Open Access Article
Ayman E. Elkholy
ac,
F. El-Taib Heakal
b and
Nageh K. Allam
*c
aDepartment of Analysis and Evaluation, Egyptian Petroleum Research Institute, 11727 Cairo, Egypt
bChemistry Department, Faculty of Science, Cairo University, 12613 Giza, Egypt
cEnergy Materials Laboratory, School of Sciences and Engineering, The American University in Cairo, 11835 New Cairo, Egypt. E-mail: nageh.allam@aucegypt.edu
First published on 8th November 2017
We report on the synthesis of manganese cobalt ferrite (MnCoFeO4) nanoparticles via a simple one-pot co-precipitation method and their characterization through energy-dispersive spectroscopy (EDS), X-ray diffraction (XRD), high-resolution transmission electron microscopy (HR-TEM), Fourier transform infrared (FT-IR) spectroscopy and N2 adsorption/desorption techniques. The MnCoFeO4 supercapacitor showed the maximum specific capacitance of 675 F g−1 at a scan rate of 1 mV s−1. Its energy and power densities were 18.85 W h kg−1 and 337.50 W kg−1, respectively, at a current density of 1.5 A g−1. The cyclic stability was scrutinized via galvanostatic charging/discharging (GCD) and electrochemical impedance spectroscopy (EIS). The degradation of the supercapacitive performance was only 7.14% after 1000 GCD cycles, indicating an excellent long-term stability. The equivalent series resistance (ESR) remained nearly constant even after 1000 GCD cycles.
According to their active materials, supercapacitors (SCs) are classified into two main types: electrochemical double layer capacitors (DCs) and electrochemical pseudocapacitors (PCs).7,8 The active material in the DCs is composed of carbonaceous (carbon-based) materials, such as activated carbon, graphite, and graphene. Energy is stored physically within a DC via charge accumulation across the electrode/electrolyte interface.8,9 As for PCs, the active material is primarily composed of a transition metal (TM) oxide, a TM nitride or a conducting polymer. Energy is stored electrochemically within a PC via the reversible interfacial redox reactions in TMs or via ion intercalation throughout the electrode in conducting polymers.5,10,11 Due to the intrinsic low specific capacitance (Csp) and low energy density stored in the current SCs, it is vital to explore new materials that simultaneously exhibit high Csp as well as high conductivity.4 TMs are characterized by being cost-effective and displaying multiple oxidation states7 in addition to their fast and reversible faradaic redox reactions.12 Therefore, TMs are engrossed as the electrode materials for supercapacitor applications because of their outstanding electrochemical performance.13 In addition, TM oxides can provide a higher theoretical Csp than that of the conventional carbon-based materials and a better electrochemical stability than that of the polymeric materials.7 Among them, RuO2 has been widely investigated as a promising candidate because of its better conductivity and high Csp; however, it is limited by its high cost, rarity,13,14 and toxicity.15 Alternative inorganic electrode materials such as MnO2,16 Co3O4,17 NiO,18 V2O5/VO4,19 WO3,20 and Fe2O3
21 have been intensively investigated in SC applications owing to their wide availability, chemical stability, mechanical strength, safety, and eco-friendliness.15 Recently, ferrites as SC materials have been explored by researchers due to their various redox states, electrochemical stability as well as remarkable magnetic, catalytic, optical, and electrical properties.22 Ferrite-based materials have been synthesized in diverse nanostructured forms, including nanoparticles,23 nanotubes,24 nanofibers,20 nanowires,25,26 nanorods,10 nanoflakes,17,27 nanomesh arrays,14 nanosheets,17 and hollow structures.28 Ferrites can be synthesized via various synthesis techniques including sol–gel method,12 co-precipitation method,29 template method,24 solvothermal method,30 microwave-assisted method,10,13 electrodeposition method,27 chemical spray method,15 spray-pyrolysis method15 and hydrothermal method.14 Among these techniques, the co-precipitation method provides a simple route for the one-pot synthesis and the reaction conditions are mild and simple.13
Usually, the general formula for a spinel is AB2O4, where A refers to a divalent metal ion (M2+) and B refers to a trivalent metal ion (M3+). In a normal spinel structure, the A ions occupy the tetrahedral sites and the B ions occupy the octahedral sites.31 Spinel ferrites, MFe2O4 or MM′FeO4 (where M or M′= Mn, Co, Ni, Zn, Cu, etc.), are fascinating materials owing to their impressive magnetic, electrical, and optical properties in addition to their ability to exhibit different redox states and electrochemical stability. In spinel ferrites, the divalent metal ion (M2+) occupies the tetrahedral site and the trivalent metal ion (M′3+ or Fe3+) occupies the octahedral position.12 According to Bernard et al.32 and Kulkarni,33 Mn atoms in MnCoFeO4 prefer to occupy the tetrahedral positions, while Co atoms prefer to occupy the octahedral sites. Besides, the valence states of Mn, Co, and Fe are majorly (+II), (+III), and (+III), respectively. In contrast, Martens34 has studied the magneto-optical properties of MnxCoFe2−xO4 (where x = 0, 0.5, and 1.0) prepared using a conventional ceramic technology with a final sinter treatment in oxygen for 24 h at 1200–1300 °C and found that Mn has the oxidation state (+III) and occupied the octahedral sites, while Co, in the form of Co2+, could occupy both octahedral and tetrahedral positions with the latter being more favorable. These observations were confirmed by two later studies based on submicron MnxCoFe2−xO4 spinel ferrites by Chassaing et al.35 and Laarj and Kacim.36 In both studies, MnxCoFe2−xO4 spinels were prepared from the oxalic precursors and subjected to annealing treatment between 600 °C and 700 °C.
Numerous binary TM ferrites have been investigated for supercapacitor applications such as ZnFe2O4,17,30 CoFe2O4,14 MnFe2O4,37 CuFe2O4,26 and SnFe2O4.10 However, few research studies were carried out on ternary TM ferrites as the electrodes for supercapacitor applications as listed in Table 1.
Herein, MnCoFeO4 was synthesized in the form of nanoparticles via the co-precipitation method and characterized by EDS, XRD, TEM, FT-IR analyses, and N2 adsorption/desorption. Its supercapacitive performance in 6 M KOH was investigated via cyclic voltammetry (CV) and galvanostatic charging/discharging (GCD). The cyclic stability was studied via GCD and electrochemical impedance spectroscopy (EIS). The materials showed an exceptional Csp as compared to the Csp of those reported in the literature.
:
20
:
10. The prepared slurry was stirred for about 48 h at room temperature, then coated onto a part of a 1 cm × 2 cm chip of nickel foam (as a supporter and current collector) and dried at 60 °C. The supercapacitive performance of the prepared MnCoFeO4 electrode was investigated using a three-electrode cell containing the working electrode (the prepared electrode under study), a counter electrode (a Pt coil) and a reference electrode (a saturated calomel electrode, SCE) in a 6 M KOH solution at room temperature. The electrochemical measurements involved cyclic voltammetry (CV), galvanostatic charging/discharging (GCD), and electrochemical impedance spectroscopy (EIS) using the electrochemical workstation (CHI 760D, CH Instruments, U.S.A.). CV measurements were performed within the potential window of 0–0.45 V at different scan rates (from 1 to 100 mV s−1). GCD measurements were carried out at different current densities (1.5–10 A g−1) within the same potential window. The EIS measurements were performed in the frequency range of 100 kHz to 0.1 Hz at the steady-state open circuit potential (0.261 V vs. SCE) with a sinusoidal perturbation amplitude of 10 mV. EIS parameters were derived using EC-Lab V10.40 software.
The specific capacitance (Csp, F g−1) was then calculated from data obtained from CV and GCD measurements according to eqn (1)10,22 and (2)14,17,38 respectively.
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
eqn (4) can be reformed as follows:
![]() | (5) |
![]() | (6) |
:
1
:
2
:
6. Fig. 1b shows the XRD pattern of the prepared material. All diffraction peaks are coincident to the Miller indices (111), (220), (222), (400), (422), (511), (440) and (622) of the standard data for MnCoFeO4 (cubic, space group: Fd
m, ICDD card number: 04-010-1895), which confirms the formation of MnCoFeO4 with the spinel structure. Fig. 1c displays the FT-IR spectrum of MnCoFeO4 recorded in the frequency range 4000–400 cm−1. The lower-frequency band (493 cm−1) is assigned to the octahedral groups ([Fe3+–O2−] and [Mn3+–O2−]), while the higher-frequency band (610 cm−1) is assigned to the stretching of the tetrahedral groups ([Co2+–O2−]) present in the spinel ferrite.12 The FT-IR absorption bands appearing at 3383 cm−1 and 1625 cm−1 are referred to adsorbed water molecules.41
Fig. 2 shows the HR-TEM images of the as-fabricated MnCoFeO4, indicating the formation of nanoparticles with sizes ranging ca. 30–40 nm. Fig. 3a shows the N2 adsorption/desorption isotherm for the as-synthesized MnCoFeO4. According to the IUPAC classification of gas adsorption/desorption isotherms,43 the obtained N2 adsorption/desorption isotherm is of type IV, which is characteristic to the mesoporous materials. The characteristic feature of this isotherm is its hysteresis loop, which arises when the adsorption and desorption curves do not coincide and is associated with the capillary condensation taking place in the mesopores. Moreover, the initial part, where adsorption/desorption curves are coincident, is attributed to the monolayer-multilayer adsorption. These features confirm the mesoporous structure of the prepared MnCoFeO4. It is worth mentioning that porous materials are generally classified according to their pore diameter (d) into three categories: macroporous (d > 50 nm), mesoporous (d = 2–50 nm), and microporous materials (d < 2 nm).5 Fig. 3b shows the pore diameter distribution curve obtained by the Barrett–Joyner–Halenda (BJH) method using the desorption branch of the nitrogen isotherm.17 It reveals an average pore diameter of 10.036 nm. The BET method likely yields a value of actual surface area if the isotherm is either of type II or type IV.43 Hence, the surface area obtained from multipoint BET is noted to be 104.963 m2 g−1.
![]() | ||
| Fig. 3 (a) N2 adsorption/desorption isotherm and (b) BJH pore size distribution for the as-prepared MnCoFeO4. | ||
Moreover, Fig. 4a demonstrates the pairs of distinct and broad redox peaks corresponding to the redox transitions of Co3+/Co2+ and Mn3+/Mn2+. The redox peaks of Co3+/Co2+ and Mn3+/Mn2+ seem to be merging together because their standard electrode potentials are comparable (1.92 and 1.50 VNHE, respectively45). The probable processes associated with the capacitive behavior of MnCoFeO4 can be related to the presence of two redox systems. The redox reactions related to both systems could proceed according to eqn (7)46 and (8)42, respectively.
| Co2+/Co3+ system: CoFe2O4 + OH− + H2O ↔ 2FeOOH + CoOOH + e | (7) |
| Mn2+/Mn3+ system: MnFe2O4 + OH− + H2O ↔ 2FeOOH + MnOOH + e | (8) |
The CV profiles (at different scan rates) demonstrate the same behavior, revealing the reversibility of the redox reactions. In addition, with the increase in the scan rate, both oxidation and reduction peaks shift towards more anodic and cathodic directions, respectively. For example, upon increasing the scan rate from 1 to 100 mV s−1, the anodic peak potential (Ep,a) increased from 0.358 to 0.378 VSCE, while the anodic peak current (Ip,a) increased from 2.063 to 52.494 (A g−1). This effect results from the unavoidable overpotential due to the increase in the internal diffusion resistance and Ip increases with υ due to the fast interfacial kinetics.14
Galvanostatic charge/discharge (GCD) measurement is an accurate technique for determining the electrochemical performance of supercapacitors particularly for those based on pseudo-capacitance.40 Fig. 4b displays the GCD curves recorded for MnCoFeO4 in 6 M KOH. Unlike the GCD curves of carbon-based materials exhibiting a semi-triangular shape, where their capacitances are mainly attributed to a pure electric double layer capacitance, the GCD curves of MnCoFeO4 show deviations from linearity due to its pseudocapacitive nature.22,47 Clearly, the MnCoFeO4 electrode demonstrates higher charging and discharging times as the current density decreases, resulting in higher values of specific capacitance (Fig. 4c). For example, on decreasing the current density from 10 A g−1 to 1.5 A g−1, Csp increases from 112 to 670 F g−1. Furthermore, the columbic efficiency (η%) increases with the current density up to 89% at 4 A g−1, after which it becomes constant (∼81%).
Power density (PD) and energy density (ED) are two important parameters for the evaluation of the electrochemical performance of supercapacitors and electrode materials.40 Fig. 5a displays the Ragone plot of MnCoFeO4 obtained from GCD measurements48 with the current densities ranging from 1.5 to 10 A g−1. Fig. 5a demonstrates that the ED stored in MnCoFeO4 and its corresponding PD (at a current density of 1.5 A g−1) are 18.85 W h kg−1 and 337.50 W kg−1, respectively. On the other hand, at a current density of 10 A g−1, ED and PD values are 3.15 W h kg−1 and 2250.00 W kg−1, respectively. Table 1 depicts the supercapacitive performance parameters, including Csp, PD, and ED for MnCoFeO4 and some ternary and quaternary TM ferrites. It should be noted that MnCoFeO4 has an excellent Csp and PD in comparison to those of the previously investigated materials.
EIS is a powerful tool used to investigate the features of an electrode/electrolyte interface as supercapacitors by evaluating the frequency behavior and equivalent series resistance (ESR).37 EIS offers information about the internal resistance of the electrode material and the resistance between the electrode and the electrolyte.40 EIS spectra (Fig. 5c) display a depressed semicircle in the high-frequency region linked to an inclined straight line in the low-frequency region. The slope of the inclined line is decreased after 1000 GCD cycles. The equivalent circuit used to analyze EIS spectra is depicted in the inset of Fig. 5c.48
The time constant (QRctW) represents the depressed semicircle in the high-frequency region, where Rct is the charge-transfer resistance caused by the faradaic process and Q is a constant phase element representing the non-ideal double-layer capacitance. Table 2 reveals that Rs, representing the ESR, remains nearly constant even after 1000 GCD cycles. The ESR includes electrolyte resistance, internal resistance of the electrode and contact resistance between the electrode and the current collector.5 Rct suffers a slight increase from 53.37 Ω to 78.74 Ω, indicating that the amount of charge stored in this material is slightly decreased. The linear part of the curve is inclined by ∼45°, indicating Warburg impedance (W) that is related to the frequency-dependent diffusion resistance of electrolyte ions.15,17,47 The value of W increases from 59.67 Ω s−0.5 to 101.90 Ω s−0.5, indicating a slowdown in the diffusion rate due to a somewhat difficult penetration of electrolyte ions through the electrode mesopores with long-term use.
| Fresh electrode | After 1000 cycles | |
|---|---|---|
| Rs (Ω) | 1.74 | 1.76 |
| Q (mF) | 2.68 | 2.57 |
| n | 0.86 | 0.87 |
| Rct (Ω) | 53.37 | 78.74 |
| W (Ω s−0.5) | 59.67 | 101.90 |
| C (mF) | 35.10 | 84.90 |
| This journal is © The Royal Society of Chemistry 2017 |