Next Article in Journal
Presence of Induced Weak Ferromagnetism in Fe-Substituted YFexCr1−xO3 Crystalline Compounds
Next Article in Special Issue
Graphene Quantum Dots: Novel Properties and Their Applications for Energy Storage Devices
Previous Article in Journal
Ultra-Broadband, Omnidirectional, High-Efficiency Metamaterial Absorber for Capturing Solar Energy
Previous Article in Special Issue
Easy Diameter Tuning of Silicon Nanowires with Low-Cost SnO2-Catalyzed Growth for Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pseudocapacitive Effects of Multi-Walled Carbon Nanotubes-Functionalised Spinel Copper Manganese Oxide

Sensor Laboratories (SensorLab), Department of Chemistry, University of the Western Cape, Robert Sobukwe Road, Bellville, Cape Town 7535, South Africa
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3514; https://doi.org/10.3390/nano12193514
Submission received: 1 July 2022 / Revised: 26 September 2022 / Accepted: 29 September 2022 / Published: 8 October 2022
(This article belongs to the Special Issue Nanomaterials for Catalysis and Energy Storage)

Abstract

:
Spinel copper manganese oxide nanoparticles combined with acid-treated multi-walled carbon nanotubes (CuMn2O4/MWCNTs) were used in the development of electrodes for pseudocapacitor applications. The CuMn2O4/MWCNTs preparation involved initial synthesis of Mn3O4 and CuMn2O4 precursors followed by an energy efficient reflux growth method for the CuMn2O4/MWCNTs. The CuMn2O4/MWCNTs in a three-electrode cell assembly and in 3 M LiOH aqueous electrolyte exhibited a specific capacitance of 1652.91 F g−1 at 0.5 A g−1 current load. Similar investigation in 3 M KOH aqueous electrolyte delivered a specific capacitance of 653.41 F g−1 at 0.5 A g−1 current load. Stability studies showed that after 6000 cycles, the CuMn2O4/MWCNTs electrode exhibited a higher capacitance retention (88%) in LiOH than in KOH (64%). The higher capacitance retention and cycling stability with a Coulombic efficiency of 99.6% observed in the LiOH is an indication of a better charge storage behaviour in this electrolyte than in the KOH electrolyte with a Coulombic efficiency of 97.3%. This superior performance in the LiOH electrolyte than in the KOH electrolyte is attributed to an intercalation/de-intercalation mechanism which occurs more easily in the LiOH electrolyte than in the KOH electrolyte.

1. Introduction

The rapid increase in energy demand and consumption of fossil fuel resources has evoked the need for developing renewable and self-sustaining energy generation systems, such as solar cells, wind and hydroelectric turbines [1]. However, due to the intermittent nature of solar and wind energy, including the expensive construction cost of hydropower plants, high performance energy storage technologies are in great demand [2]. In this regard, supercapacitors possess unique advantages, such as high specific power, extended cycle life, fast charge/discharge rates, light weight and environmental benignity [3]. However, the major limitation of supercapacitors for practical applications is their low specific energy. This therefore provides motive to develop supercapacitor electrode materials capable of generating enhanced energy storage capacities without jeopardizing their high specific power and cycling stability.
Supercapacitors are classified as electric double-layer capacitors (EDLCs) and pseudocapacitors (PCs) according to their specific charge storage mechanism. EDLCs involve the use of carbon-based materials such as activated carbon, carbon nanotubes and grapheme as electrode materials to store charges by electrostatic charge adsorption and desorption processes, whereas PCs use metal oxide/hydroxide and conducting polymer electrode materials to store charges via fast and reversible faradaic redox reactions throughout the material surface [4]. Recently, numerous transition metal oxides (TMOs) such as CuO, MnO2, Mn3O4, RuO2, ZnO, NiO and Co3O4 have been researched as electrode materials for supercapacitors [5,6,7,8,9,10,11]. Wang et al. reported the performance of Mn3O4 nanomaterials synthesized via a two-step hydrothermal method and applied in asymmetric supercapacitors [12]. Another approach to supercapacitors development is to integrate manganese oxides with other transition metals such as Ni, Zn and Co in formulating spinel structured NiMn2O4, ZnMn2O4 and CoMn2O4 nanomaterials [13,14,15,16]. Saravanakumar et al. synthesized CuMn2O4 nanoparticles with rice-like morphology, through a solvothermal method for pseudocapacitor applications [17]. However, manganese oxide-based electrodes suffer from low-rate capabilities and poor cycling stability due to crystallographic defects and heavy aggregation of particles which render fewer number of electro-active sites for diffusion of electrolyte ions [18]. A strategic approach to overcome this challenge is by designing a nanocomposite of the manganese oxide electrode with carbon-based materials for increased surface area, electro-activity and cycling stability.
Carbon nanotubes (CNTs), due to their unique properties such as excellent electrical conductivity, good thermal stability, high flexibility, high specific surface area (50–1315 m2 g−1) and high surface area-to-volume ratio, have been employed in the development of nanocomposite electrodes for supercapacitors because of their ability to enhance the specific capacitance, energy and power density of the materials [19,20]. Shakir et al. demonstrated a facile two-step fabrication method for synthesizing spinel nickel cobaltite (NiCo2O4) nanostructures anchored to MWCNTs for flexible pseudocapacitors [21]. In another investigation, Geng et al. fabricated NiCo2O4/CNT core-shell nanocomposite hybrids by in-situ synthesis of ultrafine NiCo2O4 nanoparticles onto acid-functionalized CNT-films [22]. The regulation effects of CNT pre-treatment were mainly investigated on the electrochemical characteristics of NiCo2O4/CNT nanocomposites as electrode materials for flexible pseudo-capacitors. The study concluded that the introduction of redox-active heteroatoms such as carbon, oxygen, hydrogen, nitrogen and sulphur onto the MWCNTs network promoted a pseudocapacitive mechanism in conjunction with the EDLC mechanism.
Composite electrode materials such as ZnMn2O4/MWCNTs, Mn3O4/MWCNTs, V2O5/CNT and MnCoOx/MWCNTs [23,24,25,26,27,28] have also been reported. Fabricating metal oxides with MWCNTs promotes a dual charge storage mechanism comprised of an electrochemical double layer capacitor (EDLC) mechanism from MWCNTs and the Faradaic redox reaction mechanism. The MWCNTs support structure additionally provides good structural flexibility, high aspect ratio and strong mechanical strength which helps to reduce nanoparticle agglomeration [26]. Although CuMn2O4 has been synthesized for supercapacitor applications [16], CuMn2O4/MWCNTs electrode material has not been reported for supercapacitors.
Leveraging the unique properties of MWCNTs, CuMn2O4/MWCNTs nanocomposite, in addition to the pristine electrode materials, was synthesized in this work, and the electrochemical properties of the materials investigated in two different aqueous electrolytes (3 M KOH and 3 M LiOH) in order to ascertain the electrolyte where the electrode demonstrated better cycling stability, since electrochemical stability is related to the cycle life and safety of supercapacitors [29]. The distribution of CuMn2O4 nanoparticles throughout the MWCNTs framework was made flexible by first introducing hydrophilic groups to the MWCNTs surface through functionalization in H2SO4-HNO3 acid mixture. These groups function as nucleation sites for in situ positioning of CuMn2O4 nanoparticles on the MWCNTs surface. The functionalised nanotubes provided good mechanical support, increased surface area and more efficient electronic channels for faster ion diffusion as evidenced in the improved structural and electrochemical stability of the CuMn2O4/MWCNTs nanocomposite.

2. Experimental Section and Characterizations

2.1. Materials and Reagents

The list of chemical reagents used in all electrode material synthetic procedures include potassium permanganate (KMnO4, 99.0%), diethylene glycol (C4H10O3, 99.0%), ethanol (CH3CH2OH, 99.8%), copper (II) nitrate trihydrate (Cu(NO3)2·3H2O, 99%), manganese (II) nitrate tetrahydrate (Mn(NO3)2·4H2O, 99%), nitric acid (HNO3, 65%), citric acid monohydrate (C6H8O7·H2O, 99.5%), ammonium hydroxide (NH4OH, 25%), raw multi-walled carbon nanotubes (MWCNTs, 95%), sulphuric acid (H2SO4, 98%), polyvinylpyrrolidone (PVP, average mol wt 40,000), hydrazine hydrate (N2H4·H2O, 50–60%), carbon black (CB), polyvinylidene fluoride (PVDF), N-methyl-2-pyrollidone (NMP, 99.5%), potassium hydroxide (KOH, 85%) and lithium hydroxide (LiOH, 98%). All these chemical reagents were purchased from Merck (Johannesburg, South Africa) and used without further purification. Nickel foam was purchased from Material Technology International Corporation (MTI Corp., Richmond, CA, USA).

2.2. Synthetic Procedures

2.2.1. Synthesis of Mn3O4 Electrode Material

The pristine electrode material was synthesized through a hydrothermal method. An amount of 1.50 g of KMnO4 was dispersed in 50 mL of ultra-pure water and subjected to ultrasonic treatment for 30 min followed with the addition of 10 mL of diethylene glycol. The mixture was transferred into a 100 mL Teflon-lined stainless-steel autoclave and heated at 180 °C for 12 h. After cooling to room temperature, the mixture was filtered by means of centrifugation and washed repeatedly with ethanol and distilled water. The solid product was dried overnight in a vacuum oven at 60 °C. The experimental diagram for the synthetic procedures of all the materials is displayed in Figure 1.

2.2.2. Synthesis of CuMn2O4 Nanoparticles

1.55 g of Cu(NO3)2·3H2O was dissolved in 6 mL of dilute HNO3. To this solution was added 2.5 g of Mn(NO3)2·4H2O, 0.67 g of C6H8O7·H2O and 40 mL of ultra-pure water. The mixture was allowed to stir for 30 min followed by a drop wise addition of NH4OH which brought the solution pH to 8 prior to heating in an autoclave at 180 °C for 12 h. The solid was collected by centrifuging and washing with ethanol and distilled water. This product was vacuum-dried overnight at 60 °C and subsequently annealed in a muffle furnace at 400 °C for 4 h.

2.2.3. Synthesis of CuMn2O4/MWCNTs

Firstly, MWCNTs were oxidized by adding 10 mg of raw MWCNTs to a concentrated acid- mixture of H2SO4 (98%) and HNO3 (65%) in a 3:1 ratio. The oxidation of the MWCNTs reduces their hydrophobicity and improves their wettability and solubility in polar media [30]. An amount of 5 mg of the acid-treated MWCNTs was dispersed in 60 mL of diethylene glycol for subsequent use. Thereafter, 1.55 g Cu(NO3)2·3H2O, 2.5 g Mn(NO3)2·4H2O, 0.67 g of C6H8O7·H2O and 0.4 g polyvinylpyrrolidone (PVP) were dissolved in 40 mL ultra-pure water and the solution pH adjusted by the dropwise addition of 10 mL N2H4·H2O to pH 8. To this solution was added the previously dispersed MWCNTs and the mixture subjected to ultrasonic treatment and magnetic stirring for 30 min followed by refluxing at 180 °C for 4 h. A black precipitate was collected after centrifuging and washing several times with deionised water and ethanol. The product was dried overnight at 60 °C in a vacuum oven and finally annealed in a muffle furnace at 500 °C for 5 h to obtain the CuMn2O4/MWCNTs. electrode material.

2.3. Material Characterization

The phases and crystal structures of the electrode materials were investigated by X-ray powder diffraction (XRD) using a focused D8 Advanced Diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) equipped with a copper K α radiation ( λ = 1.54056 Å) source, scanning from 10° to 90° at an operating voltage of 40 kV and current of 40 mA. The crystallite size calculations were performed from the data extrapolation of the most intense Bragg diffraction peak. The surface morphology of electrode materials were explored via scanning electron microscopy (SEM) using a Zeiss Ultra Scanning Electron Microscope (Carl Zeiss AG, Oberkochen, Germany). Internal structural images were obtained via high-resolution transmission electron microscopy (HR-TEM) by a Tecnai G2F2O X-Twin MAT 200 kV Field Emission Transmission Electron Microscope from FEI (Eindhoven, Netherlands). Functional groups present in the electrode materials were examined using a Perkin Elmer Spectrum 100 Series Attenuated Total Reflectance Fourier Transform Infrared (ATR–FTIR) spectrometer (PerkinElmer Incorporated, Waltham, MA, USA). Bond vibrations and rotation information of the functional groups were obtained by employing Raman spectroscopy using a HORIBA Scientific XploRA PLUS Raman Microscope (HORIBA, Northampton, UK) equipped with a 532 nm laser source and 1 μm resolution. Small-angle X-ray scattering (SAXS) measurements of the pristine and modified electrode materials were conducted with an Anton Paar SAXSpace spectrometer (Anton Paar, Graz, Austria), equipped with a copper K α radiation ( λ = 1.54056 Å) source.

2.4. Electrochemical Measurements

Cyclic voltammetry (CV), galvanostatic charge-discharge (GCD), and electrochemical impedance spectroscopy (EIS) experiments were conducted in a three-electrode cell configuration, in both 3 M KOH and 3 M LiOH aqueous electrolytes, at room temperature on a BioLogic VMP-300 (BioLogic, Seyssinet-Pariset, France) electrochemical workstation. The three-electrode half-cell consisted of the working electrode (active material coated onto nickel foam substrate), reference electrode (Ag/AgCl in 3 M KCl), and the platinum wire counter electrode. The working electrodes were prepared by mixing the active electrode materials with carbon black (CB) conducting agent and polyvinylidene fluoride (PVDF) binder in a 70:20:10 weight percent ratio, followed by the addition of a few drops of anhydrous N-methyl-2-pyrrolidone (NMP) to form a homogeneous slurry. The slurry was coated onto the nickel foam and dried at 80 °C for 12 h. CV curves were obtained for the three-electrode cells within the voltage ranges of 0–0.7 V and 0–0.8 V, at specific applied scan rates of 5, 10, 20, 50, and 100 mV s−1. GCD experiments were conducted within the potential range of 0–0.6 V at current loadings of 0.5, 1, 2, 3, and 4 A g−1. The experimental data for EIS analysis was acquired within the frequency range of 10 mHz to 100 kHz with a sinusoidal perturbation amplitude of 5 mV at a 0.324 V applied DC potential.

3. Results and Discussion

3.1. Scanning Electron Microscopy (SEM)

The surface morphologies of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials were investigated by SEM, as displayed in Figure 2. The Mn3O4 material shows aggregated particles with elongated and plate-like morphology as seen in Figure 2a, with average particle sizes ranging between 65–105 nm. The image also displays large secondary particles of approximately 145 nm originating from the aggregation of smaller primary particles. The particle distribution is in agreement with the SAXS profile shown in Figure 2b, representing the pair-distance distribution function (PDDF) of the Mn3O4. The PDDF profile of Mn3O4 illustrates small-angle scattering patterns showing average particle sizes between 60 and 90 nm. However, a decrease in the scattering fraction number (P) is observed within the particle size range of 110–160 nm, indicating aggregation of the bigger Mn3O4 nanoparticles. The CuMn2O4 morphology displayed in Figure 2c shows clusters of geometric nanoparticles with calculated average particle sizes ranging between 75–120 nm. The nanopores in the material allow for efficient diffusion of electrolyte ions throughout the material surface. The PDDF profile of the CuMn2O4 material, shown in Figure 2d, illustrates spherically shaped nanoparticles having particle sizes ranging between 70–110 nm and with less particle aggregation due to increased stability of CuMn2O4 nanostructures confirmed from the XRD analysis. The morphology of CuMn2O4/MWCNTs, presented in Figure 2e, shows the clustered CuMn2O4 geometric nanoparticles intertwined with the MWCNTs which serve as alternative diffusion pathways for electrolyte ions to traverse the material for faster electrochemical kinetics and rate capability [30]. The CuMn2O4/MWCNTs has a broad and symmetrical PDDF profile as shown in Figure 2f indicating uniform distribution of particles with average sizes between 50–108 nm.

3.2. High-Resolution Transmission Electron Microscopy (HR-TEM)

HR-TEM images of the Mn3O4 material in Figure 3a,b show rectangular-shaped agglomerated particles with atomic planes represented by lattice fringes. The selected area diffraction (SAED) pattern displayed as an inset in Figure 3b, illustrates a particular sequence of diffracted spots characteristic to Mn3O4 structures [31]. The HR-TEM image viewed at a 2 nm scale, in Figure 3b, shows lattice fringes representing atomic lattice planes with the indexed planes having an inter-planar spacing distance of 0.25 nm. This d-spacing distance corresponds to the d-space distance of the (211) indexed Bragg diffraction peak from the Mn3O4 XRD pattern. The CuMn2O4 nanoparticles displayed in Figure 3c, show rhombic nanostructures with particle sizes ranging between 95–140 nm. The SAED pattern of the CuMn2O4 material, inset in Figure 3d, illustrates high crystallinity resulting from the brighter and closely oriented diffracted spots. The HR-TEM image of the CuMn2O4/MWCNTs material, displayed in Figure 3e, shows the rhombic CuMn2O4 nanoparticles anchored to the MWCNTs surface. Image processing analysis revealed the CuMn2O4/MWCNTs material having average particle sizes ranging between 50–108 nm. In the SAED pattern of CuMn2O4/MWCNTs, the diffracted spots formed symmetric rings each arising from their respective atomic planes. This observation indicates that the CuMn2O4/MWCNTs material possesses a polynanocrystalline phase [32]. The indexed atomic lattice planes of (222) and (400) in the SAED pattern of CuMn2O4/MWCNTs corresponds to the CuMn2O4 phases present within the nanostructure as confirmed by the two highly intense Bragg diffraction peaks indexed in the CuMn2O4/MWCNTs XRD pattern. The EDS spectra of all the materials are shown in Figure S1. All the elements in Mn3O4 are present in the spectrum. However, the K and Si signals are due to residues and contaminants from the starting materials. The Cu in the EDS spectrum of Mn3O4 nanoparticles emanated from the Cu grid, and the C signal observed came from the carbon coated onto the sample to improve contrast during analysis. The EDS spectrum of CuMn2O4 confirmed the presence of the constituent elements with the observed Ni signal originating from the Ni grid used during analysis. The EDS spectrum of the CuMn2O4/MWCNTs showed the presence of all the constituent elements.

3.3. Structural Characterisation

The XRD patterns of the Mn3O4, CuMn2O4, and CuMn2O4/MWCNTs electrode materials, shown in Figure 4a, display sharp and intense Bragg diffraction peaks thus suggesting good crystallization of the synthesized electrode materials. The major diffraction peaks observed in the XRD pattern of Mn3O4 located at diffraction angles of 28.96°, 32.36°, 36.12° and 44.45° are indexed with miller indices of (112), (103), (211) and (220), respectively. The phase is identified as Mn3O4 from its Joint Committee on Powder Diffraction Standards (JCPDS) data file number 24–0734 (Hausmannite, syn), with a body-centred tetragonal (BCT) crystal structure and a 141/amd (141) space group. The XRD pattern of the CuMn2O4 illustrates broader diffraction peaks with higher intensities compared to the XRD pattern of Mn3O4. The high intensity diffraction peaks observed in the XRD pattern of CuMn2O4 located at 30.40°, 35.86°, 38.46° and 43.67° diffraction angles are indexed to the (220), (311), (222) and (400) planes, respectively. The XRD pattern identifies a face-centered cubic (FCC) crystal structure and Fd3m (227) space group of CuMn2O4 nanoparticles, which matches well with its JCPDS #: 84–0543 data file [33]. The XRD pattern of CuMn2O4/MWCNTs illustrates the highest intensity Bragg diffraction peaks at 2 θ angles of 30.40°, 38.72°, 43.53° and 53.81° assigned with their respective (220), (222), (400) and (422) lattice planes. These atomic lattice planes are indexed according to the existence of a uniform CuMn2O4/MWCNTs phase with an Fd3m (227) space group characteristic of FCC crystal systems [33,34]. It is observed that the diffraction peaks of CuMn2O4/MWCNTs occur at 2 θ positions analogous to that of the spinel CuMn2O4, implying they match well with one another. The sharper and narrower diffraction peaks of the composite CuMn2O4/MWCNTs indicate better crystallinity and electrochemical conductivity of the material as reflected in its lower charge transfer resistance compared to the Mn3O4 and CuMn2O4 in the EIS data. The crystallite size calculations were based on the Scherrer’s equation below [35]:
D ( h , k , l ) = K λ β cos θ
where D is the mean crystallite size (nm), (h, k, l) are the miller indices indexed to the Bragg diffraction peak, K is the Scherrer’s constant (0.89), λ is the wavelength of the X-ray beam (0.154 nm), β is the full width at half maximum (FWHM) of the Bragg diffraction peak and θ is the Bragg’s angle.
Table 1 shows the values of D, inter-planar spacing distance (d) and lattice parameters (a) and (c) of the three electrode materials with the CuMn2O4/MWCNTs exhibiting the smallest crystallite size of 51.26 nm. This decreased size as the SAXS and TEM data also show implies a larger surface area and hence more available sites for electrolyte interaction thus leading to higher electrochemical performance as shown in the CV and EIS data. The CuMn2O4 has a crystallite size of 77.94 nm, which is slightly larger than that of Mn3O4 (74.04 nm). This can be attributed to the CuMn2O4 FCC crystal structure containing eight tetrahedral and four octahedral interstitial sites, at which the Cu2+ and Mn3+ transition metal ions are situated [20].
The Raman spectra of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs were acquired for molecular bond stretching and vibration analyses. The spectrum of the Mn3O4 material shown in Figure 4b, illustrates a low intensity Raman peak at approximately 485 cm−1, characteristic to symmetric F2g(2) stretching vibration modes of Mn–O species within the octahedral MnO6 sub-lattices of the Mn3O4 crystal structure [36]. The CuMn2O4 displays a more intense Raman-active bands than that of Mn3O4. The intensity of the broad Raman shoulder band (F2g(1)) along the shift range of 900–1500 cm−1 in the CuMn2O4 spectrum is closely related to the oxidation/valence state of Mn3+ atoms within the CuMn2O4 crystal structure [37]. This increase in the intensities of the CuMn2O4 active bands compared to Mn3O4 is attributable to the higher oxidation/valence state of Mn3+ ions within the CuMn2O4 crystal lattice. The Raman spectrum of the CuMn2O4/MWCNTs material illustrates two partially distinctive peaks, enlarged on the inset graph, positioned at 1467 cm−1 and 1529 cm−1, ascribed to the D and G bands, respectively. The D band is attributed to the breathing mode of k-point phonons of A1g symmetry from the carbon aromatic rings, whereas the G band is assigned to the E2g phonons from the stretching vibrations of sp2 hybridized carbon atoms in the MWCNTs [38,39]. The incorporation of CuMn2O4 nanoparticles to the MWCNTs network can be confirmed by observing the additional Raman Eg band at 362 cm−1. The high-intensity Raman peak occurring within the shift range of 2430–2435 cm−1 represents the 2D band, which is assigned to an overtone mode of longitudinal optical phonon from the dispersion and frequency of 2k-point phonons [40].
The FT-IR spectra of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs are presented in Figure 4c. The broad absorption band, appearing consistently in each material’s spectrum, within the high frequency range of 3000–3660 cm−1 is ascribed to the bond stretching vibrational modes of hydroxyl (O–H) groups in water [41]. The intense absorption bands located at lower frequencies of 520 cm−1 and 623 cm−1 correspond to the vibrations in Mn–O bond within the Mn3O4 structure. The absorption bands located at wavenumbers 1123 cm−1 and 1720 cm−1 are characteristic to the C–O and C–C stretching vibrations, originating from diethylene glycol residues. In the spectrum of CuMn2O4, the two bands representing Mn–O bond vibrations appeared less intense due to vibrations from the neighbouring Cu–O bond, as illustrated by the faint absorption band at 660 cm−1. The FT-IR spectrum of the CuMn2O4/MWCNTs material shows the absorption bands corresponding to Mn–O and Cu–O functional groups, however, at weaker intensities due to C = O and C–O bond vibrations, at 1630 cm−1 and 1020 cm−1, derived from the carbon backbone of the MWCNTs network [14].

3.4. Electrochemical Characterization

3.4.1. Cyclic Voltammetry (CV)

Figure 5 shows the cyclic voltammograms of Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials. Each CV is characterised by a pair of redox peaks arising from the Mn2+/Mn3+ and Mn3+/Mn4+ redox couples. The redox reaction mechanisms of the electrode materials are presented as follows [42,43].
Mn 3 O 4 + OH + H 2 O 3 MnOOH + e
CuMn 2 O 4 + OH + H 2 O 2 MnOOH + CuOOH + e
CuOOH + OH CuO 2 + H 2 O + e
The CuMn2O4/MWCNTs showed better electrochemical performance in both electrolytes than the other electrode materials as observed in Figure 5a,b. This is attributed to the three-dimensional support structure of the MWCNTs that stabilized the electroactive sites and promoted more electrolyte ion diffusion. Figure 5c,d describe the effect of potential scan rates on the CV of the CuMn2O4/MWCNTs in each electrolyte. Scanning from 5–100 mV s−1, the current was observed to increase as the scan rate increases, with the anodic peaks shifting to more positive potentials and the cathodic peaks to more negative potentials; a phenomenon that shows that the electrochemical process on the electrode/electrolyte interface is quasi-reversible with the CuMn2O4/MWCNTs electrode exhibiting a higher current response.
It is known that charge storage in pseudocapacitors involves redox processes with capacitive behaviour [44], and since a capacitive current follows a linear dependence on scan rate, Figure S2a,b indicate that the pseudocapacitance in the CuMn2O4/MWCNTs electrode material is a surface-controlled electrochemical process governed by the equation [45].
i = d Q d t = C d E d t = C ν
where Q is the voltammetric charge, C is the capacitance and d E d t is the scan rate, ν . This pseudocapacitive behaviour is corroborated by the results from the log-log plot of current against scan rate as shown in Figure S2c,d, obtained according to the power law relationship [46].
i = a ν b
where a and b are adjustable parameters. The value of b provides important information about the charge-storage kinetics. When b is 1, the charge storage mechanism is highly capacitive and when b is 0.5, the process is diffusion-controlled. The CuMn2O4/MWCNTs has a b value of 0.86 and shows a more pseudocapacitive charge-storage mechanism in LiOH than in KOH where the b value is 0.58 suggesting a more diffusion-controlled kinetics. The charge storage capability of the different electrodes was evaluated by determining the specific capacitance from the integrated area under the CV curves in Figure 5c,d. Values were obtained according to the equation [47]:
C s p = V a V c I ( V ) d V ν m ( V c V a )
where Csp is the specific capacitance (F g−1) of the electrode materials, Va and Vc are the two potential limits (V) of the integrated area under the CV curves, I is the corresponding current (A) obtained from the current density, m is the average mass loading (g) of electroactive materials and ν is the potential scan rate (V s−1).
Figure S3a shows the plot of specific capacitance against scan rates for the CuMn2O4/MWCNTs electrode material in KOH and LiOH. The specific capacitance values obtained at the lowest applied scan rate of 5 mV s−1 are 659.71 F g−1 for KOH and 1267.43 F g−1 for LiOH. This also confirms the earlier observations made that electrochemical performance of the electrode materials in LiOH is better than in KOH. In their work with MnO2 in different aqueous hydroxides, Misnon et al. [48] reported the highest specific capacitance value in LiOH than in KOH and NaOH and attributed it to the smaller ionic radius of Li+ that allowed easier intercalation-de-intercalation reaction to occur. The specific capacitance is higher at lower scan rates and decreased as the scan rate progresses. This is due to the fact that at lower scan rates, more electroactive sites are accessible by the electrolyte ion whereas at higher scan rates, the penetration distance of the electrolyte ion into the material decreases and the ions are only limited to the surface. Figure S3a also shows that the CuMn2O4/MWCNTs electrode analysed in LiOH retained about 73% of its initial specific capacitance after increased scan rates from 5–100 mV s−1, whereas analysis of the nanocomposite cathode in KOH aqueous electrolyte has a capacitance retention of 43% thus indicating enhanced ionic interactions of Li+ which resulted in higher rate capabilities. Figure S3b illustrates the comparative CVs of the CuMn2O4/MWCNTs electrode before and after 6000 cycles in 3 M LiOH. The specific capacitance recorded at 100 mV s−1 gave a value of 643.12 F g−1 after 6000 cycles, which is about 70% of its initial capacitance before cycling. This result indicates a relatively good electrochemical stability of the CuMn2O4/MWCNTs electrode. Figure S3c shows the CVs of the bare MWCNTs electrodes in both electrolytes. The CVs are quasi-rectangular in shape, with no discernible electrochemistry, indicating the supercapacitor behaviour. The CVs of Mn3O4 at various scan rates are shown in Figure S4.

3.4.2. Electrochemical Impedance Spectroscopy (EIS)

Electrochemical impedance spectroscopy was used to shed more light on the charge transfer behaviour and the capacitive nature of the electrode materials. In Figure 6 the Nyquist and Bode plots of the pristine Mn3O4, spinel CuMn2O4, and CuMn2O4/MWCNTs electrode materials in both electrolytes are shown. In the high-frequency region of a typical Nyquist plot, the intercept of the curve at the x–axis refers to the bulk solution resistance of the electrode material at the electrode/electrolyte interface [43,49]. This bulk solution resistance (Rs) involves intrinsic resistances experienced by the electroactive material which include ionic resistances from electrolyte ions at the electrode/electrolyte interface, as well as contact surface resistances occurring between the electroactive species and Ni-foam (substrate) current collector. The diameter of the semi-circle of impedance is equivalent to the charge-transfer resistance (Rct) in the high frequency region, with the slanted line occurring in the low frequency region representing the Warburg impedance (Zw), which mathematically expresses the diffusion of electrolyte ions according to an equivalent circuit model. The Rct value signifies the resistance of electron transfer kinetics occurring at the electrode/electrolyte interface during charging/discharging processes. The double layer capacitance (Cdl) models the effect of charge accumulation in electrolyte ions at the electrode material’s surface [50,51]. The values of the kinetic parameters from EIS for all the analysed electrode materials in both electrolytes are summarised in Table 2 and Table 3.
The Nyquist plots of Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs in KOH electrolyte are shown in Figure 6a. The Mn3O4 is characterised by a larger semi-circle at high frequency with the highest Rct value and an indistinct Warburg diffusion line at the low frequency region, whereas the CuMn2O4 and CuMn2O4/MWCNTs have diminished semi-circles at higher frequency which is confirmed by their lower Rct values indicating better electron transfer kinetics at the electrode/electrolyte interface. Both also have the Warburg diffusion line at low frequency which is, however, more prominent in CuMn2O4/MWCNTs than in the CuMn2O4 electrode. The Warburg line that is near-parallel to the imaginary impedance axis suggests a more pseudocapacitive behaviour in the CuMn2O4/MWCNTs. The more vertical the curve, the more a supercapacitor (or pseudocapacitor) behaves like an ideal capacitor [52,53,54,55] and lower Zw values suggest that the kinetics of charge storage is capacitive [56].
The inset in the figure shows an expanded portion of the high frequency region indicating the Rs values of the electrodes. The behaviour of the electrode materials in LiOH electrolyte, shown in Figure 6b, follow an almost similar trend as observed in KOH. Each electrode is characterised by a semi-circle in the high frequency region and an inclined Warburg line at low frequency. The Rct and Rs values decrease from Mn3O4, CuMn2O4 to CuMn2O4/MWCNTs as shown in Table 2. The straight line in the low frequency region is near-parallel to the imaginary impedance axis, especially for the CuMn2O4/MWCNTs electrode which again, signifies the characteristics of a pseudocapacitor. The enlarged portion of the high frequency region is shown in the inset. Figure 6c compares the Nyquist plots of the CuMn2O4/MWCNTs in both electrolytes. Each plot is marked by a semi-circle at high frequency and a Warburg impedance at low frequency. The figure shows that the CuMn2O4/MWCNTs electrode material exhibited a more vertical Warburg line in LiOH than in KOH aqueous electrolyte, thus indicating a more pseudocapacitive behaviour in the former.
The Bode phase angle profiles of the materials analysed in KOH are shown in Figure 6d. The CuMn2O4/MWCNTs has the highest phase angle followed by the CuMn2O4 and Mn3O4 as presented in Table 2. It is well known that an ideal capacitor has a phase angle of about 90° [53,57]. The nanocomposite electrode with the highest phase angle that tends towards that of an ideal capacitor can be said to exhibit a more pseudocapacitive character than the others. The increased electrochemical performance is attributed to the synergy between the spinel CuMn2O4 and MWCNTs which increased the surface conductivity, fast ion transport and sufficient number of active sites, leading to higher adsorption of ions and better performance [21,54,56]. A similar observation was made in Figure 6e for the analysis of the materials in LiOH.
The log impedance against log frequency diagrams in Figure 7a,b illustrated a lower impedance in the electrodes for each electrolyte group and with the electrodes in LiOH having the least impedance overall. This result is further confirmed by the faster angular frequency response-time ( τ ) exhibited by the CuMn2O4/MWCNTs electrode material thus indicating more efficient ion transport routes enabled by the MWCNTs throughout the CuMn2O4/MWCNTs matrix. The equivalent circuit model, displayed as an inset in Figure 6c, was used to perform EIS data fitting using an EC-Lab EIS Z-fit software. The Nyquist and Bode impedance diagrams in Figure 7c,d, showed a 32% increase and a 27% decrease in resistance and conductivity of the CuMn2O4/MWCNTs electrode after 6000 cycles in 3 M LiOH aqueous electrolyte. The Nyquist and Bode impedance plots of the bare MWCNTs electrode, shown in Figure S6a,b, illustrates a lower charge transfer resistance and higher phase angle in LiOH compared to KOH aqueous electrolyte. The values of the kinetic parameters from EIS for the bare MWCNTs electrode in both KOH and LiOH electrolytes are summarised in Table S1. Comparative Bode plots between the CuMn2O4/MWCNTs electrode in KOH and LiOH is presented in Figure S6c with the CuMn2O4/MWCNTs in LiOH being more pseudocapacitive because of the higher phase angle value as shown in Table 2 and Table 3.

3.4.3. Galvanostatic Charge/Discharge (GCD)

Figure 8 compares the galvanostatic charge/discharge profiles of the electrode materials. Figure 8a–d show typical charge/discharge curves of the materials at different current densities and in different electrolytes. All the three electrodes exhibited better electrochemical performance in LiOH as also observed in the CV results. The equation below was used to calculate the specific capacitance at various current loadings [58]. Table 4 shows the comparison of supercapacitance values of electrode materials evaluated in LiOH and KOH aqueous electrolytes.
C sp = I Δ t Δ V m
where Csp is the specific capacitance (F g−1), I is the discharge current (A), Δt is the discharge time (s), ΔV is the cell operating potential window (V) and m is the average mass of active electrode materials coated onto the Ni-foam current collector.
The histogram shown in Figure 9a illustrates the values of the specific capacitances of the CuMn2O4/MWCNTs electrode material at various current loads ranging between 0.5–4 A g−1, for both electrolytes. At all current loads, the material analysed in LiOH gave the highest specific capacitance. The histogram shows the CuMn2O4/MWCNTs electrode retaining 74% of its initial specific capacitance when analysed in LiOH while the CuMn2O4/MWCNTs electrode examined in KOH sustained 44% of its initial specific capacitance. The GCD curves of Mn3O4 at various current loadings shown in Figure S5 illustrates the electrode material exhibiting a longer discharge time and higher specific capacitance in LiOH (450.86 s and 853.15 F g−1) compared to KOH (79.69 s and 56.58 F g−1) aqueous electrolytes. GCD curves of the bare MWCNTs, shown in Figure S6, indicates a higher specific capacitance performance in LiOH (50.37 F g−1) as compared to KOH (38.66 F g−1) at 2 A g−1.
The cycling stability of the CuMn2O4/MWCNTs was tested in both electrolytes as shown in Figure 9b. For the CuMn2O4/MWCNTs in LiOH electrolyte, a slight capacitance loss was noticed at the beginning of the cycling process but later stabilized around the 200th cycle and remained constant thereafter. The initial capacitance loss was more obvious for CuMn2O4/MWCNTs in KOH up to the 1000th cycle and became stable afterwards. After 6000 cycles, the CuMn2O4/MWCNTs electrode showed a specific capacitance retention of 88% and 64% in LiOH and KOH electrolytes, respectively, at 1 A g−1 current load. Overall, the CuMn2O4/MWCNTs electrode’s cycle stability and increased higher specific capacitance resulted in a greater charge storage capability and longer cycle life [67,68].
The Coulombic efficiency per cycle number of the CuMn2O4/MWCNTs electrode material in both 3 M KOH and 3 M LiOH aqueous electrolytes is shown in Figure 9c. It is observed from the graph that the electrode material in KOH electrolyte exhibited a 91.5% Coulombic efficiency for the first cycle, up to 97.3% at the 6000th cycle. On the other hand, the composite electrode material in LiOH electrolyte exhibited a 96.9% Coulombic efficiency at the first cycle and up to 99.6% at the 6000th cycle. This result also indicates that the cycling stability of the CuMn2O4/MWCNTs electrode material is better in the LiOH electrolyte than in the KOH electrolyte. Although the size of the hydrated cation K+ is smaller than that of Li+ (K+: 3.31 Å; Li+: 3.82 Å) which confers a higher mobility and conductivity in KOH electrolyte than in LiOH electrolyte, the higher capacitance values and superior performance in the LiOH electrolyte than in the KOH electrolyte is attributed to an easy intercalation/de-intercalation of Li+ ions into the electrode due to the smaller radius of Li+ (crystal size: 0.60 Å) than K+ (crystal size: 1.33 Å) which increased the double layer as well as pseudocapacitance [48]. Since intercalation and de-intercalation of alkaline electrolyte ions is generally involved for pseudocapacitive materials, the bare (unsolvated) ionic size is considered to have a pronounced effect on the pseudocapacitive behaviour [69]. Yuan and co-worker [70] reported a higher specific capacitance and cyclic stability in LiOH electrolyte than in KOH electrolyte and ascribed it to the Li+ insertion/extraction in MnO2 solid which is different with that electrode in KOH. This phenomenon was also corroborated in the work of Manickam et al. [71]. In a related study, Inamdar et al. [72] prepared NiO and recorded a specific capacitance that was almost two times higher in NaOH (Na+ crystal radius: 0.95 Å) than in KOH electrolyte. The authors attributed this to a higher intercalation rate of Na+ ions in the NIO electrode. It is well known that electrochemical stability is strongly related to cycle life [29], hence, the higher cyclability witnessed in LiOH indicates that the electrode has better stability and longer cycle life after repeated charge/discharge in this electrolyte compared to the KOH electrolyte. The GCD curves of CuMn2O4/MWCNTs before and after 6000 cycles in LiOH electrolyte is shown in Figure 9d and gave a discharge time and specific capacitance value of 130.4 s and 409.25 F g−1, respectively.

4. Conclusions

In this study, Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials were prepared. Results from XRD revealed the Mn3O4 material with a body-centred tetragonal (BCT) crystal structure, whereas both the CuMn2O4 and CuMn2O4/MWCNTs materials belong to the face-centred cubic (FCC) crystal systems. The crystal lattice structures of the Mn3O4 and CuMn2O4 materials are supported by their respective HR-TEM images, portraying two-dimensional rectangular-shaped Mn3O4 nanoparticles and slightly distorted rhombic CuMn2O4 nanostructures. Studies on the electrochemical properties of the electrode materials were conducted in both 3 M KOH and 3 M LiOH electrolytes. CV and EIS measurements illustrated a superior electrochemical performance in the nanocomposite electrode material compared to the CuMn2O4 and Mn3O4 electrode materials. The nanocomposite gave smaller charge transfer resistances in both electrolytes than the Mn3O4 and CuMn2O4 electrode materials and better electrochemical performance than the Mn3O4 and CuMn2O4 electrode materials. This is ascribed to the synergy between the spinel CuMn2O4 and MWCNTs which increased the surface conductivity, fast ion transport and sufficient number of active sites, leading to higher adsorption of ions and better performance. The results also revealed a better electrochemical response for all the electrode materials in the LiOH electrolyte than in the KOH electrolyte, attributed to faster intercalation/de-intercalation of Li+ ions into the materials. A specific capacitance of 1652.91 F g−1 at 0.5 A g−1 specific current was recorded for the CuMn2O4/MWCNTs electrode in the LiOH electrolyte while in the KOH electrolyte, the specific capacitance calculated was 653.41 F g−1 at 0.5 A g−1. The electrode also showed a higher cycling stability in the LiOH electrolyte with a 99.6% Coulombic efficiency, and looks more promising as an electrode material for pseudocapacitor applications in this electrolyte than in the KOH electrolyte, where it exhibited a lower charge storage behaviour.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12193514/s1, Figure S1: EDS spectra of Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs materials, with their corresponding elemental percentage composition tables; Figure S2: Graphs of current against square root of scan rate for CuMn2O4/MWCNTs electrode in 3 M KOH (a) and 3 M LiOH (b) electrolytes; Graphs of log current against log scan rate of CuMn2O4/MWCNTs electrode material in 3 M KOH (c) and 3 M LiOH (d) aqueous electrolytes; Figure S3: Histogram of the specific capacitance of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes at various scan rates (a); CV curves of CuMn2O4/MWCNTs electrode material before and after 6000 cycles at 100 mV s−1 in 3 M LiOH aqueous electrolyte (b); CV curves of the bare MWCNTs electrode at 20 mV s-1 in both 3 M KOH and 3 M LiOH aqueous electrolytes (c); Figure S4: CV curves of Mn3O4 material at various scan rates in 3 M KOH aqueous electrolyte (a) and in 3 M LiOH aqueous electrolyte (b). Graphs of corresponding specific capacitances at various scan rates in 3 M KOH aqueous electrolyte (c) and in 3 M LiOH (d); Figure S5: GCD curves of Mn3O4 material at various current loadings in 3 M KOH aqueous electrolyte (a) and its corresponding graph of specific capacitances at various current loadings (b); GCD curves of Mn3O4 in 3 M LiOH aqueous electrolyte (c) with the respective graph of specific capacitances at different current loadings (d); Figure S6: Nyquist (a) and Bode (b) plots of the bare MWCNTs electrode in both 3 M KOH and 3 M LiOH aqueous electrolytes; Comparative Bode plot of CuMn2O4/MWCNTs electrode material in both 3 M KOH and 3 M LiOH aqueous electrolytes (c); GCD curves of the bare MWCNTs electrode, obtained at 2 A g−1, in both 3 M KOH and 3 M LiOH aqueous electrolytes (d); Table S1: EIS fitted data of the bare acid treated MWCNTs material in 3 M KOH and 3 M LiOH aqueous electrolytes. References [43,48,69,70,71,72] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, E.I.I. and C.O.I.; methodology, C.N. and P.E.; validation, M.M.N., C.O.I. and E.I.I.; formal analysis, C.N., C.O.I., M.M.N. and P.E.; investigation, C.N., C.O.I. and M.M.N.; resources, E.I.I. and C.O.I.; data curation, C.N., C.O.I.,M.M.N., P.E. and E.I.I.; writing—original draft preparation, C.N.; writing—review and editing, C.O.I., C.N., M.M.N. and E.I.I.; visualization, C.N., C.O.I., P.E. and M.M.N.; supervision, C.O.I. and E.I.I.; project administration, C.O.I. and E.I.I.; funding acquisition, E.I.I. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Research Foundation (NRF) of South Africa: (i) NRF South Africa Research Chair Initiative (SARChI) Grant (UID: 85102); and (ii) NRF Blue Skies Grant (UID: Grant No: 110981) and the South African Department of Science and Innovation (DSI)-National Nanoscience Postgraduate Teaching and Training Platform (NNPTTP) Scholarship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy or ethical restrictions.

Acknowledgments

The Electron Microscopy Unit (EMU) of the University of the Western Cape and iThemba Laboratories for the Accelerator Based Science (iThemba Labs) are acknowledged for instrument use.

Conflicts of Interest

The authors do not have any conflict of interest to declare.

References

  1. Chen, Y.; Zhang, X.; Xu, C.; Xu, H. The fabrication of asymmetry supercapacitor based on MWCNTs/MnO2/PPy composites. Electrochim. Acta 2019, 309, 424–431. [Google Scholar] [CrossRef]
  2. Pratibha, G.; Srinivas, I.; Rao, K.V.; Shanker, A.K.; Raju, B.M.K.; Choudhary, D.K.; Rao, K.S.; Srinivasarao, C.; Maheswari, M. Net global warming potential and greenhouse gas intensity of conventional and conservation agriculture system in rainfed semi-arid tropics of India. Atmos. Environ. 2016, 145, 239–250. [Google Scholar] [CrossRef]
  3. Zhang, H.; Li, H.; Sun, Z.; Jia, D. One-step hydrothermal synthesis of NiCo2S4 nanoplates/nitrogen-doped mesoporous carbon composites as advanced electrodes for asymmetric supercapacitors. J. Power Sources 2019, 439, 227082. [Google Scholar] [CrossRef]
  4. Akhtar, M.S.; Gul, I.H.; Baig, M.M.; Akram, M.A. Binder-free pseudocapacitive nickel cobalt sulfide/MWCNTs hybrid electrode directly grown on nickel foam for high rate supercapacitors. Mater. Sci. Eng. B 2021, 264, 114898. [Google Scholar] [CrossRef]
  5. Rajkumar, S.; Elanthamilan, E.; Balaji, T.E.; Sathiyan, A.; Jafneel, N.E.; Merlin, J.P. Recovery of copper oxide nanoparticles from waste SIM cards for supercapacitor electrode material. J. Alloy. Compd. 2020, 849, 156582. [Google Scholar] [CrossRef]
  6. Ojha, G.P.; Pant, B.; Muthurasu, A.; Chae, S.H.; Park, S.J.; Kim, T.; Kim, H.Y. Three-dimensionally assembled manganese oxide ultrathin nanowires: Prospective electrode material for asymmetric supercapacitors. Energy 2019, 188, 116066. [Google Scholar] [CrossRef]
  7. Li, B.; Zhang, X.; Dou, J.; Hu, C. Facile synthesis of pseudocapacitive Mn3O4 nanoparticles for high-performance supercapacitor. Ceram. Int. 2019, 45, 16297–16304. [Google Scholar] [CrossRef]
  8. Zhang, Q.; Gu, D.; Li, H.; Xu, Z.; Sun, H.; Li, J.; Wang, L.; Shen, L. Energy release from RuO2/RuO2 supercapacitors under dynamic discharge conditions. Electrochim. Acta 2021, 367, 137455. [Google Scholar] [CrossRef]
  9. Pant, B.; Park, M.; Ojha, G.P.; Park, J.; Kuk, Y.S.; Lee, E.J.; Kim, H.Y.; Park, S.J. Carbon nanofibers wrapped with zinc oxide nano-flakes as promising electrode material for supercapacitors. J. Colloid Interface Sci. 2018, 522, 40–47. [Google Scholar] [CrossRef]
  10. Lang, J.W.; Kong, L.B.; Wu, W.J.; Luo, Y.C.; Kang, L. Facile approach to prepare loose-packed NiO nano-flakes materials for supercapacitors. Chem. Commun. 2008, 35, 4213–4215. [Google Scholar] [CrossRef]
  11. Xu, J.; Gao, L.; Cao, J.; Wang, W.; Chen, Z. Preparation and electrochemical capacitance of cobalt oxide (Co3O4) nanotubes as supercapacitor material. Electrochim. Acta 2010, 56, 732–736. [Google Scholar] [CrossRef]
  12. Wang, Z.; Fang, J.; Hao, Y.; Chen, C.; Zhang, D. High-performance Mn3O4 nanomaterials synthesized via a new two-step hydrothermal method in asymmetric supercapacitors. Mater. Sci. Semiond. Process. 2021, 130, 105823. [Google Scholar] [CrossRef]
  13. Ren, H.; Zhang, L.; Zhang, J.; Miao, T.; Yuan, R.; Chen, W.; Wang, Z.; Yang, J.; Zhao, B. Na+ pre-intercalated Na0.11MnO2 on three-dimensional graphene as cathode for aqueous zinc ion hybrid supercapacitor with high energy density. Carbon 2022, 198, 46–56. [Google Scholar] [CrossRef]
  14. Saha, S.; Roy, A.; Ray, A.; Das, T.; Nandi, M.; Ghosh, B.; Das, S. Effect of particle morphology on the electrochemical performance of hydrothermally synthesized NiMn2O4. Electrochim. Acta 2020, 353, 136515. [Google Scholar] [CrossRef]
  15. Yun, H.; Zhou, X.; Zhu, H.; Zhang, M. One-dimensional zinc-manganate oxide hollow nanostructures with enhanced supercapacitor performance. J. Colloid Interface Sci. 2021, 585, 138–147. [Google Scholar] [CrossRef] [PubMed]
  16. Kanaujiya, N.; Kumar, N.; Singh, M.; Sharma, Y.; Varma, G.D. CoMn2O4 Nanoparticles Decorated on 2D MoS2 Frame: A Synergetic Energy Storage Composite Material for Practical Supercapacitor Applications. J. Energy Storage 2021, 35, 102302. [Google Scholar] [CrossRef]
  17. Saravanakumar, B.; Lakshmi, S.M.; Ravi, G.; Ganesh, V.; Sakunthala, A.; Yuvakkumar, R. Electrochemical properties of rice-like copper manganese oxide (CuMn2O4) nanoparticles for pseudocapacitor applications. J. Alloys Compd. 2017, 723, 115–122. [Google Scholar] [CrossRef]
  18. Kang, L.; Hung, C.; Zhang, J.; Zhang, M.; Zhang, N.; He, Y.; Luo, C.; Wang, C.; Zhou, X.; Wu, X. A new strategy for synthesis of hierarchical MnO2-Mn3O4 nanocomposite via reduction-induced exfoliation of MnO2 nanowires and its application in high-performance asymmetric supercapacitor. Compos. B. Eng. 2019, 178, 107501. [Google Scholar] [CrossRef]
  19. Hsu, H.L.; Miah, M.; Saha, S.K.; Chen, J.H.; Chen, L.C.; Hsu, S.Y. Three-dimensional bundle-like multiwalled carbon nanotubes composite for supercapacitor electrode application. Mater. Today Chem. 2021, 22, 100569. [Google Scholar] [CrossRef]
  20. Ali, G.A.M.; Megiel, E.; Romanski, J.; Algarni, H.; Chong, K.F. A wide potential window symmetric supercapacitor by TEMPO functionalized MWCNTs. J. Mol. Liq. 2018, 271, 31–39. [Google Scholar] [CrossRef]
  21. Shakir, I. High Performance Flexible pseudocapacitors based on nano-architectured spinel nickel cobaltite anchored multiwall carbon nanotubes. Electrochim. Acta 2014, 132, 490–495. [Google Scholar] [CrossRef]
  22. Geng, L.; Xu, S.; Liu, J.; Guo, A.; Hou, F. Effects of CNT-film pretreatment on the characteristics of nico2o4/cnt core-shell hybrids as electrode material for electrochemistry capacitor. Electroanalysis 2017, 29, 778–786. [Google Scholar] [CrossRef]
  23. Aruchamy, K.; Nagaraj, R.; Manohara, H.M.; Nidhi, M.R.; Mondal, D.; Ghosh, D.; Nataraj, S.K. One-step green route synthesis of spinel ZnMn2O4 nanoparticles decorated on MWCNTs as a novel electrode material for supercapacitor. Mater. Sci. Eng. B 2020, 252, 114481. [Google Scholar] [CrossRef]
  24. Mondal, C.; Ghosh, D.; Aditya, T.; Sasmal, A.K.; Pal, T. Mn3O4 nanoparticles anchored to multiwall carbon nanotubes: A distinctive synergism for high-performance supercapacitors. New J. Chem. 2015, 39, 8373–8380. [Google Scholar] [CrossRef]
  25. Sun, G.; Ren, H.; Shi, Z.; Zhang, L.; Wang, Z.; Zhan, K.; Yan, Y.; Yang, J.; Zhao, B. V2O5/vertically-aligned carbon nanotubes as negative electrode for asymmetric supercapacitor in neutral aqueous electrolyte. J. Colloid Interface Sci. 2021, 588, 847–856. [Google Scholar] [CrossRef]
  26. Peng, W.; Su, Z.; Wang, J.; Li, S.; Chen, K.; Song, N.; Luo, S.; Xie, A. MnCoOx-multi-walled carbon nanotubes composite with ultra-high specific capacitance for supercapacitors. J. Energy Storage 2022, 51, 104519. [Google Scholar] [CrossRef]
  27. Shi, Z.; Sun, G.; Yuan, R.; Chen, W.; Wang, Z.; Zhang, L.; Zhan, K.; Zhu, M.; Yang, J.; Zhao, B. Scalable fabrication of NiCo2O4/reduced graphene oxide composites by ultrasonic spray as binder-free electrodes for supercapacitors with ultralong lifetime. J. Mater. Sci. Technol. 2022, 99, 260–269. [Google Scholar] [CrossRef]
  28. Li, L.; Jiang, G.; Ma, J. CuMn2O4/graphene nanosheets as excellent anode for lithium-ion battery. Mater. Res. Bull. 2018, 104, 53–59. [Google Scholar] [CrossRef]
  29. Pal, B.; Yang, S.; Ramesh, S.; Thangadurai, V.; Jose, R. Electrolyte selection for supercapacitive devices: A critical review. Nanoscale Adv. 2019, 1, 3807–3835. [Google Scholar] [CrossRef] [Green Version]
  30. Sezer, N.; Koc, M. Oxidative acid treatment of carbon nanotubes. Surf. Interfaces 2019, 14, 1–8. [Google Scholar] [CrossRef]
  31. Li, T.; Guo, C.; Sun, B.; Li, T.; Li, Y.; Hou, L.; Wei, Y. Well-shaped Mn3O4 tetragonal bipyramids with good performance for lithium ion batteries. J. Mater Chem. A 2015, 3, 7248–7254. [Google Scholar] [CrossRef]
  32. Kavil, J.; Anjana, P.M.; Joshy, D.; Babu, A.; Raj, G.; Periyat, P.; Rakhi, R.B. g-C3N4/CuO and g-C3N4/Co3O4 nanohybrid structures as efficient electrode materials in symmetric supercapacitors. RSC Adv. 2019, 9, 38430–38437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Abel, M.J.; Pramothkumar, A.; Senthikumar, N.; Jothivenkatachalam, K.; Inbaraj, P.F.H.; Prince, J.J. Flake-like CuMn2O4 nanoparticles synthesized via co-precipitation method for photocatalytic activity. Phys. B Condens. Matter 2019, 572, 117–124. [Google Scholar]
  34. Wan, X.; Tang, N.; Xie, Q.; Zhao, S.; Zhou, C.; Dai, Y.; Yang, Y. A CuMn2O4 spinel oxide as a superior catalyst for the aerobic oxidation of 5-hydroxymethylfurfural toward 2,5-furandicarboxylic acid in aqueous solvent. Catal. Sci. Technol. 2021, 11, 1497–1509. [Google Scholar] [CrossRef]
  35. Ndipingwi, M.M.; Ikpo, C.O.; Hlongwa, N.W.; Myalo, Z.; Ross, N.; Masikini, M.; John, S.V.; Baker, P.G.; Roos, W.D.; Iwuoha, E.I. Orthorhombic Nanostructured Li2MnSiO4/Al2O3 Supercapattery Electrode with Efficient Lithium-Ion Migratory Pathway. Batter. Supercaps 2018, 1, 223–235. [Google Scholar] [CrossRef]
  36. Gao, F.; Qin, S.H.; Zhang, Y.H.; Gu, J.F.; Qu, J.Y. Highly efficient formation of Mn3O4-graphene oxide hybrid aerogels for use as the cathode material of high performance lithium ion batteries. New Carbon Mater. 2020, 35, 121–130. [Google Scholar] [CrossRef]
  37. Ismail, F.M.; Ramadan, M.; Abdellah, A.M.; Ismail, I.; Allam, N.K. Mesoporous spinel manganese zinc ferrite for high-performance supercapacitors. J. Electroanal. Chem 2018, 817, 111–117. [Google Scholar] [CrossRef]
  38. Hlongwa, N.W.; Ikpo, C.O.; Ndipingwi, M.M.; Nolly, C.; Raleie, N.; Dywili, N.; Iwuoha, E.I. Graphene-functionalised olivine lithium manganese phosphate derivatives for high performance lithium-ion capacitors. Electroanalysis 2020, 32, 2812–2826. [Google Scholar] [CrossRef]
  39. Zhang, L.; Guo, Y.; Shen, K.; Huo, J.; Liu, Y.; Guo, S. Ion-matching porous carbons with ultra-high surface area and superior energy storage performance for supercapacitors. J. Mater. Chem. A 2019, 7, 9163–9172. [Google Scholar] [CrossRef]
  40. Brolly, C.; Parnell, J.; Bowden, S. Raman spectroscopy: Caution when interpreting organic carbon from oxidising environments. Planet. Space Sci. 2016, 121, 53–59. [Google Scholar] [CrossRef] [Green Version]
  41. Rahmanifar, M.S.; Hemmati, M.; Noori, A.; I-Kady, M.F.E.; Mousavi, M.F.; Kaner, R.B. Asymmetric supercapacitors: An alternative to activated carbon negative electrodes based on earth abundant elements. Mater. Today Energy 2019, 12, 26–36. [Google Scholar] [CrossRef]
  42. Nguyen, N.V.; Tran, T.V.; Luong, S.T.; Pham, T.M.; Nguyen, K.V.; Vu, T.D.; Nguyen, H.S.; To, N.V. Facile synthesis of a NiCo2O4 nanoparticles mesoporous carbon composite as electrode materials for supercapacitor. ChemistrySelect 2020, 5, 7060–7068. [Google Scholar] [CrossRef]
  43. Gu, Y.; Wu, J.; Wang, X.; Liu, W.; Yan, S. Producing “Symbiotic” Reduced Graphene Oxide/Mn3O4 Nanocomposites Directly from converting Graphite for High-Performance Supercapacitor Electrodes. ACS Omega 2020, 5, 18975–18986. [Google Scholar] [CrossRef]
  44. Noori, A.; El-Kady, M.F.; Rahmanifar, M.S.; Kaner, R.B.; Mousavi, M.F. Towards establishing standard performance metrics for batteries, supercapacitors and beyond. Chem. Soc. Rev. 2019, 5, 1272–1341. [Google Scholar] [CrossRef]
  45. Li, J.; Tang, Z.; Zhang, Z. Layered hydrogen titanate nanowires with novel lithium intercalation properties. Chem. Mater. 2005, 17, 5848–5855. [Google Scholar] [CrossRef]
  46. Wang, Y.; Song, Y.; Xia, Y. Electrochemical capacitors: Mechanism, materials, systems, characterization and applications. Chem. Soc. Rev. 2016, 45, 5925–5950. [Google Scholar] [CrossRef]
  47. Vadiyar, M.M.; Bhise, S.C.; Patil, S.K.; Kolekar, S.S.; Chang, J.Y.; Ghule, A.V. comparative study of individual and mixed aqueous electrolytes with ZnFe2O4 nano-flakes thin film as an electrode for supercapacitor application. ChemistrySelect 2016, 1, 959–966. [Google Scholar] [CrossRef]
  48. Misnon, I.I.; Aziz, R.A.; Zain, N.K.M.; Vidhyadharan, B.; Krishnan, S.G.; Jose, R. High performance MnO2 nanoflower electrode and the relationship between solvated ion size and specific capacitance in highly conductive electrolytes. Mater. Res. Bull. 2014, 57, 221–230. [Google Scholar] [CrossRef] [Green Version]
  49. Balamurugan, J.; Nguyen, T.T.; Aravindan, V.; Kim, N.H.; Lee, J.H. Flexible solid-state asymmetric supercapacitors based on nitrogen-doped graphene encapsulated ternary metal-nitrides with ultralong cycle life. Adv. Funct. Mater. 2018, 28, 1804663. [Google Scholar] [CrossRef]
  50. Cao, L.; Tang, G.; Mei, J.; Liu, H. Construct hierarchical electrode with NixCo3-xS4 nanosheet coated on NiCo2O4 nanowire arrays grown on carbon fiber paper for high-performance asymmetric supercapacitors. J. Power Sources 2017, 359, 262–269. [Google Scholar] [CrossRef]
  51. Qu, Z.; Shi, M.; Wu, H.; Liu, Y.; Jiang, J.; Yan, C. An efficient binder-free electrode with multiple carbonized channels wrapped by NiCo2O4 nanosheets for high-performance capacitive energy storage. J. Power Sources 2019, 410–411, 179–187. [Google Scholar] [CrossRef]
  52. Liu, C.; Yu, Z.; Neff, D.; Zhamu, A.; Jang, B.Z. Graphene-based supercapacitor with an ultrahigh energy density. Nano Lett. 2010, 10, 4863–4868. [Google Scholar] [CrossRef] [PubMed]
  53. Krishnamoorthy, K.; Pazhamalai, P.; Sahoo, S.; Kim, S.J. Titanium carbide sheet based high performance wire type solid state supercapacitors. J. Mater. Chem. A 2017, 5, 5726–5736. [Google Scholar] [CrossRef]
  54. Chhetri, K.; Tiwari, A.; Dahal, B.; Ojha, G.; Mukhiya, T.; Lee, M.; Kim, T.; Chae, S.; Muthurasu, A.; Kim, H. A ZIF-8-derived nanoporous carbon nanocomposite wrapped with Co3O4-polyaniline as an efficient electrode material for an asymmetric supercapacitor. J. Electroanal. Chem. 2020, 856, 113670. [Google Scholar] [CrossRef]
  55. Rajkumar, S.; Elanthamilan, E.; Merlin, J.P.; Sathiyan, A. Enhanced electrochemical behavior of FeCo2O4/PANI electrode material for supercapacitors. J. Alloys Compd. 2021, 874, 159876. [Google Scholar] [CrossRef]
  56. Mandal, D.; Routh, P.; Mahato, A.K.; Nandi, K. Electrochemically modified graphite paper as an advanced electrode substrate for supercapacitor application. J. Mater. Chem. A 2019, 7, 17547–17560. [Google Scholar] [CrossRef]
  57. Pazhamalai, P.; Krishnamoorthy, K.; Sahoo, S.; Kim, S.J. Two-dimensional molybdenum diselenide nanosheets as a novel electrode material for symmetric supercapacitors using organic electrolyte. Electrochim. Acta 2019, 295, 591–598. [Google Scholar] [CrossRef]
  58. Kulandaivalu, S.; Hussein, M.Z.; Jaafar, A.M.; Abdah, M.A.A.M.; Azman, N.H.N.; Sulaiman, Y. A simple strategy to prepare a layer-by-layer assembled composite of Ni-Co LDHs on polypyrrole/rGO for a high specific capacitance supercapacitor. RSC Adv. 2019, 9, 40479–40486. [Google Scholar] [CrossRef] [Green Version]
  59. Xu, L.; Wang, S.; Zhang, X.; He, T.; Lu, F.; Li, H.; Ye, J. A facile method of preparing LiMnPO4/reduced graphene oxide aerogel as cathodic material for aqueous lithium-ion hybrid supercapacitors. Appl. Surf. Sci. 2018, 428, 977–985. [Google Scholar] [CrossRef]
  60. Li, X.; Ding, R.; Yi, L.; Shi, W.; Xu, Q.; Liu, E. Mesoporous Ni-P@NiCo2O4 composite materials for high performance aqueous asymmetric supercapacitors. Electrochim. Acta 2016, 222, 1169–1175. [Google Scholar] [CrossRef]
  61. Mariappan, V.K.; Krishnamoorthy, K.; Pazhamalai, P.; Sahoo, S.; Nardekar, S.S.; Kim, S.J. Nanostructured tenary metal chalcogenide-based binder-free electrodes for high energy density asymmetric supercapacitors. Nano Energy 2019, 57, 307–316. [Google Scholar] [CrossRef]
  62. Li, Y.; Tang, F.; Wang, R.; Wang, C.; Liu, J. A novel dual-ion hybrid supercapacitor based on NiCo2O4 nanowire cathode and MoO2-C nanofilm anode. ACS Appl. Mater. Interfaces 2016, 8, 30232–30238. [Google Scholar] [CrossRef] [PubMed]
  63. Luo, Y.; Zhang, H.; Guo, D.; Ma, J.; Li, Q.; Chen, L.; Wang, T. Porous NiCo2O4-reduced graphene oxide (rGO) composite with superior capacitance retention for supercapacitors. Electrochim. Acta 2014, 132, 332–337. [Google Scholar] [CrossRef]
  64. Iqbal, N.; Wang, X.; Babar, A.A.; Yu, J.; Ding, B. Highly flexible NiCo2O4/CNTs doped carbon nanofibers for CO2 adsorption and supercapacitor electrodes. J. Colloid Interface Sci. 2016, 476, 87–93. [Google Scholar] [CrossRef] [PubMed]
  65. Mazinani, B.; Kazazi, M.; Mobarhan, G.; Shokouhimehr, M. The combination synthesis of Ag-doped MnCo2O4 nanoparticles for supercapacitor applications. JOM 2019, 71, 8. [Google Scholar] [CrossRef]
  66. Xuan, H.; Xu, Y.; Zhang, Y.; Li, H.; Han, P.; Du, Y. One-step combustion synthesis of porous CNTs/C/NiMoO4 composites for high-performance asymmetric supercapacitors. J. Alloys Compd. 2018, 745, 135–146. [Google Scholar] [CrossRef]
  67. Krishnaveni, M.; Wu, J.J.; Anandan, S.; Ashokkumar, M. Facile synthesis of SnO2 nanoparticle intercalacted unzipped multi-walled carbon nanotubes via an ultrasound-assisted route for symmetric supercapacitor devices. Sustain. Energy Fuels 2020, 4, 5120–5131. [Google Scholar] [CrossRef]
  68. Chhetri, K.; Dahal, B.; Tiwari, A.; Mukhiya, T.; Muthurasu, A.; Kim, T.; Kim, H.; Kim, H.Y. Integrated hybrid of graphitic carbon-encapsulated CuxO on multilayered mesoporous carbon from copper MOFs and polyaniline for asymmetric supercapacitor and oxygen reduction reactions. Carbon 2021, 179, 89–99. [Google Scholar] [CrossRef]
  69. 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]
  70. Yuan, A.; Zhang, Q. A novel hybrid manganese dioxide/activated carbon supercapacitor using lithium hydroxide electrolyte. Electrochem. Commun. 2006, 8, 1173–1178. [Google Scholar] [CrossRef]
  71. Manickam, M.; Singh, P.; Issa, T.B.; Thurgate, S.; Marco, R.D. Lithium insertion into manganese dioxide electrode in MnO2/Zn aqueous battery: Part I. A preliminary study. J. Power Sources 2004, 130, 254–259. [Google Scholar] [CrossRef]
  72. Inamdar, A.I.; Kim, Y.S.; Pawar, S.M.; Kim, J.H.; Im, H.; Kim, H. Chemically grown, porous, nickel oxide thin-film for electrochemical supercapacitors. J. Power Sources 2011, 196, 2393–2397. [Google Scholar] [CrossRef]
Figure 1. Experimental diagram for the synthetic procedures of Mn3O4 (a), CuMn2O4 (b) and CuMn2O4/MWCNTs (c) electrode materials.
Figure 1. Experimental diagram for the synthetic procedures of Mn3O4 (a), CuMn2O4 (b) and CuMn2O4/MWCNTs (c) electrode materials.
Nanomaterials 12 03514 g001aNanomaterials 12 03514 g001b
Figure 2. SEM image of Mn3O4 (a) viewed at a 100 nm scale, with the corresponding PDDF profile (b). SEM image of CuMn2O4 (c) at the 100 nm scale with its related PDDF profile (d). SEM image of CuMn2O4/MWCNTs (e) at the 100 nm scale with its complementary PDDF profile (f).
Figure 2. SEM image of Mn3O4 (a) viewed at a 100 nm scale, with the corresponding PDDF profile (b). SEM image of CuMn2O4 (c) at the 100 nm scale with its related PDDF profile (d). SEM image of CuMn2O4/MWCNTs (e) at the 100 nm scale with its complementary PDDF profile (f).
Nanomaterials 12 03514 g002
Figure 3. HR-TEM images of Mn3O4 viewed at 50 nm scale (a) and 2 nm scale (b). HR-TEM images of CuMn2O4 at the 50 nm scale (c) and 2 nm scale (d). HR-TEM images of CuMn2O4/MWCNTs at the 50 nm scale (e) and 2 nm scale (f).
Figure 3. HR-TEM images of Mn3O4 viewed at 50 nm scale (a) and 2 nm scale (b). HR-TEM images of CuMn2O4 at the 50 nm scale (c) and 2 nm scale (d). HR-TEM images of CuMn2O4/MWCNTs at the 50 nm scale (e) and 2 nm scale (f).
Nanomaterials 12 03514 g003
Figure 4. XRD patterns (a), Raman spectra (b) and FTIR spectra (c) of Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials.
Figure 4. XRD patterns (a), Raman spectra (b) and FTIR spectra (c) of Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials.
Nanomaterials 12 03514 g004aNanomaterials 12 03514 g004b
Figure 5. CV curves of individual materials scanned at 20 mV s−1 in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; CV curves of CuMn2O4/MWCNTs electrode material in 3 M KOH (c) and 3 M LiOH (d) aqueous electrolytes at different scan rates.
Figure 5. CV curves of individual materials scanned at 20 mV s−1 in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; CV curves of CuMn2O4/MWCNTs electrode material in 3 M KOH (c) and 3 M LiOH (d) aqueous electrolytes at different scan rates.
Nanomaterials 12 03514 g005aNanomaterials 12 03514 g005b
Figure 6. Nyquist plots of all analysed electrode materials in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; comparative Nyquist plot of CuMn2O4/MWCNTs in both 3 M KOH and 3 M LiOH aqueous electrolytes (c); corresponding Bode phase impedance plot of all electrode materials in 3 M KOH (d) and 3 M LiOH (e) aqueous electrolytes.
Figure 6. Nyquist plots of all analysed electrode materials in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; comparative Nyquist plot of CuMn2O4/MWCNTs in both 3 M KOH and 3 M LiOH aqueous electrolytes (c); corresponding Bode phase impedance plot of all electrode materials in 3 M KOH (d) and 3 M LiOH (e) aqueous electrolytes.
Nanomaterials 12 03514 g006aNanomaterials 12 03514 g006b
Figure 7. Graphs of log impedance against log frequency for all analysed electrode materials in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; Nyquist (c) and Bode (d) plots of CuMn2O4/MWCNTs electrode material before and after 6000 cycles in 3 M LiOH aqueous electrolyte.
Figure 7. Graphs of log impedance against log frequency for all analysed electrode materials in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; Nyquist (c) and Bode (d) plots of CuMn2O4/MWCNTs electrode material before and after 6000 cycles in 3 M LiOH aqueous electrolyte.
Nanomaterials 12 03514 g007
Figure 8. GCD curves of all electrode materials obtained at 0.5 A g−1 in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; GCD curves of CuMn2O4/MWCNTs electrode material in 3 M KOH (c) and 3 M LiOH (d) aqueous electrolytes.
Figure 8. GCD curves of all electrode materials obtained at 0.5 A g−1 in 3 M KOH (a) and 3 M LiOH (b) aqueous electrolytes; GCD curves of CuMn2O4/MWCNTs electrode material in 3 M KOH (c) and 3 M LiOH (d) aqueous electrolytes.
Nanomaterials 12 03514 g008
Figure 9. Histogram of the specific capacitance of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes (a); cycling performance of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes after 6000 cycles at 1 A g−1 current load (b); Coulombic efficiency of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes over 6000 cycles obtained at 1 A g−1 current loading (c); GCD curves of CuMn2O4/MWCNTs electrode material before and after 6000 cycles at 1 A g−1 in 3 M LiOH aqueous electrolyte (d).
Figure 9. Histogram of the specific capacitance of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes (a); cycling performance of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes after 6000 cycles at 1 A g−1 current load (b); Coulombic efficiency of CuMn2O4/MWCNTs electrode material in 3 M KOH and 3 M LiOH aqueous electrolytes over 6000 cycles obtained at 1 A g−1 current loading (c); GCD curves of CuMn2O4/MWCNTs electrode material before and after 6000 cycles at 1 A g−1 in 3 M LiOH aqueous electrolyte (d).
Nanomaterials 12 03514 g009
Table 1. Summary of crystallographic parameters calculated from the XRD of the Mn3O4 CuMn2O4, and CuMn2O4/MWCNTs electrode materials.
Table 1. Summary of crystallographic parameters calculated from the XRD of the Mn3O4 CuMn2O4, and CuMn2O4/MWCNTs electrode materials.
Electrode MaterialD (nm)d (nm)a (nm)c (nm)
Mn3O474.040.200.580.95
CuMn2O477.940.270.94
CuMn2O4/MWCNTs51.260.230.77
Table 2. EIS fitted data of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials in 3 M KOH aqueous electrolyte.
Table 2. EIS fitted data of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials in 3 M KOH aqueous electrolyte.
Electrode MaterialRs (Ω)Cdl (F)Rct (Ω)Zw (Ω s−1/2)τ (s rad−1) (°)
Mn3O40.982.58 × 10−31.294.984.62 × 10−346
CuMn2O40.454.64 × 10−30.832.863.83 × 10−355
CuMn2O4/MWCNTs0.367.93 × 10−30.630.863.34 × 10−364
Table 3. EIS fitted data of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials in 3 M LiOH aqueous electrolyte.
Table 3. EIS fitted data of the Mn3O4, CuMn2O4 and CuMn2O4/MWCNTs electrode materials in 3 M LiOH aqueous electrolyte.
Electrode MaterialRs (Ω)Cdl (F)Rct (Ω)Zw (Ω s−1/2)τ (s rad−1) (°)
Mn3O40.363.16 × 10−30.520.383.77 × 10−337
CuMn2O40.315.66 × 10−30.480.332.69 × 10−345
CuMn2O4/MWCNTs0.257.87 × 10−30.350.301.63 × 10−371
Table 4. Comparison of specific capacitance values of various supercapacitor electrode materials, evaluated in LiOH and KOH aqueous electrolyte solutions.
Table 4. Comparison of specific capacitance values of various supercapacitor electrode materials, evaluated in LiOH and KOH aqueous electrolyte solutions.
Electrode MaterialElectrolyteCsp (F g−1)Reference
CuMn2O4/MWCNTs3 M LiOH1652.9This work
LiMnPO4/rGO1 M LiOH464.5[59]
Ni–P/NiCo2O40.7 M LiOH1240[60]
Cu3SbS4/Ni–51 M LiOH835.2[61]
NiCo2O4//MoO2-C1 M LiOH94.9[62]
NiCo2O4/rGO2 M KOH777.1[63]
NiCo2O4/CNTs1 M KOH220[64]
MnCo2O4/Ag NPs6 M KOH942[65]
CNTs/C/NiMoO42 M KOH1037[66]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nolly, C.; Ikpo, C.O.; Ndipingwi, M.M.; Ekwere, P.; Iwuoha, E.I. Pseudocapacitive Effects of Multi-Walled Carbon Nanotubes-Functionalised Spinel Copper Manganese Oxide. Nanomaterials 2022, 12, 3514. https://doi.org/10.3390/nano12193514

AMA Style

Nolly C, Ikpo CO, Ndipingwi MM, Ekwere P, Iwuoha EI. Pseudocapacitive Effects of Multi-Walled Carbon Nanotubes-Functionalised Spinel Copper Manganese Oxide. Nanomaterials. 2022; 12(19):3514. https://doi.org/10.3390/nano12193514

Chicago/Turabian Style

Nolly, Christopher, Chinwe O. Ikpo, Miranda M. Ndipingwi, Precious Ekwere, and Emmanuel I. Iwuoha. 2022. "Pseudocapacitive Effects of Multi-Walled Carbon Nanotubes-Functionalised Spinel Copper Manganese Oxide" Nanomaterials 12, no. 19: 3514. https://doi.org/10.3390/nano12193514

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