [5,6,7]. As electrodes for pseudocapacitors, carbon materials or electrodes based on transition metal oxides (Ni, Co, Mn, and Fe-based) are employed as pseudocapacitor electrodes, while aqueous alkaline solutions are used as electrolytes [8,9]. Thus, the synthesis of electroactive nanomaterials and nanocoatings based on transition metal oxides is the key point for the development of electrode materials for highly efficient energy storage devices in partially flexible energy storage devices [10,11,12].
From this standpoint, manganese oxides are ideal candidates for pseudocapacitive electrode materials owing to their multiple valence states enabling rich redox reactions as well as their excellent resistance in alkaline media [13,14,15,16]. Additionally, manganese oxides are more abundant and environmentally friendly than other transition metal oxides. However, the achievement of high specific capacitance, power density, and cyclic stability is limited by the structure and morphology of oxides, which play a key role in ensuring high electronic conductivity and a large number of active centers on the surface, as well as improving ion diffusion at the electrode–electrolyte interface. The solution to this problem may consist of obtaining ultrathin nanocoatings with a two-dimensional graphene-like morphology, providing a set of unique physical and chemical properties [13]. Currently, the main problem in the development of electrode materials for pseudocapacitors is the complexity of obtaining uniform ultrathin nanofilms and nanocoatings of oxides using existing methods of synthesis in a way that would significantly reduce the degradation of the electroactive material and close to the values of a theoretical specific capacity [17].
In this respect, the SILD (Successive Ionic Layer Deposition) [18] or SILAR (Successive Ionic Layer Adsorption and Reaction) methods can be prospectively applied [19]. The SILD method, compared with other popular coating techniques, has unique traits and advantages. For example, it allows for the deposition of coatings on various surfaces without dimensional restrictions and for their thickness to be readily modified across a wide range by controlling the concentration of the precursor, the number of depositions, and so on [19]. Furthermore, using various synthetic routes, the deposited coatings can be tuned to exhibit preferential morphologies and crystallographic orientations. The ability to obtain coatings on the substrate surface without the use of additional binding materials is key for the synthesis of highly efficient electrode materials. Previously, we have successfully obtained electroactive nanocoatings based on transition metal oxides and hydroxides with different structures and morphologies using the SILD technique [20,21,22,23,24].
Currently, most studies use an oxidative synthesis route to obtain manganese oxide coatings via the SILAR method [25]. MnO2 is synthesized using two main methods, namely Mn2+ oxidation and a redox reaction of Mn2+ and MnO4−. In the first method, MnO2 is generated by treatment oxidants, such as H2O2, K2S2O8, NaClO, etc. In the second case, MnO2 can also be prepared by the reaction of MnO4 with Mn2+ under alkaline conditions. However, these methods are characterized by a long reaction time and require a specifically controlled pH, complicating the reaction conditions as well as introducing alkali metal impurities to the film. Previously, the effectiveness of the use of a reduction synthesis route for the production of manganese oxide was shown. For example, paper [26] describes the production of MnO2 nanoflowers by reducing potassium permanganate with hydrogen peroxide in an acidic medium. In the process of such a synthesis, peroxide is decomposed into non-toxic H2O and O2. The in situ evolution of oxygen gas bubbles also plays an important role in the formation of the unique morphology. However, previously, such a synthesis route has not been used to obtain electroactive MnO2-based nanocoatings via the SILD (SILAR) method.
In this work, we compare the oxidation and reduction routes for the synthesis of manganese oxide ultrathin nanocoatings via the SILD method. Manganese oxide nanocoatings were deposited from aqueous solutions of potassium permanganate under “soft conditions” via oxidation and reduction synthesis routes. For the first time, we used the reduction route of the SILD method for the synthesis of MnO2 nanocoatings and explored their morphology, structure, and electrochemical properties as pseudocapacitor electrodes.

2. Materials and Methods

2.1. Materials

The nanolayers of manganese oxide were fabricated using high-purity analytical grade reagents: manganese acetate tetrahydrate (Mn(OAc)2·4H2O), potassium permanganate (KMnO4), hydrogen peroxide (H2O2), hydrochloric acid (HCl), and potassium hydroxide (KOH). Nickel foam (NF) (110 PPI; size 20 × 30 mm) was used as a substrate.

2.2. Synthesis of the Nanolayers of Manganese Oxide via SILD Synthesis

In the first stage, the NF substrate was cleaned in ethanol and distilled water by sonication for 10 min to remove the impurities and then treated with a 3 M HCl solution for 10 min to remove the oxidation layer. The manganese nanolayers were fabricated by the SILD technique.
In the first case, aqueous solutions of salts Mn(OAc)2 (0.01 M) were used as a precursor for the synthesis, and an aqueous solution of KMnO4 (0.01 M) (MO1) was used as an oxidizer. In the second case, aqueous solutions of KMnO4 (0.01 M) were used as precursors for the synthesis, and an aqueous solution of H2O2 (3%) was used as a reductant (MO2). The precleaned NF substrates were first dipped in the solutions of the first reagent for 30 s and then washed from excess reagent in distilled water for 15 s. In the following step, the substrates were dipped into a solution of oxidizer or reductant for 30 s and rewashed in water (Figure 1). After repeating 50 SILD cycles, an insoluble coating of the synthesized product was formed on the surface. Finally, the coated films were annealed at 150 °C under an argon atmosphere for 3 h.

2.3. Characterization

The morphology and structure of the synthesized films were investigated by using scanning electron microscopy SEM (Zeiss Merlin microscope, Zeiss, Jena, Germany), transmission electron microscopy (TEM) (Zeiss Libra 200 microscope, Zeiss, Jena, Germany), and X-ray diffraction analysis (XRD) (Rigaku Smartlab 3, Rigaku, Tokyo, Japan, diffractometer with CuKα radiation). Qualitative and quantitative analysis of the samples was performed using the XPS (ESCALAB 250Xi electron spectrometer, ThermoFisher Scientific, Waltham, MA, USA) and EDX (Oxford INCA energy 350, Oxford Instruments, Abingdon, UK) methods.

2.4. Electrochemical Measurement

The electrochemical experiment was carried out using an Elins P45-X potentiostat with a three-electrode electrochemical cell in 1 M of KOH aqueous electrolyte solution. The working electrode was obtained through the SILD deposition of electroactive nanocoatings on an NF. An Ag/AgCl electrode and platinum foil were used as the reference and counter electrodes, respectively. The electrochemical characterization of the electrode was carried out using cyclic voltammetry (CVA), galvanostatic charge–discharge (GCD), and the EIS techniques. The specific capacitance C (F/g) at different current densities can be calculated via Equation (1):
C = Idt dVm
where I (mA) is the galvanostatic current, dV (mv) is the potential window, dt (h) is the discharge time of a cycle, and m (g) is the mass of the active material in the film electrode. The mass of the electroactive material of the working electrode was measured using a microbalance. The electrochemical impedance spectra (EIS) were measured at a voltage amplitude of 5 mV in a frequency range between 10 kHz and 0.1 Hz. The electroactive mass of manganese oxide (2.4 mg) for the working electrode was measured using a JOANLAB FA1204 microbalance.

3. Results and Discussion

Morphology and crystal structure were characterized via SEM, TEM, and XRD techniques. As can be seen (Figure 2a,b), the synthesized nanocoating of MO1 formed an agglomerate flower-like shape consisting of ultrathin (thickness 5–8 nm) “nanosheets” with porous morphologies. By contrast, the micrograph of the MO2 (Figure 2c,d) shows that the surface of the nickel foam is uniformly covered with a homogeneous ultrathin film of the synthesized compound with a “nanoflakes” morphology, whose thickness and dimensions are smaller than those of the MO1. As it follows from the obtained TEM images, both coatings representing ultrathin, two-dimensional formations have an amorphous structure without distinctive lattice fringes.
The XRD patterns (Figure 3a) of the MO1 and MO2 samples show broad featureless peaks close to the crystal structure of birnessite MnO2 (amcsd #0004947). The obtained data may indicate a low crystallinity and an amorphous nature, which is consistent with the data obtained from TEM. The poor diffraction peaks observed are a distinctive feature of birnessite (MnO2) nanocrystals with a nanosheets/nanoflakes morphology, as can be seen in a number of recent works [9,10,11,12]. The reasons for the weak peaks are the synthesis at room temperature and subsequent moderate heat treatment (150 °C).
The XPS analysis verified that the chemical compounds and oxidation states of the elements in the as-synthesized nanocoatings were identical for both samples. The two peaks (Figure 4a) at the binding energies of 642.1 eV and 653.8 eV can be assigned to Mn 2p3/2 and Mn 2p1/2, respectively, and can also be attributed to the Mn4+ oxidation state [27]. The peak of O 1s (Figure 4b) at 529.4 eV can correspond to metal–oxygen bonding in the oxide and the M–O bonding of the metal oxide, while the wide peak at 531.8 eV can correspond to OH bonding in the adsorbed H2O [27]. The EDX quantitative and qualitative analysis of the MO1 and MO2 samples demonstrated the presence of Mn, O, K (no more than 1%), and Ni (from the substrate) atoms in the synthesized nanocoatings, which agrees well with the XPS data.
Based on the analysis of the results obtained via SEM, TEM, XRD, EDX, and XPS, we can assume that the obtained nanocoating is formed by MnO2·nH2O with a different morphology and amorphous structure. Comparing the morphology of manganese oxides obtained by oxidation and reduction, a significant difference can be observed, which confirms that the synthetic method plays an important role in controlling the morphology of nanocoatings.
The formation of synthesized nanocoatings can be presented in two schemes. In the first step of the oxidation route, Mnaq2+ cations are adsorbed on the surface of a substrate, and then the excess reagent is removed by washing it with water. In the second step, when the substrate is immersed in an oxidizer solution, the Mn2+ cations are oxidized, likely only partially, to Mn4+, resulting in a redox reaction with KMnO4. As a result, an insoluble amorphous oxide is formed. Another route of synthesis is the reduction of potassium permanganate by hydrogen peroxide in a neutral medium. The MnO4 anions are adsorbed on the substrate of nickel foam, and when immersed in a solution of H2O2, they are reduced to MnO2.
The electrochemical properties of the NF electrode based on manganese oxide nanocoatings were characterized using the CVA, GCD, and EIS methods. Figure 5a shows a comparison of the CVA curves of the MO1 and MO2 electrodes. As can be seen from the figure, the area of the MO2 CVA curve is larger than the area of the MO1 CVA curve, which indicates a greater specific capacity of this electrode. Further study was focused on the MO2 obtained using the reduction synthesis route. Figure 5b presents the CVA curves of the MO2 electrodes at different scan rates (5, 10, and 20 mV/s) within a potential window from 0.2 V to 0.55 V (vs. Ag/AgCl). On the CVA curves, one pair of wide redox peaks is observed over the entire range, suggesting a pseudocapacitive behavior of this material. Figure 5c shows the galvanostatic charge/discharge curves of the MO2 electrode at different current densities (1, 2, and 5 A/g), indicating a typical pseudocapacitance, which likely originates from the fast redox reaction of Mn3+ → Mn4+ and Mn4+ → Mn3+ on the electrode/electrolyte interface. The specific capacitance of the MO2 electrode, calculated by Equation (1), is 1490 F/g at a current density of 1 A/g.
Figure 5d shows the cycling stability of the MO2 electrode at a current density of 5 A/g in 1 M KOH electrolyte. A small increase in the capacity is observed during the first 150 charge–discharge cycles, after which it begins to decrease. After 1000 charge–discharge cycles, the initial capacitance decreased by only 3%, suggesting excellent electrochemical stability. XRD was performed for the MO2 electrodes before (as-synthesized sample) and after long-term analysis (sample after cyclic tests). The results are presented in Figure 3b. No noticeable changes were found in the X-ray diffraction patterns, indicating the high stability of the obtained MnO2-based electrode material even after 1000 charge–discharge cycles. The high electrochemical performance of the sample obtained via the reduction route in comparison to the one obtained by the oxidation route can be explained by the unique porous “nanoflakes” morphology and uniform coating of MnO2. This type of morphology ensures high electric conductivity and a large number of active sites, as well as improved ion diffusion on the electrode surface. The amorphous nature of manganese oxide is likely to benefit ionic charge transport.
The Nyquist plot of the MO1 and MO2 electrodes (Figure 6) displays a linear part in the low-frequency region and an absence of a semicircle in the high-frequency region, indicating a very low transfer resistance and fast electrochemical reactions. The electrode resistance was calculated from the Nyquist plot and is equal to 1.8 Ω for MO1 and 1.5 MO2, which proves low resistance and, as a consequence, high capacity for MO2.
The results indicate that the nanocoating of the MnO2-synthesized reduction route via the SILD method has excellent potential for application as electroactive materials in pseudocapacitors. Thus, the proposed route of synthesis for uniform nanocoatings using the SILD method allowed for the obtainment of the synthesized materials on a substrate with an intricate morphology, and, in the future, they could be used in flexible energy storage devices. In our opinion, this route opens new possibilities for the fabrication of high-performance electroactive nanomaterials. The synthesis of the nanostructures of other TMOs via the SILD method will be the focus of forthcoming publications.

4. Conclusions

In this study, a MnO2-based electroactive nanocoating was successfully deposited on nickel foam through reduction routes via the SILD technique. The influence of oxidation–reduction reaction type on the structures, morphologies, and capacity performance of manganese oxide was investigated. The manganese oxide synthesized via the reduction route of the SILD formed an ultrathin uniform nanocoating with an amorphous structure and showed good capacity performance. These electrodes have shown a high specific capacity (1490 F/g at 1 A/g) and excellent cycling stability (97% capacity retention after 1000 charge–discharge cycles) as pseudocapacitors in alkaline media. Therefore, this route opens new possibilities for the fabrication of high-performance MnO2 nanocoatings, which are highly promising as capacitive materials for flexible energy storage devices in the future.

Author Contributions

Conceptualization and methodology, A.A.L. and I.A.K.; investigation, M.I.T. and I.A.K.; writing—original draft preparation, A.A.L.; writing—review and editing, A.A.L.; visualization, M.I.T.; supervision and project administration, V.I.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant from the President of the Russian Federation for Young Doctors and Candidates of Sciences, grant number MK-3864.2022.1.3.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the Centers for Physical Methods of Surface Investigation and the Centers for Nanotechnology of Saint Petersburg State University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, C.; Li, F.; Ma, L.-P.; Cheng, H.-M. Advanced Materials for Energy Storage. Adv. Mater. 2010, 22, E28–E62. [Google Scholar] [CrossRef]
  2. Zhai, Y.; Dou, Y.; Zhao, D.; Fulvio, P.F.; Mayes, R.; Dai, S. Carbon materials for chemical capacitive energy storage. Adv. Mater. 2011, 23, 4828–4850. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, G.; Li, F.; Cheng, H.-M. Progress in flexible lithium batteries and future prospects. Energy Environ. Sci. 2014, 7, 1307–1338. [Google Scholar] [CrossRef]
  4. Lee, S.-Y.; Choi, K.-H.; Choi, W.-S.; Kwon, Y.H.; Jung, H.-R.; Shin, H.-C.; Kim, J.Y. Progress in flexible energy storage and conversion systems, with a focus on cable-type lithium-ion batteries. Energy Environ. Sci. 2013, 6, 2414. [Google Scholar] [CrossRef]
  5. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [PubMed]
  6. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A review of electrolyte materials and compositions for electrochemical supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef] [PubMed]
  7. Bhojane, P. Recent advances and fundamentals of Pseudocapacitors: Materials, mechanism, and its understanding. J. Energy Storage 2022, 45, 103654. [Google Scholar] [CrossRef]
  8. Abdah, M.A.A.M.; Azman, N.H.N.; Kulandaivalu, S.; Sulaiman, Y. Review of the use of transition-metal-oxide and conducting polymer-based fibers for high-performance supercapacitors. Mater. Design 2020, 186, 108199. [Google Scholar] [CrossRef]
  9. Palem, R.; Ramesh, S.; Bathula, C.; Kakani, V.; Saratale, G.; Yadav, H.; Kim, J.-H.; Kim, H.; Lee, S.-H. Enhanced supercapacitive behavior by CuO@MnO2/carboxymethyl cellulose composites. Ceram. Int. 2021, 47, 26738–26747. [Google Scholar] [CrossRef]
  10. Wang, Y.; Wu, X.; Han, Y.; Li, T. Flexible supercapacitor: Overview and outlooks. J. Energy Storage 2021, 42, 103053. [Google Scholar] [CrossRef]
  11. Zan, G.; Wu, T.; Zhu, F.; He, P.; Cheng, Y.; Chai, S.; Wang, Y.; Huang, X.; Zhang, W.; Wan, Y.; et al. A biomimetic conductive super-foldable material. Matter 2021, 4, 3232–3247. [Google Scholar] [CrossRef]
  12. Zan, G.; Wu, T.; Dong, W.; Zhou, J.; Tu, T.; Xu, R.; Chen, Y.; Wang, Y.; Wu, Q. Two-Level Biomimetic Designs Enable Intelligent Stress Dispersion for Super-Foldable C/NiS Nanofiber Free-Standing Electrode. Adv. Fiber Mater. 2022, 4, 1117–1190. [Google Scholar] [CrossRef]
  13. Julien, C.M.; Mauger, A. Nanostructured MnO2 as Electrode Materials for Energy Storage. Nanomaterials 2017, 7, 396. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, X.-Y.; Han, L.-Q.; Sun, S.; Wang, C.-Y.; Chen, M.-M. MnO2/C composite electrodes free of conductive enhancer for supercapacitors. J. Alloys Compd. 2015, 653, 539–545. [Google Scholar] [CrossRef]
  15. Huang, M.; Li, F.; Dong, F.; Zhang, Y.X.; Zhang, L.L. MnO2-based nanostructures for high-performance supercapacitors. J. Mater. Chem. A 2015, 3, 21380–23423. [Google Scholar] [CrossRef]
  16. Yin, B.; Zhang, S.; Jiang, H.; Qu, F.; Wu, X. Phase-controlled synthesis of polymorphic MnO2 structures for electrochemical energy storage. J. Mater. Chem. 2015, 3, 5722–5729. [Google Scholar] [CrossRef]
  17. Sun, P.; Yi, H.; Peng, T.; Jing, Y.; Wang, R.; Wang, H.; Wang, X. Ultrathin MnO2 nanoflakes deposited on carbon nanotube networks for symmetrical supercapacitors with enhanced performance. J. Power Sources 2017, 341, 27–35. [Google Scholar] [CrossRef]
  18. Tolstoy, V.P. Successive ionic layer deposition. An application in nanotechnology. Russ. Chem. Rev. 2006, 75, 161. [Google Scholar] [CrossRef]
  19. Ratnayake, S.P.; Ren, J.; Colusso, E.; Guglielmi, M.; Martucci, A.; Della Gaspera, E. SILAR Deposition of Metal Oxide Nanostructured Films. Small 2021, 17, 2101666. [Google Scholar] [CrossRef]
  20. Kodintsev, I.A.; Martinson, K.D.; Lobinsky, A.A.; Popkov, V.I. Successive ionic layer deposition of co-doped Cu(OH)2 nanorods as electrode material for electrocatalytic reforming of ethanol. Nanosyst. Phys. Chem. Math. 2019, 10, 573–578. [Google Scholar] [CrossRef]
  21. Kodintsev, I.A.; Martinson, K.D.; Lobinsky, A.A.; Popkov, V.I. SILD synthesis of the efficient and stable electrocatalyst based on CoO-NiO solid solution toward hydrogen production. Nanosyst. Phys. Chem. Math. 2019, 10, 681–685. [Google Scholar] [CrossRef]
  22. Lobinsky, A.A.; Popkov, V.I. Direct SILD synthesis of efficient electroactive materials based on ultrathin nanosheets of amorphous CoCr-LDH. Mater. Lett. 2022, 322, 132472. [Google Scholar] [CrossRef]
  23. Lobinsky, A.A.; Kaneva, M.V. Synthesis Ni-doped CuO nanorods via successive ionic layer deposition method and their capacitive performance. Nanosyst. Phys. Chem. Math. 2020, 11, 608–614. [Google Scholar] [CrossRef]
  24. Lobinsky, A.A.; Tolstoy, V.P. Synthesis of 2D Zn–Co LDH nanosheets by a successive ionic layer deposition method as a material for electrodes of high-performance alkaline battery–supercapacitor hybrid devices. RSC Adv. 2018, 8, 29607–29612. [Google Scholar] [CrossRef]
  25. Soonmin, H. Recent advances in the growth and characterizations of SILAR-deposited thin films. Appl. Sci. 2022, 12, 8184. [Google Scholar] [CrossRef]
  26. Li, Y.; Jiang, G.; Ouyang, N.; Qin, Z.; Lan, S.; Zhang, Q. The controlled synthesis of birnessite nanoflowers via H2O2 reducing KMnO4 for efficient adsorption and photooxidation activity. Front. Chem. 2021, 9, 699513. [Google Scholar] [CrossRef] [PubMed]
  27. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; St, R.; Smart, C. Resolving surface chemical states in XPS analysis of first-row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
Figure 1. Scheme of oxidation (a) and reduction (b) route of manganese oxide via the SILD method.
Figure 1. Scheme of oxidation (a) and reduction (b) route of manganese oxide via the SILD method.
Coatings 13 00361 g001
Figure 2. SEM and TEM images of MO1 (a,b) and MO2 (c,d), respectively.
Figure 2. SEM and TEM images of MO1 (a,b) and MO2 (c,d), respectively.
Coatings 13 00361 g002
Figure 3. XRD patterns of MO1 and MO2 (a), XRD patterns of MO2 before (1st cycle) and after (1000st cycle) cycling charge–discharge testing (b).
Figure 3. XRD patterns of MO1 and MO2 (a), XRD patterns of MO2 before (1st cycle) and after (1000st cycle) cycling charge–discharge testing (b).
Coatings 13 00361 g003
Figure 4. XPS spectra Mn 2p (a) and O1s (b) of manganese oxide nanolayers.
Figure 4. XPS spectra Mn 2p (a) and O1s (b) of manganese oxide nanolayers.
Coatings 13 00361 g004
Figure 5. CVA curves of MO1 and MO2 (a); CVA curves at several scan rates (b) of MO2; GCD at several current density curves (c) of MO2; cycling stability of MO2 (at 5 A/g) (d).
Figure 5. CVA curves of MO1 and MO2 (a); CVA curves at several scan rates (b) of MO2; GCD at several current density curves (c) of MO2; cycling stability of MO2 (at 5 A/g) (d).
Coatings 13 00361 g005
Figure 6. Electrochemical impedance spectra (EIS) of MO1 and MO2.
Figure 6. Electrochemical impedance spectra (EIS) of MO1 and MO2.
Coatings 13 00361 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lobinsky, A.A.; Kodintzev, I.A.; Tenevich, M.I.; Popkov, V.I. A Novel Oxidation–Reduction Route for the Morphology-Controlled Synthesis of Manganese Oxide Nanocoating as Highly Effective Material for Pseudocapacitors. Coatings 2023, 13, 361. https://doi.org/10.3390/coatings13020361

AMA Style

Lobinsky AA, Kodintzev IA, Tenevich MI, Popkov VI. A Novel Oxidation–Reduction Route for the Morphology-Controlled Synthesis of Manganese Oxide Nanocoating as Highly Effective Material for Pseudocapacitors. Coatings. 2023; 13(2):361. https://doi.org/10.3390/coatings13020361

Chicago/Turabian Style

Lobinsky, Artem A., Ilya A. Kodintzev, Maxim I. Tenevich, and Vadim I. Popkov. 2023. "A Novel Oxidation–Reduction Route for the Morphology-Controlled Synthesis of Manganese Oxide Nanocoating as Highly Effective Material for Pseudocapacitors" Coatings 13, no. 2: 361. https://doi.org/10.3390/coatings13020361

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