Next Article in Journal
Efficient and Facile Synthetic Route of MoO3:MoS2 Hybrid Thin Layer via Oxidative Reaction of MoS2 Nanoflakes
Next Article in Special Issue
The Influence of Electrode Thickness on the Structure and Water Splitting Performance of Iridium Oxide Nanostructured Films
Previous Article in Journal
Electrochemical Reduction and Preparation of Cu-Se Thermoelectric Thin Film in Solutions with PEG
Previous Article in Special Issue
Poly(Thiophene)/Graphene Oxide-Modified Electrodes for Amperometric Glucose Biosensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanoparticles of Mixed-Valence Oxides MnXCO3-XO4 (0 ≤ X ≤ 1) Obtained with Agar-Agar from Red Algae (Rhodophyta) for Oxygen Evolution Reaction

by
Jakeline Raiane D. Santos
1,*,
Rafael A. Raimundo
2,
Thayse R. Silva
3,
Vinícius D. Silva
3,
Daniel A. Macedo
3,
Francisco J. A. Loureiro
4,
Marco A. M. Torres
1,
Domenica Tonelli
5 and
Uílame U. Gomes
1,*
1
Materials Science and Engineering Postgraduate Program, UFRN, Natal 59078-970, Brazil
2
Department of Physics, UFPB, João Pessoa 58051-900, Brazil
3
Materials Science and Engineering Postgraduate Program, UFPB, João Pessoa 58051-900, Brazil
4
Centre for Mechanical Technology and Automation, Mechanical Engineering Department, UA, 3810-193 Aveiro, Portugal
5
Department of Industrial Chemistry “Toso Montanari”, Industrial Chemistry, UNIBO, V.le Risorgimento 4, 40136 Bologna, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(18), 3170; https://doi.org/10.3390/nano12183170
Submission received: 8 August 2022 / Revised: 2 September 2022 / Accepted: 7 September 2022 / Published: 13 September 2022
(This article belongs to the Special Issue Nanomaterials for Electrochemical Applications)

Abstract

:
The development of efficient electrocatalysts for the oxygen evolution reaction (OER) is of paramount importance in sustainable water-splitting technology for hydrogen production. In this context, this work reports mixed-valence oxide samples of the MnXCo3-XO4 type (0 ≤ X ≤ 1) synthesized for the first time by the proteic sol-gel method using Agar-Agar as a polymerizing agent. The powders were calcined at 1173 K, characterized by FESEM, XRD, RAMAN, UV–Vis, FT-IR, VSM, and XPS analyses, and were investigated as electrocatalysts for the oxygen evolution reaction (OER). Through XRD analysis, it was observed that the pure cubic phase was obtained for all samples. The presence of Co3+, Co2+, Mn2+, Mn3+, and Mn4+ was confirmed by X-ray spectroscopy (XPS). Regarding the magnetic measurements, a paramagnetic behavior at 300 K was observed for all samples. As far as OER is concerned, it was investigated in an alkaline medium, where the best overpotential of 299 mV vs. RHE was observed for the sample (MnCo2O4), which is a lower value than those of noble metal electrocatalysts in the literature, together with a Tafel slope of 52 mV dec−1, and excellent electrochemical stability for 15 h. Therefore, the green synthesis method presented in this work showed great potential for obtaining electrocatalysts used in the oxygen evolution reaction for water splitting.

1. Introduction

In recent decades, with the exponential growth of the population, the intensification of using fossil fuels has generated great impacts on the environment [1]. Given this, the energy transition from fossil fuels to clean energy sources has become necessary, and major renewable energy technologies have been developed [2], such as geothermal [3], wind [4], solar [5,6], and biomass [7]. However, to benefit resources even more, it is necessary to use efficient devices for energy storage and conversion [8,9].
One of the energy conversion processes that has received a lot of attention is water splitting via electrolysis for the production of hydrogen [10], which consists of two semi-reactions: the hydrogen evolution reaction (HER) [11] and the oxygen evolution reaction (OER) [12,13,14]. During the water-splitting process, the kinetic reaction that takes place at the anode (OER) is slow and requires high energy consumption (overpotential) due to the four electrons transferred in the reaction, causing its efficiency to decrease [15]. Therefore, the use of effective electrocatalysts is necessary to accelerate the reaction and reduce the overpotential [16]. However, the most widely used and effective electrocatalysts for water splitting are compounds based on noble metals such as ruthenium and iridium, but the high cost and scarcity of these elements have limited their large-scale application [17,18]. Therefore, one of the main challenges is to develop new electrocatalysts made of low-cost and abundant materials while also offering high electrochemical performance [19].
In this context, mixed-valence transition metal oxides such as the MnXCo3-XO4 (Mn-Co-O) system have attracted attention, as they represent an important class of multifunctional materials for electrocatalysis [20], oxygen evolution reaction [21], oxygen reduction reaction [22], fuel cells [23], and batteries [24]. These applications are related to the intrinsic properties of Mn-Co-O due to the multiple oxidation states of the metal ions [25]. This material has a spinel-like structure, with general formula A[B]2O4, where A is represented by a cation in the tetrahedral site (Coordination Number 4), and B is a cation in the octahedral site (Coordination Number 6) [26]. Depending on the occupancy of the divalent (A) and trivalent (B) cations, spinels are classified into three categories, normal, inverse, and complex [27], which gives them interesting physical, chemical [28], and electrochemical properties [29].
Furthermore, the distribution of cations between the different coordination sites strongly depends on the synthesis conditions [30]; thus, several syntheses have been developed in recent years to obtain mixed-valence transition metal oxides. Among the most common methods for obtaining (Mn-Co-O) are the sol-gel [31], combustion [32], solvothermal [33], solid-state [34], co-precipitation [35], spray pyrolysis [36], and hydrothermal [37] methods. These processes directly influence the morphology and/or structure of the materials, especially the sol-gel method, which allows crystalline materials to be obtained [38,39].
However, another method that has been attracting the attention of many researchers is that named proteic sol-gel synthesis, which is defined as a modification of the traditional sol-gel method, and allows ceramic materials to be obtained at the nanometer scale [40]. This method involves the following steps: preparation of a composite solution, formation of an amorphous network via polymerization reactions, followed by hydrolysis, gel formation, removal of organic residues and water, and finally calcination [41,42]. In this context, the proteic sol-gel method has several advantages over the traditional method, such as simplicity, low cost, higher speed, and generation of less waste to the environment [43,44], since it employs organic precursors, which are called polymerizing agents, having the function of replacing citric acid and ethylene glycol as driving agents of the reaction. Furthermore, these agents have hydroxyl and carboxyl groups that promote the chelation of metal ions during the synthesis process [45]. Some examples of organic precursors are flavorless gelatin [46], coconut water [47], and Agar-Agar [48].
Agar-Agar is a biopolymer [49], hydrocolloid [50] and polysaccharide source [51], extracted from red algae of the class Rhodophyceae [52], which is composed of two main parts: agarose, which is responsible for the gelling ability, and agaropectin, which is the polymeric part resulting from the existence of various substituent groups, such as sulfates and methyl ethers among others [53]. Due to its gelling properties, biocompatibility, biodegradability, and non-toxicity [54], Agar-Agar is used in the food, leather, cosmetic, beverage, and pharmaceutical industries [55].
The present work aims to investigate the structural, optical, and magnetic properties through the Co/Mn ratios in the structure, as well as the electrochemical properties of MnXCo3-XO4-type mixed-valence transition metal oxide powders synthesized by the proteic sol-gel method using Agar-Agar as a polymerizing agent.

2. Experimental Section

2.1. Materials

Cobalt nitrate (Co(NO3)2·6H2O, (Sigma-Aldrich 99%, Saint-Louis, MO, USA), manganese nitrate (Mn(NO3)2·4H2O, Vetec 99%, Saint-Louis, MO, USA), and Agar-Agar (Gelialgas-Agargel, João Pessoa, Brazil) were used in the present study. Nickel foam (Ni 99.8%, porosity >95%) was purchased from QiJing Ltd., Ninghai, China.

2.2. Synthesis of the Proteic Sol-Gel Using Agar-Agar

The scheme of MnXCo3-XO4 (0 ≤ X ≤ 1) preparation is shown in Figure 1. The mixed oxides were synthesized using Co(NO3)2 6H2O and Mn(NO3)2 4H2O, while Agar-Agar was used as the polymerizing agent. First, 2.0 g of Agar-Agar was dispersed in 50 mL of distilled water at 60 °C, then the proper amounts in mols of the metal salts were added as follows for samples with X = 0.0 (cobalt nitrate: 0.0249 mol), X = 0.2 (cobalt nitrate: 0.0233 mol and manganese nitrate: 1.6673 mmol), X = 0.4 (cobalt nitrate: 0.0217 mol and manganese nitrate: 3.3437 mmol), X = 0.6 (cobalt nitrate: 0.0201 mol and manganese nitrate: 5.0337 mmol), X = 0.8 (cobalt nitrate: 0.0185 mol and manganese nitrate: 6.7332 mmol), and X = 1.0 (cobalt nitrate: 0.0169 mol and manganese nitrate: 8.4459 mmol), and the resulting solution was kept under stirring at 90 °C until the formation of a gel. The resulting gel was kept at 350 °C for 2 h. Hence, the obtained powders were ground and calcined in air at 900 °C. The samples MnXCo3-XO4 were labeled as: for X = 0.0 (Co3O4), X = 0.2 (Mn0.2Co2.8O4), X = 0.4 (Mn0.4Co2.6O4), X = 0.6 (Mn0.6Co2.4O4), X = 0.8 (Mn0.8Co2.2O4) and X = 1.0 (MnCo2O4).

2.3. Structural and Morphological Characterization

X-ray powder diffraction patterns (XRD) were obtained by a Shimadzu XRD-7000 diffractometer using Kα(Cu) = 1.5481 Å radiation. The 2θ range was investigated from 10° to 80° with a step size of 0.02° and acquisition time of 1 s. The crystallite size, lattice parameters, and atomic positions were determined by Rietveld refinement using the software Materials Analysis Using Diffraction (TOPAS). FT-IR spectra were performed by a Shimadzu IRPrestige21 spectrophotometer between 500 and 4000 cm−1, using KBr pellets. The ultraviolet–visible spectra (UV–Vis) were obtained in the UV-2600i spectrophotometer from Shimadzu. SEM images were obtained by a field-emission scanning electron microscope (FESEM, Carl Zeiss, Supra 35-VP Model) equipped with a Bruker EDS detector (XFlash 410-M). Surface chemical states were studied by X-ray photoelectron spectroscopy (XPS) using a SPECS Phoibos 150 spectrometer with a high-intensity monochromatic Al-Ka X-ray source (1486.6 eV). Samples were dispersed in acetone and deposited on silicon by drop-coating. Adventitious carbon C 1s with binding energy at 284.8 eV was used as reference energy. CasaXPS software was used for spectra deconvolution, thus obtaining the height, area, and position of the analyzed peaks. All the symmetric peaks were fitted using Gaussian and Lorentzian functions. Magnetic measurements were obtained using a vibrating sample magnetometer (VSM) from Lakeshore, model 7400, at room temperature, with a maximum magnetic field applied up to +15.0 KOe.

2.4. Electrochemical Characterization

All electrochemical studies were performed in an alkaline aqueous solution (KOH, 1 M pH = 13.6) at room temperature by a PGSTAT204 with FRA32M module (Metrohm Autolab) using a three-electrode setup with a platinum plate and Hg/HgO as counter and reference electrodes, respectively. The samples MnXCo3-XO4 (0 ≤ X ≤ 1) were used for the fabrication of the working electrodes. Catalytic inks were prepared by mixing 5 mg of each catalyst with 50 μL of Nafion solution (5 wt%) and dispersing the mixture in 500 μL of isopropyl alcohol. Then, inks were drop-casted onto Ni foams (1 × 1 cm) on clean substrates and dried at room temperature for 5 h to prepare the working electrodes. Linear sweep voltammetry (LSV) was performed at 5 mV s−1. Electrochemical impedance spectroscopy (EIS) was carried out using dc potentials (1.4 V vs. RHE) in the frequency range of 0.1 Hz–10 kHz and voltage amplitude of 10 mV. All measured potentials (with iR correction) were converted to the reversible hydrogen electrode (RHE) using the Nernst equation (ERHE = EHg/HgO + 0.059×pH + 0.098). Overpotential (ղ) values were calculated by the equation ղ = ERHE—1.23 V. The stability tests were conducted by chronopotentiometry analysis using multi-steps of 10–20 mA cm−2. Figure 2 shows the complete characterization process.

3. Results

3.1. X-ray Diffraction (XRD)

The refined X-ray diffraction patterns of the MnXCo3-XO4 (0 ≤ X ≤ 1) are shown in Figure 3a. As noted, all peaks are characteristic of the cubic phases of Co3O4 (structure of Spinel#MgAl2O4 type, with lattice parameter a = b = 8.072(3) Å, ICSD n° 36256, space group Fd-3mS (227)) [56] and MnCo2O4 (structure of Spinel#MgAl2O4 type, with lattice parameter a = b = 8.28(2) Å, ICSD n° 18544, space group Fd-3mZ (227)) [57]. No secondary phases corresponding to impurities were detected. The ICSD n° 36,256 was used to fit the MnXCo3-XO4 samples (X < 1), while the 18,544 ICSD file was applied to refine the sample with composition X = 1. The observed patterns are similar to those reported previously for pure and doped cobaltites [16,20,58]. All crystallographic parameters, including crystallite size and lattice parameter, as well as the agreement indices (Rwp, Rexp e χ2) for samples of MnXCo3-XO4 (0 ≤ X ≤ 1) are gathered in Table 1.
Figure 3b shows the magnification of the most intense diffraction peak (311), located between 36° and 37.5°. As shown, increasing Mn content shifts the (311) peak to lower angles, indicating a continuous increase in the lattice parameter from 8.0757(1) for x = 0, 8.0759(8) for x = 0.2, 8.1183(7) for x = 0.6, 8.1675(8) for x = 0.8 to 8.2381(6) for x = 1. The total width at half maximum intensity (FWHM) also increases gradually with the increase in Mn (0.1322 for x = 0, 0.1433 for x = 0.2, 0.1513 for x = 0.4, 0.1613 for x = 0.6, 0.1633 for x = 0.8 and 0.1692 for x = 1), signaling a progressive reduction in crystallite size and increase in strain (90.6 nm for x = 0, 82.5 nm for x = 0.2, 80 nm for x = 0.4, 71 nm for x = 0.6, 68.8 nm for x = 0.8, 66 nm for x = 1, with the only exception of the sample corresponding to X = 0.4 for which the 2θ angle is slightly higher than the one displayed by the sample with X = 0.2). The largest variations of FWHM occur in the 0 ≤ X ≤ 0.6 range. The Mn+2 has a radius of 0.80 Å and it is larger than the radii of Co+3 (0.63 Å) and Co+2 (0.65 Å); thus, it causes distortion and strain in the Co3O4 lattice, resulting in a decrease in crystallite size. In addition, other manganese oxidation states may be present in the samples such as Mn+3 and Mn+4. Mn+3 (3d4) is responsible for the Jahn–Teller phenomenon, which also develops a distortion in the lattice and an intrinsic strain that leads to a decrease in the crystallite size [20,59,60,61]. The trends of lattice parameter and crystallite size are in agreement with previous reports [16,20]. The maximum values of the agreement factors Rwp and Rexp from the Rietveld analyses are 8.14% and 6.92%, respectively. The low values of fitting quality (χ2 ≤ 1.30) indicate excellent agreement between the data and the refined models.

3.2. Field-Emission Scanning Electron Microscopy (FESEM)

The FESEM images of the MnXCo3-XO4 nanoparticles (0 ≤ X ≤ 1) are shown in Figure 4. A non-uniform morphology was observed, specifically polyhedral-shaped particles, and a few smaller spherical-like particles, mostly agglomerated [38]. Another observation is that as the amount of Mn increases, the morphology of particles tends to be octahedral like. The average particle size distribution was 208 nm for X = 0.0, 162 nm for X = 0.2, 145 nm for X = 0.4, 142 nm for X = 0.6, 140 nm for X = 0.8, and 133 nm for X = 1.0. From the size distribution histograms, it is evident the shift of main sizes to smaller values as the Mn content increased.

3.3. Transmission Electron Microscopy (TEM)

Additional morphological characterization was carried out by the TEM technique. Typical images of the nanoparticles are presented in Figure 5. They show particles with non-uniform morphologies of different sizes. These pictures agree well with the images acquired by FESEM (Figure 4). Figure 5b,e,h,k,n,q shows high-resolution TEM images (5 nm scale) of particles larger than 10 nm in size, with fringes related to atomic planes with spacings of 0.24, 0.296, 0.282, 0.252, 0.303, and 0.486 nm that may be due to the planes (311), (220), (220), (311), (220) and (111) for samples X = 0.0, X = 0.2, X = 0.4, X = 0.6, X = 0.8 and X = 1.0, respectively. Furthermore, it appears that the particles are coated with a carbon layer with a thickness smaller than 5 nm. Figure 5c,f,i,l,o,r shows the small-area electron diffraction (SAED) patterns of the samples. They exhibit diffraction rings originating from crystal planes (111), (220), (311), (400), (422), (333), and (440). The planes are listed beginning from the smallest ring.

3.4. Fourier-Transform Infrared (FT-IR) Spectroscopy

The FT-IR technique shows the vibrational fingerprint of the sample, with absorption peaks that correspond to the frequencies of vibrations of the bonds among the atoms that make up the material [62]. Figure 6 shows the spectra of the MnXCo3-XO4 samples (0 ≤ X ≤ 1) in the range from 400 to 4000 cm−1, where two bands with the highest intensities are located at 552–570 and 643–663 cm−1, which are related to the stretching vibrations of the metal–oxygen bond, which confirms the formation of the pure Co3O4 phase [63]. The v1 band at 552–570 cm−1 is characteristic of the vibration of Co3+ at the octahedral site, and the ν2 band at 643–663 cm−1 is related to the vibration of Co2+ at the tetrahedral site, confirming the formation of the spinel-like oxide [64], in agreement with the XRD study. The low-intensity band that appears at 1100 cm−1 is due to C-O stretching vibrations. The band at 1383 cm−1 is attributed to the symmetric deformations of C-N and CH2 groups, originating from the residues of nitrate ions and agar-agar [65,66,67]. The band at 1635 cm−1 was attributed to the angular deformation of the adsorbed water molecules [68]. The broad absorption band in the region of about 3440 cm−1 is due to OH stretching of the water molecules adsorbed from the moisture during the storage process [69]. Furthermore, it is observed that the bands at 554–643 cm−1 for sample X = 0.8 and the bands at 552–643 cm−1 for sample X = 1.0 are similar, which may be related to the oxidation states of manganese (Mn+2, Mn+3, and Mn+4) or to the vibrational intensity between the manganese and oxygen bond. In general, the frequency of the peaks of the absorption bands at (552–570) and (643–663) cm−1 (Table 2) decreases with the replacement of cobalt with manganese ions, i.e., they shift to the right as the amount of manganese increases, and this is related to the increase in the metal–oxygen distance, as indicated by the increase in the lattice parameter of the unit cells (Table 1), since Mn ions are larger than Co ions [21].

3.5. Ultraviolet–Visible Spectroscopy (UV–Vis)

The electronic properties of MnXCo3-XO4 samples, as illustrated in Figure 7, were investigated by UV–Vis spectroscopy in the wavelength range from 300 to 1400 nm. The absorptions at 528 and 792 nm for the sample X = 0 correspond to the ligand–metal O(-II) → Co(III) and O(-II) → Co(II) electron transfer, respectively [76,77,78]. The variation of absorption in the range from 1033 to 1110 nm shows that when the amount of manganese increases, the absorption peak wavelength increases, and this is related to the higher O2/O1 ratios according to the XPS results, and it will affect the overpotential in the oxygen evolution reaction for the MnCo2O4 (X = 1.0) sample, which would mean a better catalytic activity for the oxidation reactions.

3.6. Raman Spectroscopy

Figure 8 shows the Raman spectra of the MnXCo3-XO4 samples in the range from 100 to 1000 cm−1. The observed bands in the intervals 193–688 cm−1 for x = 0.0, 186–682 cm−1 for x = 0.2, 186–675 cm−1 for x = 0.4, 185–667 cm−1 for x = 0.6, 183–661 cm−1 for x = 0.8, and 182–660 cm−1 for x = 1.0 correspond to the active Raman modes A1g+Eg+3F2g (Table 3), confirming the formation of the pure phase of mixed-valence oxides of spinel-like structure [79,80,81,82,83]. The most intense band at 688–660 cm−1 is assigned to the octahedral site MO6 related to the A1g mode of the O7h spectroscopic symmetry, which corresponds to the stretching vibrational modes of these oxides M-O, where M = {Co, Mn}, thus substantiating the formation of MnCo2O4. The Raman bands with medium intensity in the intervals 468–488 cm−1 and 508–518 cm−1 are assigned to Eg and F2g, respectively; meanwhile, the Raman bands with lower intensities in the interval 603–617 cm−1 are caused by the F2g mode. Moreover, the Raman bands with very low intensity at 182–193 cm−1 are attributed to the F2g mode related to the tetrahedral sites of CoO4 [82,83,84]. In general, when comparing the positions of the peaks, it is observed that as the amount of manganese increases, the peaks shift to the left, analogously to what was noticed in the FT-IR spectra. This change may be due to the greater ionic radius of Mn2+, in comparison to that of Co2+/Co3+ [85,86,87], which when entering the structure of Co3O4, generates a large distortion in the crystalline structure and increases the distance between the metal and the oxygen, and consequently a weakening of bonds occurs. Another reason would be due to vibrations in the structure, where the Co2+ and Co3+ cations are located in tetrahedral and octahedral sites in the cubic crystal structure [88].

3.7. X-ray Photoelectron Spectroscopy (XPS)

The surface oxidation states of the samples were analyzed by XPS. Figure 9 shows the high-resolution Co 2p, Mn 2p, and O 1s spectra obtained from the analysis. All data were corrected for the carbon peak position. In the case of the Co 2p spectra (Figure 9a), four peaks were deconvoluted, at lower binding energies, corresponding to Co3+ and Co2+, as well as two satellite peaks at higher binding energies. The binding energies obtained for Co3+ were found to fall in the 779.64–780.19 eV range, while for Co2+, in the 781.19–781.85 eV range, in agreement with previous work [90]. We also found a Co2+/Co3+ ratio varying from 0.31 to 0.40 among the samples, for which the sample X = 0.8 was found to have the lowest value of Co2+/Co3+ = 0.31. Higher oxidation states can induce more bonded oxygen species, which may have a positive impact on oxidation reactions.
Conversely, in the case of the Mn 2p spectra (Figure 9b), the data were deconvoluted into three peaks, which were ascribed to Mn4+ (ranging from 644.154 eV to 645.146 eV), Mn3+ (ranging from 642.790 eV to 643.125 eV), and Mn2+ (641.223 eV to 641.459 eV), in agreement with previous reports [90,91]. The lowest oxidation state species, Mn2+, was found to represent the largest fraction of the total species present at the surface, with a relative value varying from 0.42 to 0.51 among the samples. Conversely, the presence of Mn4+ and Mn3+ oxidation states is related to the relatively high calcination temperature used in this work, i.e., 900 °C, as found in previous literature [92]. In this respect, we also noted a higher Mn3+/Mn4+ ratio for the tested samples, which correlates well with the previously discussed Jahn–Teller phenomenon, with a concurrent distortion of the crystal lattice.
Finally, the O 1s high-resolution spectra (Figure 9c) display three deconvoluted peaks: O1 (529.94–530.27 eV), O2 (531.16–531.72 eV), and O3 (532.55–533.19 eV). Based on the characteristic binding energies determined for these peaks, they are likely related to surface lattice oxygen (Olat, O2−), adsorbed oxygen species (Oads, O2−, O22−, and O), and adsorbed water species (OH2O), in agreement with earlier reports on similar compounds [91,93]. From the analysis of the O 1s high-resolution spectra (Table 4), we determined higher O2/O1 ratios with increasing Mn content, with a maximum value obtained for the X = 1.0 sample. This suggests that the compounds with the highest Mn content possess increased catalytic activity, as a likely result of increased oxygen-ion vacancies in these samples.

3.8. Vibrating Sample Magnetometer (VSM)

Magnetization measurements were done to study the magnetic behavior of samples at room temperature and to determine the cation magnetic moment in an approximate manner. For all samples, the isothermal magnetization at T = 300 K showed a linear behavior with the magnetic field, and their magnetization at a given field increased with the Mn concentration, as shown in Figure 10.
This trend was observed for all samples and it is typical of paramagnetic samples as shown in Figure 10. From the classical theory of paramagnetism we know that the relationship between the magnetization (M) and magnetic field (H) is given by the Langevin function L(a) = M/Mo = coth(a)-1/a, where M and Mo are the mass magnetizations per total amount of Mn and Co (without oxygen), a = μH/KBT, μ is the average magnetic moment per cation, KB is the Boltzmann constant (1.3807 × 10−16 cm2gK/s2), and T = 300 K. It is known that L(a) tends to a/3 when a is less than about 0.5 [94]. In the present case, if μ~4.51 × 10−20 Erg/Oe (theoretical magnetic moment for Mn3+ and Co3+) and H = 15 × 103 Oe, then, one can get a = 0.01633, which is smaller than 0.5. Thus, L(a) = M/Moa/3 and, therefore, M = [Moμ/(3KBT)]H. In a paramagnetic system Mo = Nμ/A, where N is the Avogadro’s number and A = (x*54.938 + (3-x)*58.933)/3 is the average atomic mass provided by Mn and Co in MnXCo3-XO4 (where X = {0.0, 0.2, 0.4, 0.6, 0.8, 1.0}). Thus, the DC susceptibility is χ = M/H= Nμ2/(3AKBT) [94]. One can study the M-H data by fitting the curve to a linear equation and comparing the slope to Nμ2/(3AKBT). Then, to obtain μ in Bohr magnetons (μB), one has to calculate μ/0.9274 × 10−20. The results provided an effective magnetic moment per cation of 2.204 2.294, 2.342, 2.344, 2.348, 2.378 μB for the samples prepared with x = 0.0, 0.2, 0.4, 0.6, 0.8, 1.0, respectively.
The magnetic moments expected for low spin configuration have a total spin of S(Mn2+) = 0.5, S(Mn3+) = 0, S(Mn4+) = 0.5, S(Co2+) = 0.5, S(Co3+) = 0, whose magnetic moments are, μ = 2 S ( S + 1 ) μB, i.e., μ(Mn2+) = 1.732 μB, μ(Mn3+) = 0, μ(Mn4+) = 1.732 μB, μ(Co2+) = 1.732 μB, μ(Co3+) = 0. Furthermore, for high spin configuration S(Mn2+) = 2.5, S(Mn3+) = 2.0, S(Mn4+) = 1.5, S(Co2+) = 1.5, S(Co3+) = 2, whose magnetic moments are μ(Mn2+) = 5.916 μB, μ(Mn3+) = 4.899 μB, μ(Mn4+) = 3.873 μB, μ(Co2+) = 3.873 μB, μ(Co3+) = 4.899 μB. Therefore, the magnetic moments for Co and Mn seem to be mainly in the low spin configuration; however, we cannot rule out the presence of some moments in the high spin configuration.

3.9. Oxygen Evolution Reaction (OER)

The samples were also evaluated as electrocatalysts for the oxygen evolution reaction (OER). According to the results of the anodic polarization (Figure 11a), the electrodes presented values of 515 (Ni foam), 342 (X = 0.0), 342 (X = 0.2), 339 (X = 0.4), 337 (X = 0.6), 323 (X = 0.8), 299 (X = 1.0) and 235 (RuO2/Ni foam benchmark, extracted from reference [95]) mV vs. RHE, respectively, to record a current density J = 10 mA cm2. Among the investigated materials, the MnCo2O4 (X = 1.0) samples displayed the best catalytic activity for OER, i.e., the lowest overpotential because the incorporation of Mn into the structure enhanced the defect concentrations, thus increasing the amount of catalytically active sites, which facilitated the mass transfer process, favoring OER [96]. Moreover, the crystalline size decreased with the increase of manganese content, which indicates that the Co3O4 sample (X = 0.0) has larger average crystal sizes than the other samples, especially MnCo2O4 (X = 1.0); thus, Co3O4 was the sample that had the highest overpotential. This indicates that the presence of Mn, has a suppressive effect on Co3O4 [97]. The obtained values are in agreement with others reported in the literature for MnXCo3-XO4 nanostructures, as shown in Table 5.
The electrocatalytic kinetics for OER was investigated by the Tafel plots extracted from the LSV (linear sweep voltammetry) curves (Figure 11a), using the Tafel equation (η = a + b log j), where b is the Tafel slope, η is the overpotential, j is the current density, and a is a constant. The values of the Tafel slope (Figure 11b) were 63 (X = 0.0), 73 (X = 0.2), 72 (X = 0.4), 69 (X = 0.6), 68 (X = 0.8), and 52 mV dec−1 (X = 1.0). Therefore, the results did not follow exactly a sequence like the ղ10 values (Figure 11a), but it can be observed that the electrode based on the MnCo2O4 sample (X = 1.0) exhibited the best reaction kinetics for OER, as it showed the lowest Tafel slope, which demonstrates a higher efficiency for oxygen evolution. The Tafel slope of 63 mV dec−1 for the Co3O4 sample (X = 0.0) corresponds to slightly slower kinetics, indicating limitation in charge and mass transfer processes compared to the x = 1.0 sample. The Co3O4 sample was also the one with the highest overpotential, with no distortion in the structure, which reduces defects and consequently the oxygen vacancies [106,107]. These results are consistent with the XPS values as well as the electrochemical impedance spectroscopy. The samples with X = 0.2, 0.4, 0.6, and 0.8 show values next to 70 mV dec−1, which means much slower kinetics for OER.
All this evidence can be explained by the distortion of the lattice with the increase of the amount of Mn in the Co3O4 structure, which changes the electronic charge distribution and increases the disorder in the crystalline system [106,107,108,109]. Furthermore, with increasing Mn percentages, the availability of oxygen and flexibility in the lattice is greater, which in turn is related to the M-O bond length [110]. In any case, the samples (X = 0.2), (X = 0.4), (X = 0.6) and (X = 0.8) show values of Tafel slope below 80 mV dec−1, and these results indicate the adsorption of intermediate species as the rate-determining step (rds), based on the Krasil’shchikov reaction model for OER in alkaline medium [111,112].
The double-layer capacitance (CDL) can be obtained from the relationship between the anode current density (ia) and the scan rate (ʋ), according to (ia = ʋ x CDL) [100]. Figure 11c shows the double-layer capacitance values obtained for the samples: 1.78 (X = 0.0), 2.75 (X = 0.2), 2.57 (X = 0.4), 2.41 (X = 0.6), 3.16 (X = 0.8), and 2.03 mF cm−2 (X = 1.0). These results suggest that the largest number of active sites is organized on the electrode surfaces. Although among the samples of this series, the one with X = 0.8 had the second lowest performance for OER, it displays the highest CDL value, which may be linked to the high amount of oxygen vacancies that improves the absorption of reactive species (like OH¯) [113]. For the samples (X = 0.2), (X = 0.4), and (X = 0.6), the values of 2.75, 2.57, and 2.41 mF cm−2, respectively, are consistent with the XPS data (Figure 9b), where the species in the lowest oxidation state, Mn2+, represented the largest fraction of the total species present on the surface, with a relative fraction ranging from 0.42–0.51 among the samples. However, even with a low CDL value for the sample (X = 1.0), the presence of Mn ions in the structure is essential for superior electrocatalytic properties. This was proven by the best overpotential extracted from the LSV curves (Figure 11a) and the XPS data (Figure 9c), as the substitution of Mn in spinel oxide cobalt occurs selectively in the (Co3+) lattices, and the energy required for Mn2+ to substitute Co3+ is lower than that of Co2+, [114]. Moreover, Mn+2, Mn+3, and Mn+4 have ionic ratios of 0.80, 0.66, and 0.60 Å, respectively, and the ratios of Mn+2 and Mn+3 are larger than that of Co3+ (0.63 Å). Therefore, Mn doping results in the expansion of the Co3O4 lattice, generating defects, which influences the mass diffusion and charge transfer properties, contributing to oxygen-ion vacancies, which are consistent with the O2/O1 ratio that was highest for the MnCo2O4 sample (X = 1.0), with a value of 0.940 [108,109,115,116].
Durability is another important indicator of catalytic performance. The stability of the electrocatalysts was evaluated by chronopotentiometry. Tests were performed at a current density of 10 mA cm−2 for 15 h. According to the curves shown in Figure 11d, it can be seen that the samples (X = 0) and (X = 0.6) exhibited a potential decrease until about 2 h, but then they remained stable, whereas for the samples (X = 0.8) and (X = 1.0), the potential was practically stable for the entire time period tested. In general, all samples showed satisfactory stability over 15 h of testing [117].

3.10. Electrochemical Impedance Spectroscopy (EIS)

The electrocatalytic activity also was investigated by electrochemical impedance spectroscopy (EIS). The EIS spectra of all samples were collected at 1.4 V vs. RHE. As seen in the Bode plots (Figure 12b), the OER is composed of complex processes involving electrosorption of intermediate species during the reaction progress. This suggests an equivalent circuit model (ECM) able to describe these processes [118,119] that is composed of RS (uncompensated solution resistance), RP (polarization resistance, which denotes the overall rate of the OER), QDL (double-layer pseudo-capacitance), R-ad (resistance associated with adsorption of intermediate species), and Q-ad (pseudo-capacitance of these species throughout the reaction). A constant phase element (Q) was used to model an imperfect capacitor, and its impedance was obtained by:
Z Q = ( Q ( i ω ) n ) 1
Then, the values were used to calculate true capacitance (CDL or C-ad) by:
C = R ( 1 n ) / n   Q 1 / n
In Figure 12a, the impedance of the electrodes is composed of two incomplete semicircles. The first is attributed to the polarization process (charge transfer), and the second indicates limitations on mass transfer processes, related to the intermediate species adsorption process [120]. For the electrodes, the obtained RP values were consistent with the OER performance, i.e., the X = 1.0 sample showed the lowest value (6.02 Ω), followed by X = 0.8 (9.70 Ω). The other samples revealed RP values very close, but the result was expected as their overpotential values were close. The CDL values varied slightly (Table 6) due to the oxidation peak shown in Figure 11a. The RadCad loop associated with relaxation, which was attributed to the adsorbed intermediate species, revealed the difficulty of these electrodes to work in the diffusive processes observed at low frequencies. The high R-ad (>1100 Ω) values displayed by the electrodes in those low frequency (>1 Hz) confirm that the adsorption of intermediates should be a rate-limiting step as predicted by Tafel analysis (Figure 11b) [118,121]. The values obtained from the fitting of the spectra are listed in Table 6.

4. Conclusions

The MnXCo3-XO4 (0 ≤ X ≤ 1) samples were synthesized by the proteic sol-gel method (green synthesis) using Agar-Agar as a polymerizing agent in order to investigate their structural, optical, magnetic, and electrochemical properties. X-ray diffraction indicated for all samples the obtainment of the pure cubic phase without any secondary phase, which was also confirmed from Raman, TEM, FT-IR, and UV–Vis studies. Regarding the magnetic measurements, it was observed for all samples a magnetization in a certain field increasing with the Mn concentration, which is typical of a paramagnetic behavior. From the XPS analysis, the species in the Mn2+ oxidation state represented the largest fraction of the total species present on the surface, and as the amount of Mn increased, the O2/O1 ratio also increased, reaching a value of 0.940 for the sample MnCo2O4 (X = 1.0). For OER, the same sample exhibited the best catalytic activity when compared with the others, with an overpotential of 299 mV, which is lower than those of noble metal electrocatalysts reported in the literature. In addition, the samples showed superior long-term stability for efficient water oxidation activities at J = 10 mA/cm2 per 15 h. Thus, it can be concluded that proteic sol-gel synthesis is an excellent method to produce nanosized mixed-valence oxides MnXCo3-XO4 for the fabrication of electrodes for water electrolysis.

Author Contributions

J.R.D.S.: Conceptualization, Investigation, Methodology, Software, Data curation, Writing—original draft preparation. R.R.A.: Software, Data curation, Writing—original draft preparation. T.R.S.: Software, Data curation. V.D.S.: Software, Data curation, Writing—original draft preparation. D.A.M.: Writing—review and editing. F.J.A.L.: Software, Data curation, Writing—original draft preparation. M.A.M.T.: Software, Data curation, Writing—review and editing. D.T.: Writing—review and editing. U.U.G.: Writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

CAPES—Finance code 01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The study did not report any data.

Acknowledgments

The authors thank the Coordination for the Improvement of Higher Education Personnel (CAPES, Finance Code 001). This work was funded by Public Call No. 01/2021 Produtividade em Pesquisa PROPESQ/PRPG/UFPB Proposal Code PVF 14860-2021. Uílame Umbelino Gomes—(CNPQ-Level 1B Research Productivity Fellowship, N.308894/2021-9). Rafael A. Raimundo acknowledges Paraíba State Research Foundation (FAPESQ) for the financial support through the Postdoctoral Fellowship in the country—BLD—PDRP (Public notice N° 07/2021, program 22210.19.573.5011.1680). Francisco J. A. Loureiro expresses his gratitude to the projects CEECIND/02797/2020, PTDC/CTM-CTM/2156/2020, UIDB/00481/2020 and UIDP/00481/2020 from Fundação para a Ciência e a Tecnologia (FCT) and CENTRO-01-0145-FEDER-022083—Centro Portugal Regional Operational Programme (Centro2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF). Vinícius D. Silva thanks the national research Council (CNPq, Process No. 202290/2020-4). Daniel A. Macedo and Marco A. Morales acknowledge CNPq/Brazil (309430/2019-4 and 313337/2018-7) for the financial support given. We thank Tatiane Oliveira dos Santos at LABMIC-UFG for providing TEM images.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roy, A.; Jadhav, H.S.; Cho, M.; Seo, J.G. Electrochemical deposition of self-supported bifunctional copper oxide electrocatalyst for methanol oxidation and oxygen evolution reaction. J. Ind. Eng. Chem. 2019, 76, 515–523. [Google Scholar] [CrossRef]
  2. Sivakumar, P.; Subramanian, P.; Maiyalagan, T.; Gedanken, A.; Schechter, A. Ternary nickel–cobalt–manganese spinel oxide nanoparticles as heterogeneous electrocatalysts for oxygen evolution and oxygen reduction reaction. Mater. Chem. Phys. 2019, 229, 190–196. [Google Scholar] [CrossRef]
  3. Jiang, B.; Kim, J.; Guo, Y.; Wu, K.C.W.; Alshehri, S.M.; Ahamad, T.; Alhokbany, N.; Henzie, J.; Yamachi, Y. Efficient oxygen evolution on mesoporous IrO:X nanosheets. Catal. Sci. Technol. 2019, 9, 3697–3702. [Google Scholar] [CrossRef]
  4. Tu, Q.; Mo, J.; Liu, Z.; Gong, C.; Fan, Y. Using green finance to counteract the adverse effects of COVID-19 pandemic on renewable energy investment-The case of offshore wind power in China. Energy Policy 2021, 158, 112542. [Google Scholar] [CrossRef] [PubMed]
  5. Iskandarova, M.; Dembek, A.; Fraaije, M.; Matthews, W.; Stasik, A.; Wittmayer, J.M.; Sovacool, B.K. Who finances renewable energy in Europe? Examining temporality, authority and contestation in solar and wind subsidies in Poland, the Netherlands and the United Kingdom. Energy Strateg. Rev. 2021, 38, 100730. [Google Scholar] [CrossRef]
  6. Cousse, J. Still in love with solar energy? Installation size, affect, and the social acceptance of renewable energy technologies. Renew. Sustain. Energy Rev. 2021, 145, 111107. [Google Scholar] [CrossRef]
  7. Sayed, E.T.; Wilberforce, T.; Elsaid, K.; Rabaia, M.K.H.; Abdelkareem, M.A.; Chae, K.-J.; Olabi, A. A critical review on environmental impacts of renewable energy systems and mitigation strategies: Wind, hydro, biomass and geothermal. Sci. Total Environ. 2021, 766, 144505. [Google Scholar] [CrossRef]
  8. Rashidi, S.; Karimi, N.; Sunden, B.; Kim, K.C.; Olabi, A.G.; Mahian, O. Progress and challenges on the thermal management of electrochemical energy conversion and storage technologies: Fuel cells, electrolysers, and supercapacitors. Prog. Energy Combust. Sci. 2022, 88, 100966. [Google Scholar] [CrossRef]
  9. Tomon, C.; Sarawutanukul, S.; Phattharasupakun, N.; Duangdangchote, S.; Chomkhundtod, P.; Kidkhunthod, P.; Sawangphruk, M. Insight into photoelectrocatalytic mechanisms of bifunctional cobaltite hollow-nanofibers towards oxygen evolution and oxygen reduction reactions for high-energy zinc-air batteries. Electrochim. Acta 2021, 392, 139022. [Google Scholar] [CrossRef]
  10. Vazhayil, A.; Vazhayal, L.; Thomas, J.; Ashok, C.S.; Thomas, N. A comprehensive review on the recent developments in transition metal-based electrocatalysts for oxygen evolution reaction. Appl. Surf. Sci. Adv. 2021, 6, 100184. [Google Scholar] [CrossRef]
  11. Anwar, S.; Khan, F.; Zhang, Y.; Djire, A. Recent development in electrocatalysts for hydrogen production through water electrolysis. Int. J. Hydrogen Energy 2021, 46, 32284–32317. [Google Scholar] [CrossRef]
  12. Lourenço, A.D.A.; Silva, V.D.; da Silva, R.; Silva, U.C.; Chesman, C.; Salvador, C.; Simões, T.; de Macedo, D.A.; da Silva, F.F. Metal-organic frameworks as template for synthesis of Mn3+/Mn4+ mixed valence manganese cobaltites electrocatalysts for oxygen evolution reaction. J. Colloid Interface Sci. 2021, 582, 124–136. [Google Scholar] [CrossRef] [PubMed]
  13. Song, X.-Z.; Zhang, N.; Liu, F.; Wang, Z.-H.; Zhu, W.-Y.; Zhang, G.-Z.; Niu, Z.-Y.; Wang, X.-F.; Tan, Z. Spontaneously engineering heterogeneous interface of silver nanoparticles on α-Co(OH)2 for boosting electrochemical oxygen evolution. J. Alloys Compd. 2021, 873, 159766. [Google Scholar] [CrossRef]
  14. Zhang, T.; Meng, Y.-L.; Zhao, Y.-H.; Ni, J.-C.; Pan, Y.; Dai, Y.; Tan, Z.; Wang, X.-F.; Song, X.-Z. Boosting oxygen evolution electrocatalysis of high-entropy hydroxide by high-valency nickel regulation. Chem. Commun. 2022, 58, 7682–7685. [Google Scholar] [CrossRef]
  15. Jamesh, M.I.; Sun, X. Recent progress on earth abundant electrocatalysts for oxygen evolution reaction (OER) in alkaline medium to achieve efficient water splitting—A review. J. Power Sources 2018, 400, 31–68. [Google Scholar] [CrossRef]
  16. Lankauf, K.; Cysewska, K.; Karczewski, J.; Mielewczyk-Gryń, A.; Górnicka, K.; Cempura, G.; Chen, M.; Jasiński, P.; Molin, S. MnxCo3−xO4 spinel oxides as efficient oxygen evolution reaction catalysts in alkaline media. Int. J. Hydrogen Energy 2020, 45, 14867–14879. [Google Scholar] [CrossRef]
  17. Tao, L.; Guo, P.; Zhu, W.; Li, T.; Zhou, X.; Fu, Y.; Yu, C.; Ji, H. Highly efficient mixed-metal spinel cobaltite electrocatalysts for the oxygen evolution reaction. Chin. J. Catal. 2020, 41, 1855–1863. [Google Scholar] [CrossRef]
  18. Liu, Q.; Chen, Z.; Yan, Z.; Wang, Y.; Wang, E.; Wang, S.; Wang, S.; Sun, G. Crystal-Plane-Dependent Activity of Spinel Co3O4 Towards Water Splitting and the Oxygen Reduction Reaction. ChemElectroChem 2018, 5, 1080–1086. [Google Scholar] [CrossRef]
  19. Leng, M.; Huang, X.; Xiao, W.; Ding, J.; Liu, B.; Du, Y.; Xue, J. Enhanced oxygen evolution reaction by Co-O-C bonds in rationally designed Co3O4/graphene nanocomposites. Nano Energy 2017, 33, 445–452. [Google Scholar] [CrossRef]
  20. Lee, E.; Jang, J.H.; Kwon, Y.U. Composition effects of spinel MnxCo3−xO4 nanoparticles on their electrocatalytic properties in oxygen reduction reaction in alkaline media. J. Power Sources 2015, 273, 735–741. [Google Scholar] [CrossRef]
  21. Rios, E.; Chartier, P.; Gautier, J.-L. Oxygen evolution electrocatalysis in alkaline medium at thin MnxCo3−xO4 (0 ≤ x ≤ 1) spine1 films on glass/SnO2: F prepared by spray pyrolysis. Solid State Sci. 1999, 1, 267–277. [Google Scholar] [CrossRef]
  22. Shi, C.; Ullah, S.; Li, K.; Wang, W.; Zhang, R.; Pan, L.; Zhang, X.; Zou, J.-J. Low-temperature synthesis of ultrasmall spinel MnxCo3−xO4 nanoparticles for efficient oxygen reduction. Chin. J. Catal. 2020, 41, 1818–1825. [Google Scholar] [CrossRef]
  23. Shen, Z.; Rong, J.; Yu, X. MnxCo3−xO4 spinel coatings: Controlled synthesis and high temperature oxidation resistance behavior. Ceram. Int. 2020, 46, 5821–5827. [Google Scholar] [CrossRef]
  24. Li, D.; Kang, Z.; Sun, H.; Wang, Y.; Xie, H.; Liu, J.; Zhu, J. A bifunctionalMnxCo3−xO4-decorated separator for efficient Li-LiI-O2 batteries: A novel strategy to promote redox coupling and inhibit redox shuttling. Chem. Eng. J. 2022, 428, 131105. [Google Scholar] [CrossRef]
  25. Wang, Y.; Yang, Y.; Jia, S.; Wang, X.; Lyu, K.; Peng, Y.; Zheng, H.; Wei, X.; Ren, H.; Xiao, L.; et al. Synergistic Mn-Co catalyst outperforms Pt on high-rate oxygen reduction for alkaline polymer electrolyte fuel cells. Nat. Commun. 2019, 10, 1506. [Google Scholar] [CrossRef]
  26. Talic, B.; Hendriksen, P.V.; Wiik, K.; Lein, H.L. Thermal expansion and electrical conductivity of Fe and Cu doped MnCo2O4 spinel. Solid State Ion. 2018, 326, 90–99. [Google Scholar] [CrossRef]
  27. Zhao, Q.; Yan, Z.; Chen, C.; Chen, J. Spinels: Controlled Preparation, Oxygen Reduction/Evolution Reaction Application, and beyond. Chem. Rev. 2017, 117, 10121–10211. [Google Scholar] [CrossRef]
  28. Mandal, B.; Mitra, P. Grain growth correlated complex impedance spectroscopy, modulus spectroscopy and carrier hopping mechanism in MnCo2O4, Influence of sintering temperature. Mater. Chem. Phys. 2020, 251, 123095. [Google Scholar] [CrossRef]
  29. Tajik, S.; Dubal, D.P.; Gomez-Romero, P.; Yadegari, A.; Rashidi, A.; Nasernejad, B.; Inamuddin; Asiri, A.M. Nanostructured mixed transition metal oxides for high performance asymmetric supercapacitors: Facile synthetic strategy. Int. J. Hydrogen Energy 2017, 42, 12384–12395. [Google Scholar] [CrossRef]
  30. Yang, H.; Hu, F.; Zhang, Y.; Shi, L.; Wang, Q. Controlled synthesis of porous spinel cobalt manganese oxides as efficient oxygen reduction reaction electrocatalysts. Nano Res. 2016, 9, 207–213. [Google Scholar] [CrossRef]
  31. Lavela, P.; Tirado, J.L.; Vidal-Abarca, C. Sol-gel preparation of cobalt manganese mixed oxides for their use as electrode materials in lithium cells. Electrochim. Acta 2007, 52, 7986–7995. [Google Scholar] [CrossRef]
  32. Park, K.R.; Jeon, J.E.; Ali, G.; Ko, Y.-H.; Lee, J.; Han, H.; Mhin, S. Oxygen evolution reaction of Co-Mn-O electrocatalyst prepared by solution combustion synthesis. Catalysts 2019, 9, 564. [Google Scholar] [CrossRef]
  33. Zhao, T.; Gadipelli, S.; He, G.; Ward, M.J.; Do, D.; Zhang, P.; Guo, Z. Tunable Bifunctional Activity of MnxCo3−xO4 Nanocrystals Decorated on Carbon Nanotubes for Oxygen Electrocatalysis. ChemSusChem 2018, 11, 1295–1304. [Google Scholar] [CrossRef]
  34. Wei, C.; Feng, Z.; Scherer, G.G.; Barber, J.; Shao-Horn, Y.; Xu, Z.J. Cations in Octahedral Sites: A Descriptor for Oxygen Electrocatalysis on Transition-Metal Spinels. Adv. Mater. 2017, 29, 1606800. [Google Scholar] [CrossRef]
  35. Bobruk, M.; Brylewska, K.; Durczak, K.; Wojciechowski, K.; Adamczyk, A.; Brylewski, T. Synthesis of manganese-cobalt spinel via wet chemistry methods and its properties. Ceram. Int. 2017, 43, 15597–15609. [Google Scholar] [CrossRef]
  36. Ríos, E.; Peña, O.; Guizouarn, T.; Gautier, J.-L. Thin films of Co3O4, MnCo2O4 and their solid solution as electrocatalyst: Study of their magnetic properties. Phys. Status Solidi 2004, 1, S108–S111. [Google Scholar] [CrossRef]
  37. Rebekah, A.; Anantharaj, S.; Viswanthan, C.; Ponpandian, N. ScienceDirect Zn-substituted MnCo2O4 nanostructure anchored over rGO for boosting the electrocatalytic performance towards methanol oxidation and oxygen evolution reaction (OER). Int. J. Hydrogen Energy 2020, 45, 14713–14727. [Google Scholar] [CrossRef]
  38. Itteboina, R.; Sau, T.K. Sol-gel synthesis and characterizations of morphology-controlled Co3O4 particles. Mater. Today Proc. 2019, 9, 458–467. [Google Scholar] [CrossRef]
  39. Guo, J.; Chen, L.; Zhang, X.; Jiang, B.; Ma, L. Sol-gel synthesis of mesoporous Co3O4 octahedra toward high-performance anodes for lithium-ion batteries. Electrochim. Acta 2014, 129, 410–415. [Google Scholar] [CrossRef]
  40. Silva, R.M.; Raimundo, R.A.; Fernandes, W.V.; Torres, S.M.; Silva, V.D.; Grilo, J.P.; Morales, M.A.; Macedo, D.A. Proteic sol-gel synthesis, structure and magnetic properties of Ni/NiO core-shell powders. Ceram. Int. 2018, 44, 6152–6156. [Google Scholar] [CrossRef]
  41. Amri, A.; Jiang, Z.T.; Pryor, T.; Yin, C.Y.; Djordjevic, S. Developments in the synthesis of flat plate solar selective absorber materials via sol-gel methods: A review. Renew. Sustain. Energy Rev. 2014, 36, 316–328. [Google Scholar] [CrossRef]
  42. Nair, P.A.K.; Vasconcelos, W.L.; Paine, K.; Calabria-Holley, J. A review on applications of sol-gel science in cement. Constr. Build. Mater. 2021, 291, 123065. [Google Scholar] [CrossRef]
  43. da Silva, M.V.; Fajardo, H.V.; Rodrigues, T.S.; e Silva, F.A.; Bergamaschi, V.S.; Dias, A.; Siqueira, K.P. Synthesis of NiMoO4 ceramics by proteic sol-gel method and investigation of their catalytic properties in hydrogen production. Mater. Chem. Phys. 2021, 262, 124301. [Google Scholar] [CrossRef]
  44. Pimentel, P.M.; Lima, S.V.M.; Costa, A.F.; Câmara, M.S.C.; Carregosa, J.D.C.; Oliveira, R.M.P.B. Gelatin synthesis and color properties of (La, Pr, Nd) lanthanide aluminates. Ceram. Int. 2017, 43, 6592–6596. [Google Scholar] [CrossRef]
  45. Santos, J.R.; Loureiro, F.J.; Grilo, J.P.; Silva, V.D.; Simões, T.A.; Fagg, D.P.; Macedo, D.A. Understanding the cathodic polarisation behaviour of the misfit [Ca2CoO3−δ]q[CoO2] (C349) as oxygen electrode for IT-SOFC. Electrochim. Acta 2018, 285, 214–220. [Google Scholar] [CrossRef]
  46. Ferreira, L.S.; Silva, T.R.; Silva, V.D.; Simões, T.A.; Araújo, A.J.; Morales, M.A.; Macedo, D.A. Proteic sol-gel synthesis, structure and battery-type behavior of Fe-based spinels (MFe2O4, M = Cu, Co, Ni). Adv. Powder Technol. 2020, 31, 604–613. [Google Scholar] [CrossRef]
  47. Santos, J.V.A.; Macedo, M.A.; Cunha, F.; Sasaki, J.M.; Duque, J.G.S. BaFe12O19 thin film grown by an aqueous sol-gel process. Microelectron. J. 2003, 34, 565–567. [Google Scholar] [CrossRef]
  48. Ferreira, L.S.; Silva, T.R.; Santos, J.R.; Silva, V.D.; Raimundo, R.A.; Morales, M.A.; Macedo, D.A. Structure, magnetic behavior and OER activity of CoFe2O4 powders obtained using agar-agar from red seaweed (Rhodophyta). Mater. Chem. Phys. 2019, 237, 121847. [Google Scholar] [CrossRef]
  49. Rees, D.A. Structure, Conformation, and Mechanism in the Formation of Polysaccharide Gels and Networks. Adv. Carbohydr. Chem. Biochem. 1969, 24, 267–332. [Google Scholar] [CrossRef]
  50. Rhein-Knudsen, N.; Ale, M.T.; Ajalloueian, F.; Yu, L.; Meyer, A.S. Rheological properties of agar and carrageenan from Ghanaian red seaweeds. Food Hydrocoll. 2017, 63, 50–58. [Google Scholar] [CrossRef]
  51. Kartik, A.; Akhil, D.; Lakshmi, D.; Gopinath, K.P.; Arun, J.; Sivaramakrishnan, R.; Pugazhendhi, A. A critical review on production of biopolymers from algae biomass and their applications. Bioresour. Technol. 2021, 329, 124868. [Google Scholar] [CrossRef] [PubMed]
  52. Malagurski, I.; Levic, S.; Nesic, A.; Mitric, M.; Pavlovic, V.; Dimitrijevic-Brankovic, S. Mineralized agar-based nanocomposite films: Potential food packaging materials with antimicrobial properties. Carbohydr. Polym. 2017, 175, 55–62. [Google Scholar] [CrossRef]
  53. Francavilla, M.; Pineda, A.; Lin, C.S.; Franchi, M.; Trotta, P.; Romero, A.A.; Luque, R. Natural porous agar materials from macroalgae. Carbohydr. Polym. 2013, 92, 1555–1560. [Google Scholar] [CrossRef]
  54. Arnott, S.; Fulmer, A.; Scott, W.E.; Dea, I.C.M.; Moorhouse, R.; Rees, D.A. The Agarose Double Helix and Its Function in Agarose Gel Structure. J. Mol. Biol. 1974, 90, 269–284. [Google Scholar] [CrossRef]
  55. Pradhan, B.; Bhuyan, P.P.; Patra, S.; Nayak, R.; Behera, P.K.; Behera, C.; Behera, A.K.; Ki, J.-S.; Jena, M. Beneficial effects of seaweeds and seaweed-derived bioactive compounds: Current evidence and future prospective. Biocatal. Agric. Biotechnol. 2022, 39, 102242. [Google Scholar] [CrossRef]
  56. Picard, J.P.; Baud, G.; Besse, J.P.; Chevalier, R. Croissance cristalline et étude structurale de Co3O4. J. Less-Common Met. 1980, 75, 99–104. [Google Scholar] [CrossRef]
  57. Dang, W.; Tang, X.; Wang, W.; Yang, Y.; Li, X.; Huang, L.; Zhang, Y. Micro-nano NiO-MnCo2O4heterostructure with optimal interfacial electronic environment for high performance and enhanced lithium storage kinetics. Dalton Trans. 2020, 49, 10994–11004. [Google Scholar] [CrossRef]
  58. Restovic, A.; Ríos, E.; Barbato, S.; Ortiz, J.; Gautier, J. Oxygen reduction in alkaline medium at thin MnxCo3−xO4 (0 ≤ x ≤ 1) spinel films prepared by spray pyrolysis. Effect of oxide cation composition on the reaction kinetics. J. Electroanal. Chem. 2002, 522, 141–151. [Google Scholar] [CrossRef]
  59. Raj, S.; Anantharaj, S.; Kundu, S.; Roy, P. In Situ Mn-Doping-Promoted Conversion of Co(OH)2 to Co3O4 as an Active Electrocatalyst for Oxygen Evolution Reaction. ACS Sustain. Chem. Eng. 2019, 7, 9690–9698. [Google Scholar] [CrossRef]
  60. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. In Situ DRIFTs Inestigation of the Low-Temperature Reaction Mechanism over Mn-Doped Co3O4 for the Selective Catalytic Reduction of NO with NH. J. Phys. Chem. C 2015, 119, 22924–22933. [Google Scholar] [CrossRef]
  61. Wyckoff, R.W.G. Crystal Structures, 1st ed.; Interseience: New York, NY, USA; London, UK; Sydney, Australia, 1963; Volume 1. [Google Scholar] [CrossRef]
  62. Dutta, A. Fourier Transform Infrared Spectroscopy; Elsevier Inc.: Amsterdam, The Netherlands, 2017; Volume 2. [Google Scholar] [CrossRef]
  63. Ai, L.H.; Jiang, J. Rapid synthesis of nanocrystalline Co3O4 by a microwave-assisted combustion method. Powder Technol. 2009, 195, 11–14. [Google Scholar] [CrossRef]
  64. Yin, J.; Cao, H.; Lu, Y. Self-assembly into magnetic Co3O4 complex nanostructures as peroxidase. J. Mater. Chem. 2012, 22, 527–534. [Google Scholar] [CrossRef]
  65. Tholkappiyan, R.; Naveen, A.N.; Sumithra, S.; Vishista, K. Investigation on spinel MnCo2O4 electrode material prepared via controlled and uncontrolled synthesis route for supercapacitor application. J. Mater. Sci. 2015, 50, 5833–5843. [Google Scholar] [CrossRef]
  66. Petit, S.; Madejova, J. Fourier Transform Infrared Spectroscopy, 2nd ed.; Elsevier Ltd.: Amsterdam, The Netherlands, 2013; Volume 5. [Google Scholar] [CrossRef]
  67. Basumatary, K.; Daimary, P.; Das, S.K.; Thapa, M.; Singh, M.; Mukherjee, A.; Kumar, S. Lagerstroemia speciosa fruit-mediated synthesis of silver nanoparticles and its application as filler in agar based nanocomposite films for antimicrobial food packaging. Food Packag. Shelf Life 2018, 17, 99–106. [Google Scholar] [CrossRef]
  68. Naveen, A.N.; Selladurai, S. Investigation on physiochemical properties of Mn substituted spinel cobalt oxide for supercapacitor applications. Electrochim. Acta 2014, 125, 404–414. [Google Scholar] [CrossRef]
  69. Xie, L.; Li, K.; Sun, G.; Hu, Z.; Lv, C.; Wang, J.; Zhang, C. Preparation and electrochemical performance of the layered cobalt oxide (Co3O4) as supercapacitor electrode material. J. Solid State Electrochem. 2013, 17, 55–61. [Google Scholar] [CrossRef]
  70. Habibi, M.H.; Bagheri, P. Enhanced photo-catalytic degradation of naphthol blue black on nano-structure MnCo2O4, charge separation of the photo-generated electron–hole pair. J. Mater. Sci. Mater. Electron. 2017, 28, 289–294. [Google Scholar] [CrossRef]
  71. Uke, S.J.; Chaudhari, G.N.; Kumar, Y.; Mardikar, S.P. Tri-Ethanolamine-Ethoxylate assisted hydrothermal synthesis of nanostructured MnCo2O4 with superior electrochemical performance for high energy density supercapacitor application. Mater. Today Proc. J. 2021, 43, 2792–2799. [Google Scholar] [CrossRef]
  72. Mu, J.; Zhang, L.; Zhao, M.; Wang, Y. Co3O4 nanoparticles as an efficient catalase mimic: Properties, mechanism and its electrocatalytic sensing application for hydrogen peroxide. J. Mol. Catal. A Chem. 2013, 378, 30–37. [Google Scholar] [CrossRef]
  73. Deng, J.; Cheng, Y.-Q.; Lu, Y.-A.; Crittenden, J.C.; Zhou, S.-Q.; Gao, N.-Y.; Li, J. Mesoporous Manganese Cobaltite Nanocages as Effective and Reusable Heterogeneous Peroxymonosulfate Activators for Carbamazepine Degradation. Chem. Eng. J. 2017, 330, 505–517. [Google Scholar] [CrossRef]
  74. Mohamed, M.A.; Jaafar, J.; Ismail, A.F.; Othman, M.H.D.; Rahman, M.A. Fourier Transform Infrared (FTIR) Spectroscopy; Elsevier B.V.: Amsterdam, The Netherlands, 2017. [Google Scholar] [CrossRef]
  75. Rojas, R.M.; Vila, E.; García, O.; de Vidales, J.L.M. Thermal Behaviour and Reactivity of Manganese Cobaltites MnxCo3-xO4 (0.0 < x < 1.0) obtained at Low Temperature. J. Mater. Chem. 1994, 4, 1635–1639. [Google Scholar] [CrossRef]
  76. Zhang, N.; Shi, J.; Mao, S.S.; Guo, L. Co3O4 quantum dots: Reverse micelle synthesis and visible-light-driven photocatalytic overall water splitting. Chem. Commun. 2014, 50, 2002–2004. [Google Scholar] [CrossRef] [PubMed]
  77. Barreca, D.; Massignan, C.; Daolio, S.; Fabrizio, M.; Piccirillo, C.; Armelao, L.; Tondello, E. Composition and Microstructure of Cobalt Oxide Thin Films Obtained from a Novel Cobalt (II) Precursor by Chemical Vapor Deposition. Chem. Mater. 2001, 13, 588–593. [Google Scholar] [CrossRef]
  78. Cook, J.G.; Meer, M.P.V.D. The Optical Properties of Sputtered Co3O4 Films. Thin Solid Films 1986, 144, 165–176. [Google Scholar] [CrossRef]
  79. SSun, C.; Yang, J.; Dai, Z.; Wang, X.; Zhang, Y.; Li, L.; Chen, P.; Huang, W.; Dong, X. Nanowires assembled from MnCo2O4@C nanoparticles for water splitting and all-solid-state supercapacitor. Nano Res. 2016, 9, 1300–1309. [Google Scholar] [CrossRef]
  80. Saravanakumar, B.; Ravi, G.; Ganesh, V.; Guduru, R.K.; Yuvakkumar, R. MnCo2O4 nanosphere synthesis for electrochemical applications. Mater. Sci. Energy Technol. 2018, 2, 130–138. [Google Scholar] [CrossRef]
  81. Hadjiev, V.G.; Iliev, M.; Vergilov, I.V. The Raman spectra of Co3O4. J. Phys. C Solid State Phys. 1988, 21, L199–L201. [Google Scholar] [CrossRef]
  82. Jiang, J.; Li, L. Synthesis of sphere-like Co3O4 nanocrystals via a simple polyol route. Mater. Lett. 2007, 61, 4894–4896. [Google Scholar] [CrossRef]
  83. Zhao, Y.; Li, Y.; Ma, C.; Shao, Z. Micro-/nano-structured hybrid of exfoliated graphite and Co3O4 nanoparticles as high-performance anode material for Li-ion batteries. Electrochim. Acta 2016, 213, 98–106. [Google Scholar] [CrossRef]
  84. Du, S.; Ren, Z.; Qu, Y.; Wu, J.; Xi, W.; Zhu, J.; Fu, H. Co3O4 nanosheets as the high-performance catalyst for oxygen evolution proceeding via the twice two-electron process. Chem. Commun. 2016, 52, 6705–6708. [Google Scholar] [CrossRef]
  85. Islam, M.N.; Pradhan, A.; Kumar, S. Effects of crystallite size distribution on the Raman-scattering profiles of silicon nanostructures. J. Appl. Phys. 2005, 98, 024309. [Google Scholar] [CrossRef]
  86. Zhang, S.Y.; Yang, Y.Y.; Zheng, Y.Q.; Zhu, H.L. Ag-doped Co3O4 catalyst derived from heterometallic MOF for syngas production by electrocatalytic reduction of CO2 in water. J. Solid State Chem. 2018, 263, 44–51. [Google Scholar] [CrossRef]
  87. Kim, N.-I.; Sa, Y.J.; Cho, S.-H.; So, I.; Kwon, K.; Joo, S.H.; Park, J.-Y. Enhancing Activity and Stability of Cobalt Oxide Electrocatalysts for the Oxygen Evolution Reaction via Transition Metal Doping. J. Electrochem. Soc. 2016, 163, F3020–F3028. [Google Scholar] [CrossRef]
  88. Wang, G.; Shen, X.; Horvat, J.; Wang, B.; Liu, H.; Wexler, D.; Yao, J. Hydrothermal synthesis and optical, magnetic, and supercapacitance properties of nanoporous cobalt oxide nanorods. J. Phys. Chem. C 2009, 113, 4357–4361. [Google Scholar] [CrossRef]
  89. Alvarez, A.; Ivanova, S.; Centeno, M.A.; Odriozola, J.A. Sub-ambient CO oxidation over mesoporous Co3O4: Effect of morphology on its reduction behavior and catalytic performance. Appl. Catal. A Gen. 2012, 431–432, 9–17. [Google Scholar] [CrossRef]
  90. Tu, S.; Chen, Y.; Zhang, X.; Yao, J.; Wu, Y.; Wu, H.; Zhang, J.; Wang, J.; Mu, B.; Li, Z.; et al. Complete catalytic oxidation of formaldehyde at room temperature on MnxCo3−xO4 catalysts derived from metal-organic frameworks. Appl. Catal. A Gen. 2021, 611, 117975. [Google Scholar] [CrossRef]
  91. Park, K.R.; Jeon, J.E.; Kim, K.; Oh, N.; Ko, Y.H.; Lee, J.; Lee, S.H.; Ryu, J.H.; Han, H.; Mhin, S. Synthesis of rod-type Co2.4Mn0.6O4 via oxalate precipitation for water splitting catalysts. Appl. Surf. Sci. 2020, 510, 145390. [Google Scholar] [CrossRef]
  92. Gautier, J.L.; Rios, E.; Gracia, M.; Marco, J.F.; Gancedo, J.R. Characterisation by X-ray photoelectron spectroscopy of thin ž/Mn Co O 1 G x G 0 spinel films prepared by low-temperature spray x 3yx 4 pyrolysis. Thin Solid Films 1997, 311, 51–57. [Google Scholar] [CrossRef]
  93. Tang, W.; Wang, S.; Xiao, W.; Du, S.; Lu, X.; Hoang, S.; Ding, J.; Gao, P.-X. Pre-surface leached cordierite honeycombs for MnxCo3-xO4 nano-sheet array integration with enhanced hydrocarbons combustion. Catal. Today 2019, 320, 196–203. [Google Scholar] [CrossRef]
  94. Cullity, B.D.; Graham, C.D. Introduction to Magnetic Materials, 2nd ed.; IEEE: Piscataway, NJ, USA, 2008; Volume 59. [Google Scholar]
  95. Xue, Z.; Liu, K.; Liu, Q.; Li, Y.; Li, M.; Su, C.-Y.; Ogiwara, N.; Kobayashi, H.; Kitagawa, H.; Liu, M.; et al. Missing-linker metal-organic frameworks for oxygen evolution reaction. Nat. Commun. 2019, 10, 5048. [Google Scholar] [CrossRef]
  96. Shit, S.; Bolar, S.; Murmu, N.C.; Kuila, T. An account of the strategies to enhance the water splitting efficiency of noble-metal-free electrocatalysts. J. Energy Chem. 2021, 59, 160–190. [Google Scholar] [CrossRef]
  97. Li, J.; Liang, X.; Xu, S.; Hao, J. Catalytic performance of manganese cobalt oxides on methane combustion at low temperature. Appl. Catal. B Environ. 2009, 90, 307–312. [Google Scholar] [CrossRef]
  98. Menezes, P.W.; Indra, A.; Sahraie, R.; Bergmann, A. Cobalt–Manganese-Based Spinels as Multifunctional Materials that Unify Catalytic Water Oxidation and Oxygen Reduction Reactions. ChemSusChem 2015, 8, 164–171. [Google Scholar] [CrossRef]
  99. Anne, R.; Acedera, E.; Gupta, G.; Mamlouk, M.; Donnabelle, M.; Balela, L. Solution combustion synthesis of porous Co3O4 nanoparticles as oxygen evolution reaction (OER) electrocatalysts in alkaline medium. J. Alloys Compd. 2020, 836, 154919. [Google Scholar] [CrossRef]
  100. Yu, Y.; Zhang, J.; Zhong, M.; Guo, S. Co3O4 Nanosheet Arrays on Ni Foam as Electrocatalyst for Oxygen Evolution Reaction. Electrocatalysis 2018, 9, 653–661. [Google Scholar] [CrossRef]
  101. Menezes, P.W.; Indra, A.; Gutkin, V.; Driess, M. Boosting electrochemical water oxidation through replacement of Oh Co sites in cobalt oxide spinel with manganese. Chem. Commun. 2017, 53, 8018–8021. [Google Scholar] [CrossRef]
  102. Alqahtani, D.M.; Zequine, C.; Ranaweera, C.; Siam, K.; Kahol, P.K.; Poudel, T.P.; Mishra, S.; Gupta, R.K. Effect of metal ion substitution on electrochemical properties of cobalt oxide. J. Alloys Compd. 2019, 771, 951–959. [Google Scholar] [CrossRef]
  103. Ma, Y.; Zha, M.; Dong, Y.; Li, L.; Hu, G. Mn-doped Co3O4 nanoarrays as a promising electrocatalyst for oxygen evolution reaction Mn-doped Co3O4 nanoarrays as a promising electrocatalyst for oxygen evolution reaction. Mater. Res. Express 2019, 6, 115033. [Google Scholar] [CrossRef]
  104. Wang, W.; Kuai, L.; Cao, W.; Huttula, M.; Ollikkala, S.; Ahopelto, T.; Honkanen, A.-P.; Huotari, S.; Yu, M.; Geng, B. Mass-production of Mesoporous MnCo2O4 Spinel with MnIV- and CoII-rich Surface for Superior Bifunctional Oxygen Electrocatalysis. Angew. Chem. Int. Ed. 2017, 47, 14977–14981. [Google Scholar] [CrossRef]
  105. Natarajan, S.; Anantharaj, S.; Tayade, R.J.; Bajaj, H.C.; Kundu, S. Recovered Spinel MnCo2O4 from Spent Lithium Ion Batteries for Enhanced Electrocatalytic Oxygen Evolution in Alkaline Medium. Dalton Trans. 2017, 46, 14382–14392. [Google Scholar] [CrossRef]
  106. Yan, D.; Li, Y.; Huo, J.; Chen, R.; Dai, L.; Wang, S. Defect Chemistry of Nonprecious-Metal Electrocatalysts for Oxygen Reactions. Adv. Mater. 2017, 29, 1606459. [Google Scholar] [CrossRef] [PubMed]
  107. Cai, Z.; Bi, Y.; Hu, E.; Liu, W.; Dwarica, N.; Tian, Y.; Li, X.; Kuang, Y.; Li, Y.; Yang, X.; et al. Single-Crystalline Ultrathin Co3O4 Nanosheets with Massive Vacancy Defects for Enhanced Electrocatalysis. Adv. Energy Mater. 2018, 8, 1701694. [Google Scholar] [CrossRef]
  108. Zhang, W.; Li, M.; Wang, X.; Zhang, X.; Niu, X.; Zhu, Y. Boosting catalytic toluene combustion over Mn doped Co3O4 spinel catalysts: Improved mobility of surface oxygen due to formation of Mn-O-Co bonds. Appl. Surf. Sci. 2022, 590, 153140. [Google Scholar] [CrossRef]
  109. Qin, C.; Wang, B.; Wang, Y. Metal-organic frameworks-derived Mn-doped Co3O4 porous nanosheets and enhanced CO sensing performance. Sens. Actuators B Chem. 2022, 351, 130943. [Google Scholar] [CrossRef]
  110. Biswas, S.; Dutta, B.; Mannodi-Kanakkithodi, A.; Clarke, R.; Song, W.; Ramprasad, R.; Suib, S.L. Heterogeneous mesoporous manganese/cobalt oxide catalysts for selective oxidation of 5-hydroxymethylfurfural to 2,5-diformylfuran. Chem. Commun. 2017, 53, 11751–11754. [Google Scholar] [CrossRef]
  111. Hamdani, M.; Singh, R.N.; Chartier, P. Co3O4 and Co-Based Spinel Oxides Bifunctional Oxygen Electrodes. Int. J. Electrochem. Sci. 2010, 5, 556–577. [Google Scholar]
  112. Bockris, J.O.M.; Otagawa, T. Mechanism of oxygen evolution on perovskites. J. Phys. Chem. 1983, 87, 2960–2971. [Google Scholar] [CrossRef]
  113. Li, G.; Anderson, L.; Chen, Y.; Pan, M.; Chuang, P.-Y.A. New insights into evaluating catalyst activity and stability of oxygen evolution reactions in alkaline media. Sustain. Energy Fuels 2018, 2, 237–251. [Google Scholar] [CrossRef]
  114. Indra, A.; Menezes, P.W.; Zaharieva, I.; Baktash, E.; Pfrommer, J.; Schwarze, M.; Dau, H.; Driess, M. Active mixed-valent MnOx water oxidation catalysts through partial oxidation (corrosion) of nanostructured MnO particles. Angew. Chem.-Int. Ed. 2013, 52, 13206–13210. [Google Scholar] [CrossRef]
  115. Zeng, K.; Li, W.; Zhou, Y.; Sun, Z.; Lu, C.; Yan, J.; Choi, J.-H.; Yang, R. Multilayer hollow MnCo2O4 microsphere with oxygen vacancies as efficient electrocatalyst for oxygen evolution reaction. Chem. Eng. J. 2021, 421, 127831. [Google Scholar] [CrossRef]
  116. Fan, C.; Wu, X.; Li, M.; Wang, X.; Zhu, Y.; Fu, G.; Ma, T.; Tang, Y. Surface chemical reconstruction of hierarchical hollow inverse-spinel manganese cobalt oxide boosting oxygen evolution reaction. Chem. Eng. J. 2022, 431, 133829. [Google Scholar] [CrossRef]
  117. Zhang, L.; Shi, L.; Huang, L.; Zhang, J.; Gao, R.; Zhang, D. Rational Design of High-Performance DeNOx Catalysts Based on MnxCo3−xO4 Nanocages Derived from Metal−Organic Frameworks. ACS Catal. 2014, 4, 1753–1763. [Google Scholar] [CrossRef]
  118. Lyons, M.E.G.; Brandon, M.P. The significance of electrochemical impedance spectra recorded during active oxygen evolution for oxide covered Ni, Co and Fe electrodes in alkaline solution. J. Electroanal. Chem. 2009, 631, 62–70. [Google Scholar] [CrossRef]
  119. Silva, V.D.; Raimundo, R.A.; Simões, T.A.; Loureiro, F.J.; Fagg, D.P.; Morales, M.A.; Macedo, D.A.; Medeiros, E.S. ScienceDirect Nonwoven Ni e NiO/carbon fibers for electrochemical water oxidation. Int. J. Hydrogen Energy 2020, 46, 3798–3810. [Google Scholar] [CrossRef]
  120. Silva, T.; Silva, V.; Santos, J.; Simões, T.; Macedo, D. Effect of Cu-doping on the activity of calcium cobaltite for oxygen evolution reaction. Mater. Lett. 2021, 298, 130026. [Google Scholar] [CrossRef]
  121. Silva, V.D.; Simões, T.A.; Loureiro, F.J.; Fagg, D.P.; Figueiredo, F.M.; Medeiros, E.S.; Macedo, D.A. Solution blow spun nickel oxide/carbon nanocomposite hollow fibres as an efficient oxygen evolution reaction electrocatalyst. Int. J. Hydrogen Energy 2019, 44, 14877–14888. [Google Scholar] [CrossRef]
Figure 1. Procedure for the synthesis of samples MnXCo3-XO4 (0 ≤ X ≤ 1).
Figure 1. Procedure for the synthesis of samples MnXCo3-XO4 (0 ≤ X ≤ 1).
Nanomaterials 12 03170 g001
Figure 2. Physical, chemical, and electrochemical characterization of MnXCo3-XO4 (0 ≤ X ≤ 1).
Figure 2. Physical, chemical, and electrochemical characterization of MnXCo3-XO4 (0 ≤ X ≤ 1).
Nanomaterials 12 03170 g002
Figure 3. (a) XRD patterns of MnXCo3-XO4 (0 ≤ X ≤ 1) samples (b) Peaks (311) for each sample and their positions and FWHM. Blue lines below the diffractograms are the difference between the calculated and experimental data.
Figure 3. (a) XRD patterns of MnXCo3-XO4 (0 ≤ X ≤ 1) samples (b) Peaks (311) for each sample and their positions and FWHM. Blue lines below the diffractograms are the difference between the calculated and experimental data.
Nanomaterials 12 03170 g003
Figure 4. FESEM images and particle size distribution of MnXCo3-XO4 (0 ≤ X ≤ 1) samples (a): X = 0.0, (b) X = 0.2, (c) X = 0.4, (d) X = 0.6, (e) X = 0.8, and (f) X = 1.0.
Figure 4. FESEM images and particle size distribution of MnXCo3-XO4 (0 ≤ X ≤ 1) samples (a): X = 0.0, (b) X = 0.2, (c) X = 0.4, (d) X = 0.6, (e) X = 0.8, and (f) X = 1.0.
Nanomaterials 12 03170 g004
Figure 5. TEM images and selected area electron diffraction (SAED) of MnXCo3-XO4 (0 ≤ X ≤ 1) samples: (ac) X = 0.0, (df) X = 0.2, (gi) X = 0.4, (jl) X = 0.6, (mo) X = 0.8, and (pr) X = 1.0.
Figure 5. TEM images and selected area electron diffraction (SAED) of MnXCo3-XO4 (0 ≤ X ≤ 1) samples: (ac) X = 0.0, (df) X = 0.2, (gi) X = 0.4, (jl) X = 0.6, (mo) X = 0.8, and (pr) X = 1.0.
Nanomaterials 12 03170 g005
Figure 6. FT-IR spectra for all MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Figure 6. FT-IR spectra for all MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Nanomaterials 12 03170 g006
Figure 7. UV–Vis Absorbance spectra of MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Figure 7. UV–Vis Absorbance spectra of MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Nanomaterials 12 03170 g007
Figure 8. Raman spectra for all MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Figure 8. Raman spectra for all MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Nanomaterials 12 03170 g008
Figure 9. High-resolution XPS spectra of MnXCo3-XO4 (0 ≤ X ≤ 1) for (a) Co 2p, (b) Mn 2p and (c) O 1s.
Figure 9. High-resolution XPS spectra of MnXCo3-XO4 (0 ≤ X ≤ 1) for (a) Co 2p, (b) Mn 2p and (c) O 1s.
Nanomaterials 12 03170 g009
Figure 10. Magnetization curve of MnXCo3-XO4 (0 ≤ X ≤ 1).
Figure 10. Magnetization curve of MnXCo3-XO4 (0 ≤ X ≤ 1).
Nanomaterials 12 03170 g010
Figure 11. (a) LSV collected at 5 mV s−1 in 1 mol L−1 KOH for MnXCo3-XO4 (0 ≤ X ≤ 1) electrodes, where X = 0.0, X = 0.2, X = 0.4, X = 0.6, X = 0.8, and X = 1.0, and (b) the corresponding Tafel slopes; (c) anodic current (ia) versus scan rate to determine CDL; (d) chronopotentiometry analysis measured at 10 mA cm−2.
Figure 11. (a) LSV collected at 5 mV s−1 in 1 mol L−1 KOH for MnXCo3-XO4 (0 ≤ X ≤ 1) electrodes, where X = 0.0, X = 0.2, X = 0.4, X = 0.6, X = 0.8, and X = 1.0, and (b) the corresponding Tafel slopes; (c) anodic current (ia) versus scan rate to determine CDL; (d) chronopotentiometry analysis measured at 10 mA cm−2.
Nanomaterials 12 03170 g011
Figure 12. EIS—Nyquist (a) and Bode (b) plots for the tested electrocatalysts.
Figure 12. EIS—Nyquist (a) and Bode (b) plots for the tested electrocatalysts.
Nanomaterials 12 03170 g012
Table 1. Crystallographic parameters, including crystallite size, lattice parameter, and agreement indices (Rwp, Rexp e χ2) for samples of MnXCo3-XO4 (0 ≤ X ≤ 1).
Table 1. Crystallographic parameters, including crystallite size, lattice parameter, and agreement indices (Rwp, Rexp e χ2) for samples of MnXCo3-XO4 (0 ≤ X ≤ 1).
SamplesCo3O4—ICSD 36256MnCo2O4—ICSD 18544Agreement Factors
DXRD (nm)a (Å)DXRD (nm)a (Å)Rwp (%)Rexp (%)χ2
Co3O4—ICSD 36256---8.072(3)---------------
MnCo2O4—ICSD 18544---------8.28(2)---------
X = 090.6 [100%]8.0757(1)------7.166.921.04
X = 0.282.5 [100%]8.0759(8)------7.956.591.21
X = 0.480 [100%]8.0754(1)------7.976.601.21
X = 0.671 [100%]8.1183(7)------8.146.671.22
X = 0.868.8 [100%]8.1675(8)------8.016.661.20
X = 1------66 [100%]8.2381(6)8.696.701.30
Table 2. Assignment of the FT-IR band frequencies observed for MnXCo3-XO4 (0 ≤ X ≤ 1) samples and their comparison with the literature.
Table 2. Assignment of the FT-IR band frequencies observed for MnXCo3-XO4 (0 ≤ X ≤ 1) samples and their comparison with the literature.
Observed Frequencies (cm−1)
(X = 0.0)(X = 0.2)(X = 0.4)(X = 0.6)(X = 0.8)(X = 1.0)ReferenceMode Assignment
344034403440344034403440[68,69,70,71,72]O-H stretching vibration
163516351635163516351635[65,68,71,73,74]Angular deformation of adsorbed water molecules
138313831383138313831383[65,66,67]Deformations of C-N and CH2 groups
110011001100110011001100[62,68,74]C-O stretching vibrations
663662661653643643[65,68,70,71,75]Stretching vibrations of Mn–O
570568568561554552[21,63,64,72,73]Stretching vibrations of Co–O
Table 3. Raman active band positions for MnXCo3-XO4 (0 ≤ X ≤ 1) samples and their comparison with the literature.
Table 3. Raman active band positions for MnXCo3-XO4 (0 ≤ X ≤ 1) samples and their comparison with the literature.
Raman Band Position (cm−1)
MnXCo3-XO4 (0 ≤ X ≤ 1)F2gEgF2gF2gA1gReference
(X = 0.0)193480518617688[82,83,84]
(X = 0.2)186473512611682[20,88]
(X = 0.4)186471510605675[81,82]
(X = 0.6)185468508603667[20,83]
(X = 0.8)183480--661[82,89]
(X = 1.0)182488--660[79,80]
Table 4. Gaussian-fitted peaks for O 1s XPS spectra for MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
Table 4. Gaussian-fitted peaks for O 1s XPS spectra for MnXCo3-XO4 (0 ≤ X ≤ 1) samples.
SampleX = 0X = 0.2X = 0.4X = 0.6X = 0.8X = 1
Peak (eV)531.717531.386 531.448531.158 531.195 531.618
O2 area (nm)9859.48712887.3913100.6513363.59015330.31019922.580
O2/O10.4950.4980.5020.5170.6930.940
Table 5. Comparison of OER performance of nanostructured MnXCo3-XO4 (0 ≤ X ≤ 1) catalysts reported in the literature. Data refer to an overpotential to generate j = 10 mA cm210).
Table 5. Comparison of OER performance of nanostructured MnXCo3-XO4 (0 ≤ X ≤ 1) catalysts reported in the literature. Data refer to an overpotential to generate j = 10 mA cm210).
CatalystSubstrate *ղ10 (mV vs. RHE)Tafel Slope (mV dec−1)Reference
MnXCo3-XO4(0 ≤ X ≤ 1) powders (ágar-ágar)Ni foam29955This work
MnXCo3-XO4(0 ≤ X ≤ 2) powdersNi foam32779[16]
MnCo2O4GC510123[98]
Co3O4 nanoparticlesCFP36187.5[99]
Co3O4 NanosheetNi foam190103[100]
MnXCo3-XO4 (X = 0.3)Ni foam390N.R.[101]
MnXCo3-XO4 (X = 0.6)GC36550.6[91]
MnCo2O4Ni foam358N.R.[102]
MnXCo3-XO4 (1:3 ratio)Ni foam222162[103]
MnCo2O4Ni foam40090[104]
MnCo2O4carbon cloth400190[105]
MnCo2O4GC510123[98]
* CFP (carbon fiber paper); GC (glassy carbon).
Table 6. EIS—Results of fitting of the impedance spectra reported in Figure 12.
Table 6. EIS—Results of fitting of the impedance spectra reported in Figure 12.
SampleRS (Ω)RP (Ω)CDL (mF)R-ad (Ω)C-ad (mF)
X = 1.00.416.022.5512926.07
X = 0.00.6228.151.4912763.36
X = 0.20.4830.696.0119178.47
X = 0.40.5220.294.5513748.86
X = 0.60.4319.213.88214910.34
X = 0.80.469.708.27113511.77
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Santos, J.R.D.; Raimundo, R.A.; Silva, T.R.; Silva, V.D.; Macedo, D.A.; Loureiro, F.J.A.; Torres, M.A.M.; Tonelli, D.; Gomes, U.U. Nanoparticles of Mixed-Valence Oxides MnXCO3-XO4 (0 ≤ X ≤ 1) Obtained with Agar-Agar from Red Algae (Rhodophyta) for Oxygen Evolution Reaction. Nanomaterials 2022, 12, 3170. https://doi.org/10.3390/nano12183170

AMA Style

Santos JRD, Raimundo RA, Silva TR, Silva VD, Macedo DA, Loureiro FJA, Torres MAM, Tonelli D, Gomes UU. Nanoparticles of Mixed-Valence Oxides MnXCO3-XO4 (0 ≤ X ≤ 1) Obtained with Agar-Agar from Red Algae (Rhodophyta) for Oxygen Evolution Reaction. Nanomaterials. 2022; 12(18):3170. https://doi.org/10.3390/nano12183170

Chicago/Turabian Style

Santos, Jakeline Raiane D., Rafael A. Raimundo, Thayse R. Silva, Vinícius D. Silva, Daniel A. Macedo, Francisco J. A. Loureiro, Marco A. M. Torres, Domenica Tonelli, and Uílame U. Gomes. 2022. "Nanoparticles of Mixed-Valence Oxides MnXCO3-XO4 (0 ≤ X ≤ 1) Obtained with Agar-Agar from Red Algae (Rhodophyta) for Oxygen Evolution Reaction" Nanomaterials 12, no. 18: 3170. https://doi.org/10.3390/nano12183170

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