Next Article in Journal
Impact of Grinding Depth on Dislocation Structures and Surface Hardening in C45 Steel
Previous Article in Journal
Dynamical Characterization of Plates Containing Plane Cracks with Functional Gradient Materials
Previous Article in Special Issue
The Design of Diatomite/TiO2/MoS2/Nitrogen-Doped Carbon Nanofiber Composite Separators for Lithium–Sulfur Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molten Salt Synthesis and Electrochemical Evaluation of Na/Ag-Containing MnxOy Composites for Pseudocapacitor Applications

by
Carmen Martínez-Morales
1,
Antonio Romero-Serrano
1,
Josué López-Rodríguez
1,* and
Paulina Arellanes-Lozada
2,*
1
Departamento de Metalurgia y Materiales, Instituto Politécnico Nacional-ESIQIE, UPALM-Zacatenco, Av. Instituto Politécnico Nacional s/n, Lindavista, G. A. Madero, CDMX, Ciudad de Mexico 07738, Mexico
2
Dirección de Investigación, Instituto Mexicano del Petróleo. Eje Central Norte Lázaro Cárdenas No. 152, San Bartolo Atepehuacan, G. A. Madero, CDMX, Ciudad de Mexico 07730, Mexico
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(16), 3869; https://doi.org/10.3390/ma18163869
Submission received: 2 July 2025 / Revised: 2 August 2025 / Accepted: 7 August 2025 / Published: 18 August 2025

Abstract

Different composites of manganese oxides (MnxOy) containing sodium (Na) and silver (Ag) were synthesized by the molten salt method with various MnSO4·H2O/NaNO3 (M/N) molar ratios (between 0.3 and 1), and different AgNO3 and NaOH amounts, obtaining two groups of materials: without the addition of AgNO3 (labeled as M/N) and with AgNO3 (labeled as M/N-A). As for the M/N group, the system with the lowest M/N ratio yielded the highest specific capacitance (160.5 F g 1 ), attributed to the formation of Mn3O4 and sodium birnessite. In the M/N-A group, the 1 M/N-0.5A system, produced with M/N ratio of 1 and addition of 0.5 g of AgNO3, exhibited the highest specific capacitance (229.1 F g 1 ), associated with the presence of Mn2O3, silver hollandite, and metallic Ag. This enhancement is attributed to the synergistic effects of Na+ and Ag+ ions, which improve charge transfer kinetics and electrochemical performance. It was demonstrated that decreasing the MnSO4·H2O/NaNO3 ratio in the M/N group and increasing AgNO3 content in the M/N-A group enhances the electrochemically active surface area. Galvanostatic charge–discharge (GCD) and electrochemical impedance spectroscopy (EIS) techniques confirmed that the 1 M/N-0.5A system exhibited the best performance, characterized by high energy retention, stable cycling behavior, and low capacitance dispersion, indicating its strong potential as an active material for pseudocapacitor applications.

Graphical Abstract

1. Introduction

Currently, the world is facing a significant energy crisis triggered by the accelerated demographic growth and over-exploitation of non-renewable energy sources such as fossil fuels [1,2]. Before this challenge, the scientific community had focused its efforts on developing and optimizing advanced energy storing technologies that include electrochemical batteries, fuel cells, and supercapacitors in order to improve the efficiency, sustainability, and feasibility of the energy systems [3,4]. Supercapacitors (SCs) are storing systems of electrochemical energy that follow a Faradaic process based on redox reactions on the electrode material, and due to their nature, these materials have high power, energy density, and good cyclability capacity [5,6]. There are two types of SCs: (a) the electric double-layer capacitors, which accumulate and separate charge at the electrode–electrolyte interface without involving electrochemical reactions and (b) pseudocapacitors, which store charge through redox reactions or insert ions between the electrode and electrolyte by means of their high power density, fast charge–discharge response, and good electrochemical stability [7,8]. The different types of pseudocapacitors can be identified by their chemical nature, which can be represented by transition metal oxides ( MO x ; M = Fe, Co, Ni, Mn, etc.), metal hydroxides, and conducting polymers [9,10]. Numerous transition metal oxides have been employed as pseudocapacitors, where RuO x , MnO x , NiO, CoO x , MoO x , and VO x have stood out [11,12]. MnO x materials such as MnO, MnO 2 , Mn 2 O 3 , Mn 3 O 4 , etc., have been widely studied because of their high theoretical capacitance (755–1232 mAh g 1 ) [13]. MnO 2 and Mn 2 O 3 stand out due to characteristics, such as high theoretical specific capacitance, excellent reversibility, high specific surface area, relatively low conversion potential, small voltage hysteresis, multi-oxidation states, low cost, high availability, and ecofriendly nature [2,5,7,9,13,14]. However, these Mn x O y materials have a low electrical conductivity compared to others electrochemical capacitors [5,13,14]. One strategy to overcome their kinetic and conductivity limitations is to insert alkyl cations such as Na, K, Ag, and transition metals into the Mn x O y crystal lattice, thereby improving the material’s electrochemical behavior [12]. Thanigai et al. performed a comparative analysis of the electrochemical behavior of Mn 3 O 4 , Na-MnO 2 , and Na-MnO 2 x ; the latter was obtained by introducing oxygen vacancies in Na + ions inserted in MnO 2 . They concluded that the improvement of electrochemical properties of materials is related to electrochemical kinetics and the increase in the active surface area of Na-MnO 2 x [9]. Vashisth et al. conducted a study using the solid-state reaction method to synthesize α -Na 0.5 Mn0.9302@Na0.91MnO2 (NaMnO) and K 0.48 Mn1.9405@Na0.91MnO2 (KNaMnO) nanocomposites; they indicated that no Na-K phases were formed; rather, these cations interacted independently with the MnO 2 crystal lattice, which generated high specific capacitance values for NaMnO and KNaMnO of approximately 361 and 143 F g 1 , respectively [15]. In contrast, the synthesis of Ag 2 -MnO 2 ultrathin nanosheets has also been studied by Rahman et al., in 2021, who indicated that one-pot hydrothermal synthesis process at 150 °C can favor the electrode–electrolyte diffusive process by the good electrical conductivity of Ag ions, and the increased active sites at the boundaries of the MnO 2 crystal lattice [12].
It is well known that the properties and performance of materials are greatly affected by the synthesis method and conditions used to produce them. Among the different synthesis methods for obtaining electrode materials such as hydrothermal [16,17,18], sol-gel [19], co-precipitation [20], microemulsion [21], and combustion, the molten salt synthesis has proven to be a cost-effective, simple to operate, easy to scale, generalizable, and environmentally friendly process [16,22,23]. Additionally, molten salt synthesis has certain advantages over other methods, such as solid-state reactions. With a liquid phase of salts, faster mass transport by convection and diffusion is achieved [23,24,25]. This is evident in the study conducted by Zhao et al., who synthesized Mn 2 O 3 by the molten salt method employing KNO 3 -NaNO 2 -NaNO 3 and manganese acetate as reaction medium and manganese source, respectively; these authors claimed that such a synthesis method could achieve electrode materials with outstanding electrochemical performance related to a high specific surface area [22].
In this work, ten electrode materials were synthesized by the molten salt method, modifying the MnSO4· H 2 O/NaNO 3 ratio and AgNO 3 concentration, in order to obtain Mn x O y materials with Na + and Ag + residual ions that improve the pseudocapacitive properties of Mn x O y . X-ray diffraction (XRD) and scanning electron microscopy–energy dispersive X-ray spectroscopy (SEM-EDS) were used to analyze the morphology and crystalline structure of the electrode materials, whereas the performance of the electrodes was evaluated using the cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS) techniques with 0.1 M Na 2 SO 4 as electrolyte.

2. Materials and Methods

2.1. Synthesis

Ten electrode materials were synthesized by the molten salt method, whose abbreviations and compositions are presented in Table 1. All the reagents used in the experiments were of analytical grade and used without further purification. Different mass ratios of MnSO4·H2O, NaNO 3 , AgNO 3 , and NaOH were placed in an agate mortar, and ground and homogenized for 30 min to promote interaction between the reactants during precipitation. Subsequently, a nickel crucible containing each salt mixture (MnSO4·H2O/NaNO3AgNO3/NaOH) was put in a tube furnace supplied with Ar at 420 °C for 30 min, at a heating rate of 10 °C/min. Once the mixed salt reached the melting point, the samples were cooled down inside the furnace until they reached room temperature. Then, the precipitated powder was collected, washed, and centrifuged with deionised water eight times [26,27]. This procedure was carried out to remove sub-products of the molten salt synthesis (i.e., remaining water-soluble salts) and separate the Mn x O y composites. After drying them in an oven at 90 °C for 6 h, a black solid was finally obtained.

2.2. Characterization of Materials

XRD analysis of the different materials was carried out using a Bruker D8 Advance diffractometer within a 2 θ interval ranging from 0 to 90° and Cu k α radiation with a wavelength of approximately λ = 1.54056 Å. The XRD results were processed using the Match software version 4.1 Build 311. The surface morphology and elemental composition of the materials were analyzed by SEM using a high-resolution microscope JEOL JSM 6701F coupled with EDS. ImageJ software Version 1.54f (Wayne Rasband, MD, USA) was used to analyze the particle size distribution on the surface of the Mn x O y composites. The electrochemical behavior of the synthesized materials was analyzed by CV, GCD, and EIS, employing a Biologic SP-300 potentiostat/galvanostat (BioLogic Science Instruments, Seyssinet-Pariset, France) and the EC-Lab software version V11.34. Electrochemical measurements were performed in a three-electrode glass cell with Ag/AgCl/NaCl [3 M] as reference electrode, graphite as counter electrode, and Mn x O y /glassy carbon as working electrode. The latter was prepared by using Mn x O y -based ink on a glassy carbon substrate. The ink was made by mixing 0.03 g of synthesized composites, 0.0015 g of carbon black vulcan XC 72, and 300 µL of acetone with 60 µL of a 5 wt.% Nafion 117 solution. The mixture was placed in an ultrasonic bath for 30 min to form a homogeneous black suspension, which was then carefully pipetted onto a polished glassy carbon electrode of 3 mm in diameter using 1000 and 2000 grit SiC paper and diamond paste until achieving a mirror finish to ensure proper adhesion of the materials. The mass loading of Na/Ag-containing Mn x O y electrodes was 0.279 mg. Finally, the ink was dried at room temperature for 1 h [28,29]. The employed electrolyte was a 0.1 M Na 2 SO 4 solution. CV measurements were conducted within a potential interval ranging from 0.0 to 0.9 V vs. Ag/AgCl/NaCl [3 M] at scan rates of 5, 10, 20, 50, and 100 mV s 1 . GCD profiles were obtained at current densities of 0.5, 1.0, 1.5, and 2.0 A g 1 within a potential range of 0.0 to 1.0 V. Cycling performance was studied at a scan rate of 1 A g 1 for over 500 cycles. Regarding the EIS technique, experimental tests were run at open-circuit potential within a frequency interval ranging from 100 kHz to 10 mHz with sine wave amplitude of 10 mV.

3. Results

3.1. Materials Characterization Using XRD and SEM-EDS

Figure 1 shows the XRD spectra of Na/Ag-containing Mn x O y composites synthesized by the molten salt method with different MnSO4·H2O/NaNO3 ratios (M/N). The 0.5 M/N and 0.7 M/N spectra show characteristic peaks at 2 θ of 23.1, 32.8, 38.1, 55.0, and 65. 6 , indicating the formation of Mn 2 O 3 (Bixbyite) (COD 96-151-4114) [30]; this compound has a cubic phase and lattice parameter of a = 9.43000 Å [30,31]. This crystal phase, obtained by hydrothermal synthesis, has been reported to exhibit good electrocatalytic behavior [31]. Regarding the 0.5 M/N system, an additional peak at 36. 7 was associated with the formation of Na(OH)( H 2 O) (COD 96-231-0701), which could reduce the electrochemical behavior of the electrode material. In contrast, a new phase was formed when the M/N ratio decreased; this event occurred in the 0.3 M/N and 0.4 M/N systems, where Mn 3 O 4 (Hausmannita) was present as confirmed by the peaks at 28.8, 31.0, 32.2, 38.0, 44.4, 50.8, 58.5, 59.8, and 64. 6 (COD 96-151-4121), see Figure 1 [32]. Firstly, the molten salt synthesis process allowed the ionization of MnSO 4 (Reaction (1)). Additionally, the synthesis of Mn 3 O 4 was achieved thanks to the active oxygen released from the nitrate decomposition reaction (Reactions (2) and (3)) and the presence of hydroxyl ions from NaOH ionization, as observed in Reactions (4)–(6) [9,33]:
MnSO 4 Mn 2 + + SO 4 2 +
2 NaNO 3 2 NaNO 2 + O 2
2 AgNO 3 2 Ag + 2 NO 2 + O 2
2 Mn 2 + + 5 OH Mn ( OH ) 2 + 2 Mn ( OH ) 3
Mn ( OH ) 2 + 2 Mn ( OH ) 3 Mn 3 O 4 + 4 H 2 O
6 Mn 2 + + 12 OH + O 2 2 Mn 3 O 4 + 6 H 2 O
Mn 3 O 4 has a tetragonal arrangement and belongs to the spinel class, where Mn 2 + and Mn 3 + cations occupy tetrahedral and octahedral sites, respectively [32,34]. Therefore, manganese is surrounded by oxygen atoms at tetrahedral and octahedral positions [33]. Additionally, the 0.3 M/N system shows a characteristic peak at 24. 9 , related to Na 0.31 1 + ( Mn 0.69 4 + Mn 0.31 3 + )O2 ·  0.4 H 2 O (sodium birnessite) formation (COD 96-153-1680). This compound has also been obtained by hydrothermal synthesis and has a tricyclic structure when sodium cations and water molecules are distributed by vacancy-free Mn-bearing octahedral layers [35]. The presence of sodium birnessite could result in better electrochemical behavior due to the increased conductivity and/or active surface area with respect to the other systems [9,35].
Figure 2 shows XRD spectra of MnSO4· H 2 O/NaNO 3 systems synthesized with different concentrations of AgNO 3 (M/N-A). All the samples presented peaks of the Mn 2 O 3 phase at 2 θ of 23.1, 32.8, 38.1, 45.1, 49.2, 55.0, and 65.6 ° (COD 96-151-4114) [30]. The low concentration of AgNO 3 in the 0.5 M/N-0.1A, 0.5 M/N-0.3A, 1 M/N-0.1A, and 1 M/N-0.3A systems did not promote the doping of Ag into the material lattice. Since the amount of NaOH was reduced to 1 g in electrode materials synthesized with AgNO 3 , the Mn 2 O 3 synthesis was favored by the reaction with oxygen released from the precursors (Reactions (2) and (3)) and Mn ( OH ) 2 (Reaction (7)) [36,37].
4 Mn ( OH ) 2 + O 2 2 Mn 2 O 3 + 4 H 2 O
Additionally, XRD spectra of the 0.5 M/N-0.5A and 1 M/N-0.5A systems exhibited characteristic peaks of Ag 1.15 Mn 8 O 16 (silver hollandite) at 2 θ of 28.7, 36.7, and 37. 4 (COD 96-450-6592) [38]. Hollandite-type materials consist of MnO 6 octahedra with a tunnel structure of approximately ∼4.7 Å, which facilitates the insertion of small ions such as Na + , K + , Rb + , Ag + , and Ba 2 + [39,40]. Particularly, silver hollandite is an electrochemically active material because Ag + ions are located at the center of the unit cell, as reported by Takeuchi et al. [38]. Its synthesis by solid-state and hydrothermal methods is difficult to achieve, as reported by different research works [41,42,43,44]. However, in this study, better ionic diffusion of molten salts facilitated its production. The insertion of Ag atoms into the crystal lattice could enhance the electron transport efficiency and specific capacitance of the 0.5 M/N-0.5A and 1 M/N-0.5A systems due to the excellent electrical conductivity and surface activity of Ag atoms [12]. Finally, the XRD spectrum of the 1 M/N-0.5A system shows additional peaks at 2 θ of 38.0, 44.2, 64.4, and 77. 4 , which are associated with Ag (COD 96-900-8460) [38]. Since the relative intensity of the Ag peaks is higher than that of the Ag 1.15 Mn 8 O 16 peaks, it can be suggested that the doping of Mn 2 O 3 with Ag atoms was favored due to the high and low concentration of AgNO 3 and NaNO 3 precursors, respectively.
The surface morphology of the synthesized materials was analyzed by SEM-EDS (Figure 3 and Table 2). Figure 3a shows that the 0.3 M/N sample consists of disordered stacked bars of Mn 3 O 4 and sodium birnessite [45]. These bars reach a diameter of up to 293 nm and are covered with smaller particles, as shown in the corresponding histogram. The elemental mapping and EDS results indicate a uniform distribution of Mn, O, and Na, suggesting that the Na 0.31 1 + ( Mn 0.69 4 + Mn 0.31 3 + )O2 · 0.4 H 2 O phase is dispersed uniformly. The 0.5 M/N-0.5A and 1 M/N-0.5A samples, synthesized with Ag/NO 3 , are presented in Figure 3b and c, respectively. Their surface morphology has particles of Mn 2 O 3 , silver hollandite, and Ag of approximately 42 nm in size (see corresponding histogram) [46,47]. The elemental mapping reveals the uniform distribution of Ag, confirming the presence of Ag 1.15 Mn 8 O 16 and Ag. These results suggest that the molten salt synthesizing process enabled the nitrite and hydroxide salts to completely interact with the Mn source, thereby preventing product agglomeration.

3.2. Cyclic Voltammetry

The ion dynamics at the electrode–electrolyte interface of the Na/Ag-containing Mn x O y composites were studied by the CV technique. Figure 4 shows the CV curves of the electrode materials in 0.1 M Na 2 SO 4 at varying scan rates ranging from 0 to 0.9 V vs. Ag/AgCl/NaCl [3 M]. The presence of peaks at Ep1 ≈ 0.56 V and Ep2 ≈ 0.26 V is evident, which is attributed to redox reaction potentials of Mn ions (see Figure 4a). These characteristic peaks are less noticeable in the 0.3 M/N and 1 M/N-0.5A systems, suggesting a fast insertion/de-insertion of sodium ions [15]. Additionally, the quasi-rectangular shape of the CV profiles indicates that the analyzed systems can store energy as pseudocapacitors; this characteristic is more evident in 0.5 M/N-0.5A and 1 M/N-0.5A than in 0.3 M/N. These results reveal that doping with a high concentration of AgNO 3 improves the pseudocapacitive properties of Mn x O y systems. Figure 4b,c show CV profiles of 0.3 M/N and 1 M/N-0.5A in 0.1 M Na 2 SO 4 at different scan rates, respectively. Increasing the scan rate resulted in area increase in the CV curve and modification of the quasi-rectangular shape of the CV profiles. Specific capacitance values ( C s , CV ) were determined from CV curves as follows:
C s , CV = V 2 V 1 i · d V v · Δ V · m
where i is the current, Δ V is the applied potential window ( Δ V = V 2 V 1 ), v is the voltage scan rate, and m is the mass of the electrode active material. The integral of Equation (8) was determined by using the internal area of each CV curve.
Figure 4d,e show the C s , CV values as functions of the scan rate of Na and Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 , respectively. As expected, increasing the scan rate reduced the C s , CV values due to more efficient ion diffusion between the electrolyte and each site of the active material at low scan rate [15]. Figure 4d presents C s , CV values of Na-containing Mn x O y composites, M/N systems, where the system with the lowest M/N ratio obtained the highest value of 160.5 F g 1 . In contrast, among M/N-A systems, the 0.5 M/N-0.5A and 1 M/N-0.5A systems at 5 mV s 1 achieved higher C s , CV values of 223.7 and 229.1 F g 1 , respectively (see Figure 4e). Sodium and silver ions contributed to the charge storage process by providing higher current density in these composites. These results are in good agreement with those reported by Gu et al., who indicated that Mn 2 O 3 -based electrodes had higher specific capacitance than Mn 3 O 4 -based electrodes [13].
In order to relate the electrochemically active surface area (ECSA) to the electrochemical behavior of the electrodes, the double-layer capacitance ( C d l ) was first determined using CV experiments. C d l and ECSA values were estimated using Equations (9) and (10), respectively [48].
C d l = i a i c 2 v
E C S A = C d l C s
where i a and i c are anodic and cathodic current at Δ E = 0.44 V (non-faradaic potential window) and C s is in F cm 2 . Figure 4f plots the results obtained from the different systems, revealing a clear trend. Decreasing the MnSO4· H 2 O/NaNO 3 ratio (M/N group) and increasing the AgNO 3 concentration (M/N-A group) increased the ECSA values. The 0.3 M/N, 0.5 M/N-0.5A, and 1 M/N-0.5A systems presented the highest values; therefore, increasing the number of active sites in these systems improves electrochemical kinetics and enables high charge storage.
The main energy storage mechanisms of pseudocapacitor materials involve electrochemical reactions in the electrode bulk and on the electrode surface [5]. In the case of Mn x O y composites, it has been reported that an electron–proton transfer systems involve the insertion/de-insertion of alkali metal cations through a fast and reversible charge storage process, although reduction and oxidation reactions have been reported for Mn x O y nanocomposites [9,15]. For the 0.5 M/N-0.5A and 1 M/N-0.5A systems, the presence of Ag 1.15 Mn 8 O 16 allows this electrochemical process to be represented as follows:
Ag 1.15 Mn 8 O 16 + Na + + e Ag 1.15 Mn 8 O 16 Na
Additionally, sodium cations from the electrolyte can be adsorbed on the Ag 1.15 Mn 8 O 16 surface, which contributes to the charge–discharge process:
( Ag 1.15 Mn 8 O 16 ) surface + Na + + e ( Ag 1.15 Mn 8 O 16 Na ) surface

3.3. Galvanostatic Charge–Discharge

Figure 5a illustrates the galvanostatic charge–discharge (GCD) profiles of the 0.3 M/N, 0.5 M/N-0.5A, and 1 M/N-0.5A systems within a potential range of 0.0 to 1.0 V and a current density of 0.5 A g 1 . The 1 M/N-0.5A stands out due to its relatively linear and semi-geometric behavior, which indicates high reversibility and high columbic efficiency. Additionally, a small IR drop is observed in the GCD profiles. This is associated with two resistances: R E S R , which is related to the resistance of the electrolyte, cables, and electrode; and R E D R , which is related to the resistance of ions accessing the outer pore of the electrode [49]. Therefore, at a current density of 0.5 A g 1 , the energy stored in the double electric layer during charging of the 0.5 M/N-0.5 A and 1 M/N-0.5 A systems is not lost during discharge. This reflects the good capacitive behavior and electrical conductivity of the Ag-containing Mn x O y composites. Furthermore, the 1 M/N-0.5A system exhibited a longer discharge time than the 0.3 M/N and 0.5 M/N-0.5A systems, suggesting that the presence of Ag and silver hollandite atoms enables higher specific capacitance.
Figure 5b shows the GCD profiles of the 1 M/N-0.5A system obtained at different current densities (from 0.5 to 2 A g 1 ). As expected, the lower the applied current density, the longer the total discharge time ( t d ). Using the t d values obtained and reported in Table 3, the specific capacitance by GCD ( C s , GCD ) was calculated according to the following equation [2]:
C s , GCD = i · t d ( Δ V I R ) · m
where IR is the potential drop during discharge. The C s , GCD values are consistent with those reported by the CV technique (see Table 3). The specific energy and power of the 1 M/N-0.5A system were also determined from the GCD tests using the relationships outlined by the Equations (14) and (15), respectively [2,50]:
E = C s , GCD · Δ V 2 2
P = E t d
The specific energy and power of the 1 M/N-0.5A system at a current of 0.5 A g 1 were 31.3 Wh kg 1 and 253.8 W kg 1 , respectively (see Table 1). These values are related to the good specific capacitance and rapid charge–discharge capacity in 0.1 M NaSO 4 of the composite [50]. Finally, the Figure 5 shows the results obtained for the cyclic stability of the 1 M/N-0.5A system at 1 A g 1 . The system exhibits adequate cyclic stability during the initial cycles; however, an increase in the number of cycles results in a slight decrease in coulombic efficiency, reaching 88.6 % after 500 cycles.

3.4. Electrochemical Impedance Spectroscopy

Figure 6 shows Nyquist plots of Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 . High- (≈100 kHz) and low-frequency (≈0 mHz) regions are observed, which are related to low and high real impedance values, respectively. Similar impedance behavior was obtained for all the systems: at high frequencies, the impedance spectra present incomplete semicircles; at medium frequencies, linear behavior begins at an angle of approximately 45° with respect to the real impedance axis (x-axis); finally, at low frequencies, a linear tendency is maintained, and angle values are greater than 45°. In the high frequency region, the incomplete semicircles represent the retrieved internal resistance, which is the sum of the electrode and bulk electrolyte resistances [51,52], where the impedance value at the intersection of the semicircle and real impedance is related to the system’s ability to transport charge [51]. For M/N systems, the lowest MnSO4· H 2 O/NaNO 3 ratio (0.3 M/N) shows the lowest value, see Figure 6a; in contrast, the 1 M/N-0.5A and 0.5 M/N-0.5A systems (Figure 6b) achieved lower values at high silver concentrations, indicating a fast charge transport process, a desired characteristic in pseudocapacitors. In addition, at intermediate frequencies, the transition of the curves to linear behavior with 45° slope angles indicates that all reactive sites at the solid–liquid interface interact rapidly with the ions in the electrolyte [53]. The low-frequency region of the Nyquist plots shows lines with slope angles that are mostly greater than 45°. Ideal capacitors are known to exhibit linear behavior with an angle of 90°. Therefore, it can be concluded that the 1 M/N-0.5A system has superior pseudocapacitive properties compared to the rest of the systems [54].
In order to interpret quantitatively the EIS results and analyze the properties of these supercapacitive materials, the experimental data from the impedance spectra were fitted to an EEC, which combined capacitive and resistive elements (Figure 6c). The EEC is composed of the following electrical elements: an electrode resistance ( R e ); an R C c t component corresponding to the charge transfer process at the current collector/active material interface, related to a resistor (R), and a constant phase element (C) that describes a pseudocapacitive element (C); an R C d l component related to the diffuse layer of the interface; and finally, equilibrium differential capacitance ( C e q ) [51,52,55]. The electrical elements obtained from the different systems by the EEC are reported in Table 4. The low R e l values suggest that the methodology used to prepare the working electrode was effective. The C d l values were calculated using the C P E d l and R d l components [56,57]:
C d l = ( Y d l R d l 1 n ) 1 n
where Y d l and n are the proportionality factor and the exponent of the C P E , respectively. The 0.3 M/N, 0.5 M/N-A, 1 M/N-0.3A, and 1 M/N-0.5A systems presented the highest C d l values (see Table 4). As C d l contributes to the charge storage capacity of the electrode, it can be concluded that the low MnSO4· H 2 O/NaNO 3 ratio and high AgNO 3 concentration improved the pseudocapacitive properties of these systems. The values of the proportionality factor of the equilibrium differential capacitance ( Y e q ) related to the impedance response at low frequencies suggest lower capacitance dispersion in the analyzed systems. The R c t values obtained by the EEC are shown in Figure 6d. An increase in MnSO4· H 2 O increased the R c t values for the M/N composite group. However, for the M/N-A systems, a larger amount of AgNO 3 provoked a reduction in R c t . For the last composite group, the charge transport process from the liquid phase (electrolyte) to solid phase (electrode material) was accelerated due to the increase in conductivity caused by Ag doping [4,52]. The lowest R c t value of 60.5 Ω was obtained with the 1 M/N-0.5A system, which elucidated its high specific power. Additionally, Figure 6d displays the inclination angles relative to the real impedance axis at low frequencies. The results suggested that the 1 M/N-0.5A system presented an angle greater than 81. 5 , confirming its lower capacitance dispersion [49]. Using a smaller amount of NaNO 3 precursor during the molten salt synthesis process could favor the Mn 2 O 3 production with larger surface area, which facilitates the electron transfer [58].

4. Conclusions

The molten salt synthesis of Na/Ag-containing Mn x O y composites produced two groups of electrode materials: the first composite group synthesized without AgNO 3 (M/N) mainly presented Mn 2 O 3 , Mn 3 O 4 , and Na 0.31 1 + ( Mn 0.69 4 + Mn 0.31 3 + )O2· 0.4 H 2 O, while the second group, which was synthesized with AgNO 3 (M/N-A), showed Mn 2 O 3 , Ag 1.15 Mn 8 O 16 , and Ag. SEM-EDS analysis confirmed uniform surface structures typical of Na/Ag-doped Mn x O y composites. The 0.5 M/N-0.5A and 1 M/N-0.5A systems exhibited the highest specific capacitances, 223.7 and 229.1 F g 1 , respectively. The incorporation of Ag improved electrochemical performance by enhancing charge transfer kinetics, conductivity, and electroactive surface area. In GCD profiles, the 1 M/N-0.5A system showed a near-linear and semi-symmetric shape, indicating high reversibility and coulombic efficiency. At 0.5 A g 1 , it delivered a specific energy of 31.3 Wh kg 1 and power of 253.8 W kg 1 , attributed to its good capacitance and fast charge–discharge capability. Its cyclic performance at 1 A g 1 remained stable in early cycles, with coulombic efficiency reaching 88.7% after 500 cycles. The presence of Ag 1.15 Mn 8 O 16 and Ag in the 1 M/N-0.5A system caused lower R c t values, as well as an inclination angle greater than 81.5° in the EIS spectra. This good electrochemical behavior was related to the low and high amounts of NaNO 3 and AgNO 3 , respectively.

Author Contributions

Conceptualization, J.L.-R.; methodology, C.M.-M., A.R.-S. and J.L.-R.; software, C.M.-M.; validation, J.L.-R.; formal analysis, A.R.-S., P.A.-L. and J.L.-R.; investigation, A.R.-S. and P.A.-L.; resources, A.R.-S.; data curation, C.M.-M.; writing—original draft preparation, C.M.-M. and P.A.-L.; writing—review and editing, P.A.-L.; visualization, J.L.-R.; supervision, P.A.-L.; project administration, P.A.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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.

Acknowledgments

The authors are grateful to IPN for the financial support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Philip, A.; Kumar, A.R. Electrochemical kinetics of a novel electrode material comprising different phases of MnO2 (Mn2O3, γ-, λ-, and δ-), graphite, and PVDF. Mater. Chem. Phys. 2024, 325, 129793. [Google Scholar] [CrossRef]
  2. Chowdhury, A.; Shukla, R.; Bhattacharyya, K.; Tyagi, A.K.; Chandra, A.; Grover, V. Electrochemical performance of K+-intercalated MnO2 nano-cauliflowers and their Na-ion-based pseudocapacitors. Mater. Sci. Eng. B 2023, 295, 116581. [Google Scholar] [CrossRef]
  3. Mulla, N.R.; Patel, N.N.; Bhosale, S.B.; Patil, U.M.; Patil, R.S. Morphologically tuned MnO2 thin film electrodes prepared by growth kinetic dependent SILAR approach for high-performance extrinsic pseudocapacitors. J. Alloys Compd. 2024, 1006, 176261. [Google Scholar] [CrossRef]
  4. Aadil, M.; Taki, A.G.; Zulfiqar, S.; Rahman, A.; Shahid, M.; Warsi, M.F.; Ahmad, Z.; Alothman, A.A.; Mohammad, S. Gadolinium doped zinc ferrite nanoarchitecture reinforced with a carbonaceous matrix: A novel hybrid material for next-generation flexible capacitors. RSC Adv. 2023, 13, 28063–28075. [Google Scholar] [CrossRef]
  5. Wei, W.; Cui, X.; Chen, W.; Ivey, D.G. Manganese oxide-based materials as electrochemical supercapacitor electrodes. Chem. Soc. Rev. 2011, 40, 1697–1721. [Google Scholar] [CrossRef]
  6. Ahn, J.; Chang, W.; Song, Y.; Son, Y.; Ko, Y.; Cho, J. Binder-free, multidentate bonding-induced carbon nano-oligomer assembly for boosting charge transfer and capacitance of energy nanoparticle-based textile pseudocapacitors. Energy Storage Mater. 2024, 69, 103396. [Google Scholar] [CrossRef]
  7. Rusi; Majid, S.R. Controllable synthesis of flowerlike α-lMnO2 as electrode for pseudocapacitor application. Solid State Ionics 2014, 262, 220–225. [Google Scholar] [CrossRef]
  8. Rudra, M.; Saha, S.; Sinha, T.P. Phase transitions in oxygen-intercalated pseudocapacitor Pr2MgMnO6 electrode: A combined structural and conductivity analysis. Mater. Sci. Eng. B 2024, 307, 117517. [Google Scholar] [CrossRef]
  9. Thanigai Vetrikarasan, B.; Nair, A.R.; Shinde, S.K.; Kim, D.Y.; Kim, J.M.; Bulakhe, R.N.; Sawant, S.N.; Jagadale, A.D. Oxygen vacancy enriched Na-intercalated MnO2 for high-performance MXene (Ti3C2TX)-based flexible supercapacitor and electrocatalysis. J. Energy Storage 2024, 94, 112457. [Google Scholar] [CrossRef]
  10. Siddiqui, R.; Rani, M.; Ibrahim, A.; Shah, A.A.; Razaq, A.; Bano, S.; Ajmal Khan, M. Enhanced specific capacitance of supercapacitors using wide band gap NdCrO3 and NdCrO3/graphene oxide nanocomposites. J. Rare Earths 2024. [Google Scholar] [CrossRef]
  11. Roberts, A.J.; Slade, R.C.T. Effect of specific surface area on capacitance in asymmetric carbon/α-MnO2 supercapacitors. Electrochim. Acta 2010, 55, 7460–7469. [Google Scholar] [CrossRef]
  12. Rahman, A.U.; Zarshad, N.; Wu, J.; Faiz, F.; Raziq, F.; Ali, A.; Li, G.; Ni, H. Fabrication of Ag-doped MnO2 nanosheets@carbon cloth for energy storage device. Mater. Sci. Eng. B 2021, 269, 115150. [Google Scholar] [CrossRef]
  13. Xin, G.; Jie, Y.; Liangjun, L.; Haitao, X.; Jian, Y.; Xuebo, Z. General Synthesis of MnOx (MnO2, Mn2O3, Mn3O4, MnO) Hierarchical Microspheres as Lithium-ion Battery Anodes. Electrochim. Acta 2015, 184, 250–256. [Google Scholar] [CrossRef]
  14. Karuppaiah, M.; Sakthivel, P.; Asaithambi, S.; Murugan, R.; Babu, G.A.; Yuvakkumar, R.; Ravi, G. Solvent dependent morphological modification of micro-nano assembled Mn2O3/NiO composites for high performance supercapacitor applications. Ceram. Int. 2019, 45, 4298–4307. [Google Scholar] [CrossRef]
  15. Vashisth, P.; Sharma, A.; Nasit, M.; Singh, J.P.; Anjali; Varshney, M.; Kumar, S.; Won, S.O.; Shin, H.J. Fabrication of Na and K based MnO(2) nanocomposites for supercapacitive applications. Heliyon 2024, 10, e35360. [Google Scholar] [CrossRef]
  16. Wang, D.; Lu, J.; Gou, J.; Wang, Z.; Wang, M.; Gong, X.; Hao, S. A rapid method for the synthesis of perovskite (ATiO3, A=Ca, Sr, Ba) in molten chloride. Ceram. Int. 2019, 45, 19547–19549. [Google Scholar] [CrossRef]
  17. Sun, S.; Huang, M.; Wang, P.; Lu, M. Controllable Hydrothermal Synthesis of Ni/Co MOF as Hybrid Advanced Electrode Materials for Supercapacitor. J. Electrochem. Soc. 2019, 166, A1799. [Google Scholar] [CrossRef]
  18. Riyas, Z.; Manimuthu, R.; Sankaranarayanan, K. Hydrothermal synthesis of La2O3–ZnO nanocomposites as electrode material for asymmetric supercapacitor applications. J. Mater. Sci. Mater. Electron. 2023, 34, 1612. [Google Scholar] [CrossRef]
  19. Myasoedova, T.N.; Kalusulingam, R.; Mikhailova, T.S. Sol-Gel Materials for Electrochemical Applications: Recent Advances. Coatings 2022, 12, 1625. [Google Scholar] [CrossRef]
  20. Lv, G.; Tong, B.; Shi, W.; Li, N.; Wang, L. Characterization of LiNi0.8Co0.15Al0.05O2 cathode material synthesized via co-precipitation method. J. Phys. Conf. Ser. 2024, 2789, 012006. [Google Scholar] [CrossRef]
  21. Ganguli, A.K.; Ahmad, T.; Vaidya, S.; Ahmed, J. Microemulsion route to the synthesis of nanoparticles. Pure Appl. Chem. 2008, 80, 2451–2477. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Ji, D.B.; Wang, P.; Yan, Y.D.; Xue, Y.; Xu, H.B.; Liang, Y.; Luo, H.J.; Zhang, M.L.; Han, W. Molten salt synthesis of Mn2O3 nanoparticle as a battery type positive electrode material for hybrid capacitor in KNO3-NaNO2-NaNO3 melts. Chem. Eng. J. 2018, 349, 613–621. [Google Scholar] [CrossRef]
  23. Gupta, S.K.; Mao, Y. Recent Developments on Molten Salt Synthesis of Inorganic Nanomaterials: A Review. J. Phys. Chem. C 2021, 125, 6508–6533. [Google Scholar] [CrossRef]
  24. Li, L.; Deng, J.; Chen, J.; Xing, X. Topochemical molten salt synthesis for functional perovskite compounds. Chem. Sci. 2016, 7, 855–865. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, Y.; Zhu, S.; Zhang, Z.; Sun, Q.; Wu, C.; Gong, W.; Kang, L.; Yang, Y. Molten salt synthesis of high quality 2D δ-MnO2 nanosheets for advanced aqueous Zn/MnO2 batteries. J. Alloys Compd. 2023, 957, 170362. [Google Scholar] [CrossRef]
  26. Zuniga, J.P.; Abdou, M.; Gupta, S.K.; Mao, Y. Molten-Salt Synthesis of Complex Metal Oxide Nanoparticles. J. Vis. Exp. 2018, 140, e58482. [Google Scholar] [CrossRef]
  27. Huang, A.; Zhou, W.; Wang, A.; Chen, M.; Tian, Q.; Chen, J. Molten salt synthesis of α-MnO2/Mn2O3 nanocomposite as a high-performance cathode material for aqueous zinc-ion batteries. J. Energy Chem. 2021, 54, 475–481. [Google Scholar] [CrossRef]
  28. Rossi, F.; Marini, E.; Boniardi, M.; Casaroli, A.; Bassi, A.L.; Macrelli, A.; Mele, C.; Bozzini, B. What Happens to MnO2 When It Comes in Contact with Zn2+? An Electrochemical Study in Aid of Zn/MnO2-Based Rechargeable Batteries. Energy Technol. 2022, 10, 2200084. [Google Scholar] [CrossRef]
  29. Huang, H.; Chen, Q.; He, M.; Sun, X.; Wang, X. A ternary Pt/MnO2/graphene nanohybrid with an ultrahigh electrocatalytic activity toward methanol oxidation. J. Power Sources 2013, 239, 189–195. [Google Scholar] [CrossRef]
  30. Hazarika, K.K.; Goswami, C.; Saikia, H.; Borah, B.J.; Bharali, P. Cubic Mn2O3 nanoparticles on carbon as bifunctional electrocatalyst for oxygen reduction and oxygen evolution reactions. Mol. Catal. 2018, 451, 153–160. [Google Scholar] [CrossRef]
  31. Balamurugan, S.; Ashika, S.A.; Fathima, T.K.S. Moderate temperature production of manganese oxides via thermal treatment – Structural and thermal properties. Chem. Inorg. Mater. 2024, 2, 100039. [Google Scholar] [CrossRef]
  32. Boucher, B.; Buhl, R.; Perrin, M. Proprietes et structure magnetique de Mn3O4. J. Phys. Chem. Solids 1971, 32, 2429–2437. [Google Scholar] [CrossRef]
  33. Jamil, S.; Khan, S.R.; Sultana, B.; Hashmi, M.; Haroon, M.; Janjua, M.R.S.A. Synthesis of Saucer Shaped Manganese Oxide Nanoparticles by Co-precipitation Method and the Application as Fuel Additive. J. Clust. Sci. 2018, 29, 1099–1106. [Google Scholar] [CrossRef]
  34. Garces Gonzales, P.R., Jr.; De Abreu, H.A.; Duarte, H.A. Stability, Structural, and Electronic Properties of Hausmannite (Mn3O4) Surfaces and Their Interaction with Water. J. Phys. Chem. C 2018, 122, 20841–20849. [Google Scholar] [CrossRef]
  35. Lanson, B.; Drits, V.A.; Feng, Q.; Manceau, A. Structure of synthetic Na-birnessite: Evidence for a triclinic one-layer unit cell. Am. Mineral. 2002, 87, 1662–1671. [Google Scholar] [CrossRef]
  36. Zhang, X.; Qian, Y.; Zhu, Y.; Tang, K. Synthesis of Mn2O3 nanomaterials with controllable porosity and thickness for enhanced lithium-ion batteries performance. Nanoscale 2014, 6, 1725–1731. [Google Scholar] [CrossRef] [PubMed]
  37. Song, Y.L.; Wang, Z.C.; Yan, Y.D.; Zhang, M.L.; Wang, G.L.; Yin, T.Q.; Xue, Y.; Gao, F.; Qiu, M. Molten salt synthesis and supercapacitor properties of oxygen-vacancy LaMnO3. J. Energy Chem. 2020, 43, 173–181. [Google Scholar] [CrossRef]
  38. Takeuchi, K.J.; Yau, S.Z.; Menard, M.C.; Marschilok, A.C.; Takeuchi, E.S. Synthetic Control of Composition and Crystallite Size of Silver Hollandite, AgxMn8O16: Impact on Electrochemistry. ACS Appl. Mater. Interfaces 2012, 4, 5547–5554. [Google Scholar] [CrossRef]
  39. Brady, A.B.; Huang, J.; Durham, J.L.; Smith, P.F.; Bai, J.; Takeuchi, E.S.; Marschilok, A.C.; Takeuchi, K.J. The Effect of Silver Ion Occupancy on Hollandite Lattice Structure. MRS Adv. 2018, 3, 547–552. [Google Scholar] [CrossRef]
  40. Sánchez-Ochoa, F.; Springborg, M. Silver hollandite (AgxMn8O16, x ≤ 2): A highly anisotropic half-metal for spintronics. Phys. Rev. Mater. 2021, 5, 095001. [Google Scholar] [CrossRef]
  41. Li, L.; King, D.L. Synthesis and Characterization of Silver Hollandite and Its Application in Emission Control. Chem. Mater. 2005, 17, 4335–4343. [Google Scholar] [CrossRef]
  42. Chen, J.; Li, J.; Liu, Q.; Huang, X.; Shen, W. Facile Synthesis of Ag-Hollandite Nanofibers and Their Catalytic Activity for Ethanol Selective Oxidation. Chin. J. Catal. 2007, 28, 1034–1036. [Google Scholar] [CrossRef]
  43. Chen, J.; Li, J.; Li, H.; Huang, X.; Shen, W. Facile synthesis of Ag–OMS-2 nanorods and their catalytic applications in CO oxidation. Microporous Mesoporous Mater. 2008, 116, 586–592. [Google Scholar] [CrossRef]
  44. Sun, Y.; Hu, X.; Zhang, W.; Yuan, L.; Huang, Y. Large-scale synthesis of Ag1.8Mn8O16 nanorods and their electrochemical lithium-storage properties. J. Nanopart. Res. 2011, 13, 3139–3148. [Google Scholar] [CrossRef]
  45. Sukhdev, A.; Challa, M.; Narayani, L.; Manjunatha, A.S.; Deepthi, P.R.; Angadi, J.V.; Mohan Kumar, P.; Pasha, M. Synthesis, phase transformation, and morphology of hausmannite Mn3O4 nanoparticles: Photocatalytic and antibacterial investigations. Heliyon 2020, 6, e03245. [Google Scholar] [CrossRef]
  46. Amirtharaj, S.N.; Mariappan, M. Rapid and controllable synthesis of Mn2O3 nanorods via a sonochemical method for supercapacitor electrode application. Appl. Phys. A 2021, 127, 607. [Google Scholar] [CrossRef]
  47. Nagamuthu, S.; Ryu, K.S. Synthesis of Silver Hollandite Nanorectangular Cuboids as Negative Electrode Material for High-Performance Asymmetric Supercapacitors and Lithium-Ion Capacitors. Batter. Supercaps 2019, 2, 91–103. [Google Scholar] [CrossRef]
  48. Shrestha, N.K.; Patil, S.A.; Han, J.; Cho, S.; Inamdar, A.I.; Kim, H.; Im, H. Chemical etching induced microporous nickel backbones decorated with metallic Fe@hydroxide nanocatalysts: An efficient and sustainable OER anode toward industrial alkaline water-splitting. J. Mater. Chem. A 2022, 10, 8989–9000. [Google Scholar] [CrossRef]
  49. dos Santos, J.P.A.; Rufino, F.C.; Ota, J.I.Y.; Fernandes, R.C.; Vicentini, R.; Pagan, C.J.; Silva, L.M.D.; Zanin, H. Best practices for electrochemical characterization of supercapacitors. J. Energy Chem. 2023, 80, 265–283. [Google Scholar] [CrossRef]
  50. Cui, Z.; Wang, D.; Xu, T.; Yao, T.; Shen, L. Enabling extreme low-temperature proton pseudocapacitor with tailored pseudocapacitive electrodes and antifreezing electrolytes engineering. Chem. Eng. J. 2024, 495, 153347. [Google Scholar] [CrossRef]
  51. Mei, B.A.; Munteshari, O.; Lau, J.; Dunn, B.; Pilon, L. Physical Interpretations of Nyquist Plots for EDLC Electrodes and Devices. J. Phys. Chem. C 2017, 122, 194–206. [Google Scholar] [CrossRef]
  52. Nithya, V.D.; Selvan, R.K.; Vasylechko, L.; Sanjeeviraja, C. Surfactant assisted sonochemical synthesis of Bi2WO6 nanoparticles and their improved electrochemical properties for use in pseudocapacitors. RSC Adv. 2014, 4, 4343–4352. [Google Scholar] [CrossRef]
  53. Mathis, T.S.; Kurra, N.; Wang, X.; Pinto, D.; Simon, P.; Gogotsi, Y. Energy Storage Data Reporting in Perspective—Guidelines for Interpreting the Performance of Electrochemical Energy Storage Systems. Adv. Energy Mater. 2019, 9, 1902007. [Google Scholar] [CrossRef]
  54. Abdullin, K.A.; Gabdullin, M.T.; Kalkozova, Z.K.; Nurbolat, S.T.; Mirzaeian, M. Efficient Recovery Annealing of the Pseudocapacitive Electrode with a High Loading of Cobalt Oxide Nanoparticles for Hybrid Supercapacitor Applications. Nanomaterials 2022, 12, 3669. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, J.; Burke, A.F. Electrochemical Capacitors: Performance Metrics and Evaluation by Testing and Analysis. Adv. Energy Mater. 2020, 11, 2002192. [Google Scholar] [CrossRef]
  56. Martínez-Hincapié, R.; Wegner, J.; Anwar, M.U.; Raza-Khan, A.; Franzka, S.; Kleszczynski, S.; Colic, V. The determination of the electrochemically active surface area and its effects on the electrocatalytic properties of structured nickel electrodes produced by additive manufacturing. Electrochim. Acta 2024, 476, 143663. [Google Scholar] [CrossRef]
  57. Rivera-Benítez, A.; Luna-Sánchez, R.; Rafael-Alonso, A.; Arce-Estrada, E.; Cabrea-Sierra, R.; Henández-Ramírez, A.; Arellanes-Lozada, P.; Cuellar-Herrera, L.; López-Rodríguez, J. Enhancing the specific capacitance of α-MnO2 through quenching-induced changes in the crystal phase structure. Int. J. Electrochem. Sci. 2024, 19, 100609. [Google Scholar] [CrossRef]
  58. Saju, S.M.; Vargeese, A.A. Synthesis of perforated single-crystalline Mn2O3 microcubes for thermocatalytic applications. New J. Chem. 2024, 48, 17150–17158. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of Na-containing Mn x O y composites: 0.3 M/N, 0.4 M/N, 0.5 M/N, and 0.7 M/N.
Figure 1. XRD pattern of Na-containing Mn x O y composites: 0.3 M/N, 0.4 M/N, 0.5 M/N, and 0.7 M/N.
Materials 18 03869 g001
Figure 2. XRD pattern of Na/Ag-containing Mn x O y composites: 0.5 M/N-0.1A, 0.5 M/N-0.3A, 0.5 M/N-0.5A, 1 M/N-0.1A, 1 M/N-0.3A, and 1 M/N-0.5A.
Figure 2. XRD pattern of Na/Ag-containing Mn x O y composites: 0.5 M/N-0.1A, 0.5 M/N-0.3A, 0.5 M/N-0.5A, 1 M/N-0.1A, 1 M/N-0.3A, and 1 M/N-0.5A.
Materials 18 03869 g002
Figure 3. SEM micrographs at 80,000x and particle size histograms of Na/Ag-containing Mn x O y composites: (a) 0.3 M/N, (b) 0.5 M/N-0.5A, and (c) 1 M/N-0.5A. Elemental mapping of (d) 0.3 M/N, (e) 0.5 M/N-0.5A, and (f) 1 M/N-0.5A.
Figure 3. SEM micrographs at 80,000x and particle size histograms of Na/Ag-containing Mn x O y composites: (a) 0.3 M/N, (b) 0.5 M/N-0.5A, and (c) 1 M/N-0.5A. Elemental mapping of (d) 0.3 M/N, (e) 0.5 M/N-0.5A, and (f) 1 M/N-0.5A.
Materials 18 03869 g003
Figure 4. Electrochemical behavior of Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 by CV: (a) 5 mV s 1 ; scan rates of (b) 0.3 M/N and (c) 1 M/N-0.5A. Specific capacitance plots at different scan rates of Mn x O y systems containing (d) Na and (e) Ag. (f) Electrochemically active surface area (ECSA) of the different electrodes.
Figure 4. Electrochemical behavior of Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 by CV: (a) 5 mV s 1 ; scan rates of (b) 0.3 M/N and (c) 1 M/N-0.5A. Specific capacitance plots at different scan rates of Mn x O y systems containing (d) Na and (e) Ag. (f) Electrochemically active surface area (ECSA) of the different electrodes.
Materials 18 03869 g004
Figure 5. The GCD curves of (a) 0.3 M/N, 0.5 M/N-0.5A, and 1 M/N-0.5A systems at 0.5 A g 1 and (b) 1 M/N-0.5A system at different current densities in 0.1 M Na 2 SO 4 . (c) The cyclic stability test of 1 M/N-0.5A at current density of 1 A g 1 .
Figure 5. The GCD curves of (a) 0.3 M/N, 0.5 M/N-0.5A, and 1 M/N-0.5A systems at 0.5 A g 1 and (b) 1 M/N-0.5A system at different current densities in 0.1 M Na 2 SO 4 . (c) The cyclic stability test of 1 M/N-0.5A at current density of 1 A g 1 .
Materials 18 03869 g005
Figure 6. Electrochemical behavior of Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 by EIS: Nyquist plots of Mn x O y containing (a) Na and (b) Ag; (c) EEC used to fit the experimental data; (d) R c t and inclination angle values at low frequencies.
Figure 6. Electrochemical behavior of Na/Ag-containing Mn x O y composites in 0.1 M Na 2 SO 4 by EIS: Nyquist plots of Mn x O y containing (a) Na and (b) Ag; (c) EEC used to fit the experimental data; (d) R c t and inclination angle values at low frequencies.
Materials 18 03869 g006
Table 1. Synthesis conditions and abbreviations of Na/Ag-containing Mn x O y composites.
Table 1. Synthesis conditions and abbreviations of Na/Ag-containing Mn x O y composites.
Sample AbbreviationMnSO4·H2O/NaNO3MnSO4·H2O
(g)
NaNO 3
(g)
AgNO 3
(g)
NaOH
(g)
0.3 M/N0.33903
0.4 M/N0.44903
0.5 M/N0.55903
0.7 M/N0.76903
0.5 M/N-0.1A0.5360.11
0.5 M/N-0.3A0.5360.31
0.5 M/N-0.5A0.5360.51
1 M/N-0.1A1330.11
1 M/N-0.3A1330.31
1 M/N-0.5A1330.51
Table 2. Elemental composition of Na/Ag-containing Mn x O y composites by EDS.
Table 2. Elemental composition of Na/Ag-containing Mn x O y composites by EDS.
SampleONaMnAg
0.3 M/N69.535.6924.780.00
0.5 M/N-0.5Ag73.014.9618.513.53
1 M/N-0.5Ag72.223.419.345.04
Table 3. Results obtained from the 1 M/N-0.5A system in 0.1 M NaSO 4 at 0.5 A g 1 by GCD.
Table 3. Results obtained from the 1 M/N-0.5A system in 0.1 M NaSO 4 at 0.5 A g 1 by GCD.
I (A g 1 ) t s (s) C s , GCD (F g 1 )E (Wh kg 1 )P (W kg 1 )
0.5443.3225.131.3253.8
1.0159.8162.222.5507.5
1.533.751.37.1761.3
2.02.95.90.81015.0
Table 4. Impedance parameters of Na/Ag-containing Mn x O y composites in 0.1 M NaSO 4 .
Table 4. Impedance parameters of Na/Ag-containing Mn x O y composites in 0.1 M NaSO 4 .
Sample R el ( Ω ) C dl (mF) Y eq (mS s n ) n eq R ct ( Ω )Angle (°)
0.3 M/N0.08346.814.40.8468.267.0
0.4 M/N0.01720.33.30.7267.169.7
0.5 M/N0.86912.17.30.8672.569.1
0.7 M/N0.02118.04.10.6280.456.5
0.5 M/N-0.1A0.02735.85.40.8679.570.1
0.5 M/N-0.3A0.00328.96.10.9159.070.3
0.5 M/N-0.5A0.00249.19.90.8367.470.6
1 M/N-0.1A0.00018.88.00.8171.163.4
1 M/N-0.3A0.05057.57.90.7769.162.6
1 M/N-0.5A0.08341.722.00.9560.681.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Martínez-Morales, C.; Romero-Serrano, A.; López-Rodríguez, J.; Arellanes-Lozada, P. Molten Salt Synthesis and Electrochemical Evaluation of Na/Ag-Containing MnxOy Composites for Pseudocapacitor Applications. Materials 2025, 18, 3869. https://doi.org/10.3390/ma18163869

AMA Style

Martínez-Morales C, Romero-Serrano A, López-Rodríguez J, Arellanes-Lozada P. Molten Salt Synthesis and Electrochemical Evaluation of Na/Ag-Containing MnxOy Composites for Pseudocapacitor Applications. Materials. 2025; 18(16):3869. https://doi.org/10.3390/ma18163869

Chicago/Turabian Style

Martínez-Morales, Carmen, Antonio Romero-Serrano, Josué López-Rodríguez, and Paulina Arellanes-Lozada. 2025. "Molten Salt Synthesis and Electrochemical Evaluation of Na/Ag-Containing MnxOy Composites for Pseudocapacitor Applications" Materials 18, no. 16: 3869. https://doi.org/10.3390/ma18163869

APA Style

Martínez-Morales, C., Romero-Serrano, A., López-Rodríguez, J., & Arellanes-Lozada, P. (2025). Molten Salt Synthesis and Electrochemical Evaluation of Na/Ag-Containing MnxOy Composites for Pseudocapacitor Applications. Materials, 18(16), 3869. https://doi.org/10.3390/ma18163869

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