Next Article in Journal
Investigation of Scaling and Materials’ Performance in Simulated Geothermal Brine
Previous Article in Journal
The Gypsum Influence on the Formation of Secondary Phases During Autoclave Leaching of Gold-Bearing Concentrates and the Silver Recovery Using Cyanidation
Previous Article in Special Issue
Friction and Wear Behavior of Double-Walled Carbon Nanotube-Yttria-Stabilized ZrO2 Nanocomposites Prepared by Spark Plasma Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic CuFe2O4 Spinel–Polypyrrole Pseudocapacitive Composites for Energy Storage

Department of Materials Science and Engineering, McMaster University, Hamilton, ON L8S 4L7, Canada
*
Author to whom correspondence should be addressed.
Materials 2024, 17(21), 5249; https://doi.org/10.3390/ma17215249
Submission received: 2 October 2024 / Revised: 23 October 2024 / Accepted: 26 October 2024 / Published: 28 October 2024

Abstract

:
This investigation focused on the fabrication of ceramic ferrimagnetic CuFe2O4–conductive polypyrrole (PPy) composites for energy storage. CuFe2O4 with a crystal size of 20–30 nm and saturation magnetization of 31.4 emu g−1 was prepared by hydrothermal synthesis, and PPy was prepared by chemical polymerization. High-active-mass composite electrodes were fabricated for energy storage in supercapacitors for operation in a sodium sulfate electrolyte. The addition of PPy to CuFe2O4 resulted in a decrease in charge transfer resistance and an increase in capacitance in the range from 1.20 F cm−2 (31 F g−1) to 4.52 F cm−2 (117.4 F g−1) at a 1 mV s−1 sweep rate and from 1.17 F cm−2 (29.9 F g−1) to 4.60 F cm−2 (120.1 F g−1) at a 3 mA cm−2 current density. The composites showed higher capacitance than other magnetic ceramic composites of the same mass containing PPy in the same potential range and exhibited improved cyclic stability. The magnetic behavior of the composites was influenced by the magnetic properties of ferrimagnetic CuFe2O4 and paramagnetic PPy. The composites showed a valuable combination of capacitive and magnetic properties and enriched materials science of magnetic supercapacitors for novel applications based on magnetoelectric and magnetocapacitive properties.

1. Introduction

Ceramic–conductive polymer composites have attracted rising attention due to their advanced multifunctional properties and interesting interface effects, which result in enhanced magnetoresistance, conductivity, magnetization, charge storage, and other properties [1,2,3,4]. Moreover, the addition of conductive polymers to ceramics improves the processability of ceramic–polymer nanocomposites [4]. Conductive polymers show valuable capacitive properties for energy storage applications [5,6,7,8,9,10]. (LaSr)MnO3 (LSMO)–polyaniline (PANI) composites showed enhanced conductivity and magnetoresistance in [1]. It was demonstrated that the synergistic effects in the composites resulted from the influence of the PANI coating on the LSMO particles on the electric transport and magnetoelectric phenomena in the magnetic fields. A magnetoresistance increase of ~73% was reported in the composites containing LSMO particles coated with PANI [11]. Polypyrrole (PPy) coatings on LSMO particles were developed in [3], and such coatings influenced the charge transfer, double-exchange interactions, and magnetization. Of particular interest was the significant enhancement in magnetization from 49.4 to 102.0 emu g−1 of the core–shell particles due to the decreased spin disorder in the LSMO core [3]. The magnetization increase resulted from the transfer of charge and spin from the conductive polymer layer to the LSMO. An enhancement of saturation magnetization was also reported for PPy-coated Fe3O4 particles and attributed to charge transfer, which increased the surface magnetic moment of the particles [12]. An investigation of the charge storage properties of LSMO-PPy composites showed a considerable increase in electrochemical capacitance compared with pure PPy [13]. Magnetoelectric composites have been developed for diverse applications in solar cells, memory devices, and sensors [4]. CoFe2O4-PPy composites showed advanced microwave absorption properties in [14]. ZnO-PEDOT composites exhibited promising properties for applications in electrochromic displays in [15].
Efficient CuO-PPy anodes were developed for energy storage in lithium-ion batteries in [16,17,18,19]. The composites offered the advantages of increased conductivity and capacity. Moreover, the detrimental volume changes related to charging and discharging were reduced. CuO-PPy composites showed promising performance for applications in supercapacitors with enhanced capacitance [20,21,22]. The investigations in [23] revealed a synergy of the charge storage properties of CuO and PPy, which led to the high capacitance of the composites. Hexagonal ferrite [24] and spinel [25] ferrimagnetic ceramics were combined with PPy for the fabrication of advanced supercapacitor composite anodes. In addition to high capacitance, such composites exhibited valuable magnetic properties.
The progress achieved in the development of ceramic–conductive polymer composites indicates that further advances in this area can result in the development of novel advanced composites containing various functional ceramic materials. Composites exhibit valuable functional properties, such as advanced mechanical properties, photovoltaic properties, high conductivity, and advanced mechanical and electrical properties, which have been utilized for novel applications [26,27,28,29,30,31,32]. CuFe2O4 is a promising multifunctional material for the development of ceramic–polymer composites. CuFe2O4 is a spinel ferrimagnetic compound [33,34] that exhibits high magnetization due to superexchange interactions between Cu2+ and Fe3+ ions. It is an advanced material for biomedicine [35], catalytic applications [36], water purification [37], photovoltaic [38] devices, and energy storage applications in supercapacitors [39,40,41,42,43]. Therefore, the development of CuFe2O4–conductive polymer composites can result in materials with advanced functionality.
This investigation was motivated by the interesting multifunctional properties of ceramic–PPy composites. Moreover, the research motivation was fueled by the multifunctional properties of CuFe2O4 and recent advances in the development of multifunctional materials [44], combining high magnetization and remarkable pseudocapacitive properties. Such materials show novel magnetoelectric effects, which are currently under consideration for the fabrication of novel multifunctional devices. Of particular interest are the results regarding capacitance increases and charge transfer resistance decreases in a magnetic field, electric potential-controlled magnetization reversal, electric potential-controlled shift of the Curie point, periodic variations in magnetization by periodic galvanostatic charge–discharge, electric field control of saturation magnetization, and other magnetocapacitive and magnetoelectric effects [44]. Different mechanisms that provide a direct link between magnetization and redox reactions were described in [44]. As pointed out above, the combination of pseudocapacitive conductive polymers and ceramic magnetic particles generates new effects related to spin transfer from the paramagnetic polymers to the magnetically ordered ceramic materials, which results in enhanced magnetization.
The objective of this work was the fabrication and testing of new CuFe2O4-PPy composites. In contrast with other studies on magnetic complex oxide ceramic–PPy composites [24,25], which involved high-energy ball milling for the particle size reduction of the ceramic phase, the fabrication of CuFe2O4 nanoparticles was performed by hydrothermal synthesis. The composite materials combined high magnetic and pseudocapacitive properties due to the synergy of the contributions of CuFe2O4 and PPy. It was found that the magnetic properties of the composites were influenced by the weakly paramagnetic PPy. The combination of CuFe2O4 and PPy resulted in improved capacitance and cyclic stability.

2. Materials and Methods

Iron (III) chloride hexahydrate (FeCl3·6H2O), copper (II) chloride dihydrate (CuCl2·2H2O), trisodium citrate dihydrate (C6H9Na3O9) (TSCD), NaOH, polyvinyl butyral (PVB), pyrrole (Py), ammonium persulfate (APS), and 4,5-dihydroxy-1,3-benzenedisulfonic acid disodium salt monohydrate (Tiron) were supplied by Aldrich. Multiwalled carbon nanotubes (MWCNTs) (with outer and inner diameters of 12 and 5 nm, respectively, and an average length of 1.5 µm) were provided by Bayer, Germany. Ni foams (with a thickness of 1.5 mm and 95% porosity) were supplied by Vale, Ontario, Canada.
CuFe2O4 (CFO) and PPy were used for the fabrication of composite materials. CFO was selected due to its high magnetization and promising pseudocapacitive properties. PPy is an advanced pseudocapacitive material that offers the benefit of high conductivity.
The synthesis of CFO nanoparticles was performed with 7 mM of FeCl3 and 3.3 mM of CuCl2 as precursors. Both were dissolved in 50 mL of DI water. Then, TSCD was added as a chelating agent to the solution at 15% according to the weight of the precursors. The solution was stirred until completely dissolved. An aqueous solution of NaOH was used to adjust the pH to 12. The hydrothermal process was conducted for 8 h at 180 °C. The product was washed with DI water and ethanol several times to remove the unwanted by-products and dried for 12h at 55 °C.
The synthesis of PPy was achieved with a 0.1 M Py reagent added to DI water cooled with an ice bath to 4 °C. Tiron was added as a dopant to the solution with a Py:Tiron molar ratio of 10:1. Then, 0.1 M APS was added dropwise as an oxidizing agent. The mixture was kept under stirring for 2 h to complete the chemical reaction. Subsequently, the Tiron-doped PPy was washed with DI water to remove excess materials and dried until a constant mass was reached.
The synthesis methods used in this investigation facilitated the fabrication of stable slurries of PPyand CFO. The sedimentation tests showed the good colloidal stability of the slurries in ethanol, which facilitated improved the mixing of the individual components. Moreover, the use of Tiron as a dopant for PPy was beneficial for the improved mixing of PPy and CFO. It is known that catecholate-type materials, such as Tiron, provide strong chelating bonding to metal atoms on the surface of inorganic nanoparticles [45]. The slurries in ethanol for the impregnation of Ni foam current collectors contained the materials in a ratio of (CFO + PPY):MWCNT:PVB = 80:20:3. CFO100 samples were prepared without PPy as a control sample. Different composite materials were prepared by variations in the PPy and CFO contents in their composition to analyze the influence of the electrode composition on their properties. The preparation of electrodes with the same active mass loading and same thickness was important for the comparison of their properties. However, the limitations of this preparation method resulted from the low density of PPy. As a result, the maximum content of the PPy phase in the electrodes of the same mass and thickness was 70%. The samples CFO90, CFO70, CFO50, and CFO30 contained CFO and PPy. The CFO:PPy mass ratios were 90:10, 70:30, 50:50, and 30:70 for CFO90, CFO70, CFO50, and CFO30, respectively. The mass of the impregnated materials after drying was 40 mg cm−2. To increase physical contact, the impregnated foams were pressed to 37% of their original thickness. All electrodes had an area of 1 cm2.
Transmission Electron Microscopy (TEM) analysis was conducted using a Talos L120C system, while X-ray diffraction (XRD) patterns were recorded with a Bruker D8 Advance diffractometer (Cu-Kα radiation). Scanning Electron Microscopy (SEM) imaging was carried out with a JSM 6610LV (JEOL, Tokyo, Japan). Electrochemical investigations, including cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and galvanostatic charge–discharge (GCD) studies, were performed using a VMP 300 potentiostat (BioLogic, Seyssinet-Pariset, France). The magnetic data were obtained with MPMS SQUID magnetometer (Quantum Design, San Diego, CA, USA). The simulations of the EIS data were performed using the ZSimpWin software v3.6 (AMETEK, Oak Ridge, TN, USA). All electrochemical tests of the electrodes were conducted in a three-electrode setup using a 0.5 M Na2SO4 aqueous electrolyte. A saturated calomel electrode (SCE) served as the reference electrode, while a platinum gauze was used as the counter electrode. Equations (1) and (2) were used to calculate the integral capacitances (CS and Cm) by normalizing the total capacitance (C) by the electrode area and mass, respectively [46]:
C = Δ Q Δ U = | 0 t ( U m a x ) I d t | + | t ( U m a x ) 0 I d t | 2 U m a x
where I is the current, ΔQ is the charge, and ΔU is the potential range, and from the chronopotentiometry data:
C = IΔt/ΔU
The complex capacitance C*(ω) = C′(ω) − iC″(ω) was derived from the EIS data at various frequencies (ω) of complex impedance, where Z*(ω) = Z′(ω) + i Z″(ω) [47]:
C ω = Z ( ω ) ω | Z ( ω ) | 2
C ω = Z ( ω ) ω | Z ( ω ) | 2

3. Results and Discussion

The XRD diffraction pattern of the as-prepared material (Figure 1A) confirmed the formation of a pure spinel phase of CuFe2O4. The diffraction angles and peak intensities corresponded to JCPDS file 00-034-0425. The relatively wide peaks indicated a small particle size. The magnetic measurements (Figure 1B) showed a saturation magnetization of 31.4 emu g−1, which was higher than the saturation magnetization of CuFe2O4 prepared by some other methods [48,49]. It is known that the magnetization of CuFe2O4 depends on the occupancy of the tetrahedral and octahedral positions of the spinel crystallographic unit cells by the Cu2+ and Fe3+ ions. The literature data for this material show that saturation magnetization can be in the range of 1–2.3 µB per molecule [34]. The saturation magnetization of magnetic materials usually decreases with decreasing nanoparticle size [50]. The analysis of magnetization versus magnetic field dependence in low fields (Figure 1B, inset) did not show hysteresis. The zero net magnetic moment in the absence of a magnetic field indicated a superparamagnetic behavior, which can result from a small crystal size.
The TEM studies (Figure 2) showed that the hydrothermal synthesis resulted in the formation of CuFe2O4 nanoparticles with a size of 20–30 nm. The TEM images showed well-defined crystal faces. The high-resolution TEM image demonstrated the crystallinity of the obtained material. The small size of the CuFe2O4 particles was beneficial for the fabrication of composites.
The composite materials were tested for applications in supercapacitor electrodes. Figure 3 compares the electrochemical testing results for the different composites, and the results are summarized in Table 1. It is obvious that the CV areas increased with the increasing PPy content in the composites (Figure 3A), which indicated a capacitance increase (Figure 3B). The CVs did not show redox peaks, which are typical for battery-type behavior [51,52]. The charge storage mechanism of CuFe2O4 is related to the reduction in Fe3+ and Cu2+ ions [40]. The charge and discharge times also increased with the increasing PPy content (Figure 3C) due to the capacitance increase. The electrodes showed good capacitance retention with the increasing sweep rate and current density (Figure 3 and Table 1). CFO50 and CFO30 showed CS values at 1 mV s−1 of 3.60 and 4.52 F cm−2, respectively, which were the highest capacitances of the composites studied in this investigation. The Cs values at 3 mA cm−2 of 3.90 and 4.60 F cm−2 were achieved for CFO50 and CFO30, respectively. As pointed out above, the fabrication of electrodes with the same thickness and same active mass loading presented difficulties for higher PPy content due to the lower density of PPy compared with the density of CuFe2O4.
Figure 4 shows the EIS data for the different composites. The Nyquist plot (Figure 4A) shows the impedance measurement data together with the simulation results obtained using the equivalent circuit presented in Figure 4B. The impedance decreased with the increasing PPy content. The real part of the impedance decreased due to the increased conductivity. The reduction in the imaginary part resulted from the increased capacitance. The equivalent circuit used for the simulation was developed for high-active-mass loading bulk electrodes [53]. It is known that porous electrodes can be described by transmission lines containing several RC elements [54,55,56,57]. The circuit presented in Figure 4B contains two RC(RQ) elements, where R, C, and Q represent the resistances, capacitance, and capacitive constant phase element, respectively. RE, RCT, and CDL represent the electrolyte resistance, charge transfer resistance, and double-layer capacitance, respectively. The C′ increased with increasing PPy content in the composites. The relaxation frequency, given by the frequency of the C″ maximum, was relatively high; however, it decreased with increasing PPy content in the composites. The high capacitance retention observed in the CV and GCD data correlated with the relatively high EIS relaxation frequencies.
The analysis of the EIS data showed a significant reduction in the charge transfer resistance with the increasing PPy content in the range of 0–20% (Figure 5). As a result, at higher PPy contents, a significant capacitance increase was observed with increasing PPy concentration (Figure 5). This can result from the enhanced charge transfer at the interface of CuFe2O4 and PPy and the high capacitance of PPy. Moreover, PPy created an additional conductive percolating network in the composites. One of the problems limiting PPy applications in supercapacitor devices containing neutral electrolytes is poor cyclic stability [58]. A significant reduction in capacitance was reported during the first 1000 cycles in other composites containing magnetic particles and PPy [24]. This resulted from the poor cyclic stability of PPy. However, the investigation of the composites with high PPy contents revealed their good cyclic stability. Figure 6 shows the cycling stability test results for CFO50 and CFO30. The capacitance increased slightly at the beginning of cycling and remained nearly constant. The capacitance retention was 104.9% and 104.3% for CFO50 and CFO30, respectively, after 2000 cycles. The slight initial increase in the capacitance could be attributed to the microstructural changes within the electrode bulk.
Figure 7 presents SEM images of the surface microstructures of the CFO50 and CFO30 composite electrodes, which showed the highest capacitances. The electrodes showed porous microstructures, which was beneficial for the electrolyte’s diffusion. The typical pore size was <1 µm. The electrode porosity resulted from the packing of CuFe2O4, PPy, and MWCNT.
The relatively high capacitance of CFO50 and CFO30 is promising for energy storage in supercapacitors. Such composites showed higher capacitances compared with the composites containing other spinel and hexagonal ferrite ferrimagnetic ceramics with the same mass loading in the same potential range [24,25]. Moreover, CFO50 and CFO30 showed relatively high magnetization. Figure 8 shows the M–H curves plotted from the vibration magnetometer data. The CFO50 and CFO30 composites showed saturation magnetization values of 17.5 and 7.6 emu g−1, respectively. Those values were lower than the CFO100 magnetization of 31.5 emu g−1 due to the lower CuFe2O4 content in the electrode. The coercivity values (Hc) from the hysteresis curves indicate that the CFO50 and CFO30 samples showed soft ferrimagnetic behavior, with Hc values between 50 and 100 Oe and a small remanent magnetization. This can result from the interaction of CuFe2O4 with the paramagnetic PPy [59,60]. The results of this study show that the properties of CuFe2O4 composites can be varied. Such composites belong to the category of magnetic pseudocapacitors [44], which are promising for various applications based on magnetoelectric and magnetocapacitive effects. The observation of capacitive properties in high-active-mass composites depends on using advanced Ni foam current collectors, which offer high corrosion resistance, conductivity, and porosity. However, the investigation of magnetocapacitive effects in electrodes prepared in this investigation presented difficulties since Ni is ferromagnetic, which generates its own magnetic field in an external magnetic field. Thus, non-magnetic current collectors with similar properties are needed. It should be noted that the composite CuFe2O4 spinel–PPy materials prepared in this investigation showed higher capacitances compared with other polymer composites containing various ferrimagnetic spinel compounds [40,61,62,63].

4. Conclusions

Spinel CuFe2O4 nanoparticles with a crystal size of 20–30 nm and saturation magnetization of 31.4 emu g−1 were prepared by hydrothermal synthesis and used for the fabrication of CuFe2O4-PPy composites. The maximum PPy content in the composite electrodes of the same mass and thickness was 70%. The addition of PPy to CuFe2O4 reduced the charge transfer resistance and improved the material utilization at a high-active-mass loading of 40 mg cm−2, enabling electrode fabrication with high CS and Cm values. The capacitance increased with the increasing PPy content in the composites. The CFO50 and CFO30 composites showed the highest capacitances, which were 3.60 and 4.52 F cm−2 at 1 mVs−1 or 3.90 and 4.60 at 3 mA cm−2, respectively. The obtained capacitances were higher compared with the capacitances of other composites of magnetic ceramic particles and PPy. The capacitance and magnetic properties of the composites can be varied. The magnetic behavior of the composites is influenced not only by the ferrimagnetic CuFe2O4 but also by the paramagnetic PPy. The composites belong to the category of magnetic supercapacitor materials and can potentially be used for energy storage and applications based on their magnetoelectric and magnetocapacitive properties. The use of ferromagnetic Ni for the fabrication of current collectors generates difficulties in the investigation of such properties. Future research must be focused on the replacement of commercial Ni foam current collectors with non-magnetic conductive current collectors for the investigation of magnetocapacitive and magnetoelectric effects. It is also expected that the development of new composites of conductive redox-active polymers with other magnetic materials and their solid solutions can result in advanced electrodes for energy storage and the capacitive deionization of water.

Author Contributions

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

Funding

This research was funded by the Natural Sciences and Engineering Research Council of Canada (grant RGPIN-2024-03968) and the CRC program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available upon request.

Acknowledgments

The electron microscopy studies were performed at the Canadian Centre for Electron Microscopy.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Romero, M.; Faccio, R.; Montenegro, B.; Tumelero, M.A.; Cid, C.C.P.; Pasa, A.A.; Mombrú, A.W. Role of conducting polyaniline interphase on the low field magnetoresistance for LSMO-PANI nanocomposites. J. Magn. Magn. Mater. 2018, 466, 446–451. [Google Scholar] [CrossRef]
  2. Romero, M.; Faccio, R.; Pardo, H.; Tumelero, M.A.; Cid, C.C.P.; Pasa, A.A.; Mombrú, Á.W. Microstructure, interparticle interactions and magnetotransport of manganite-polyaniline nanocomposites. Mater. Chem. Phys. 2016, 171, 178–184. [Google Scholar] [CrossRef]
  3. Pana, O.; Soran, M.; Leostean, C.; Macavei, S.; Gautron, E.; Teodorescu, C.; Gheorghe, N.; Chauvet, O. Interface charge transfer in polypyrrole coated perovskite manganite magnetic nanoparticles. J. Appl. Phys. 2012, 111, 044309. [Google Scholar] [CrossRef]
  4. Vinayak, V.V.; Deshmukh, K.; Murthy, V.; Pasha, S.K. Conducting polymer based nanocomposites for supercapacitor applications: A review of recent advances, challenges and future prospects. J. Energy Storage 2024, 100, 113551. [Google Scholar] [CrossRef]
  5. Snook, G.A.; Kao, P.; Best, A.S. Conducting-polymer-based supercapacitor devices and electrodes. J. Power Sources 2011, 196, 1–12. [Google Scholar] [CrossRef]
  6. Dhandapani, E.; Thangarasu, S.; Ramesh, S.; Ramesh, K.; Vasudevan, R.; Duraisamy, N. Recent development and prospective of carbonaceous material, conducting polymer and their composite electrode materials for supercapacitor—A review. J. Energy Storage 2022, 52, 104937. [Google Scholar] [CrossRef]
  7. Meng, Q.; Cai, K.; Chen, Y.; Chen, L. Research progress on conducting polymer based supercapacitor electrode materials. Nano Energy 2017, 36, 268–285. [Google Scholar] [CrossRef]
  8. Tadesse, M.G.; Ahmmed, A.S.; Lübben, J.F. Review on conductive polymer composites for supercapacitor applications. J. Compos. Sci. 2024, 8, 53. [Google Scholar] [CrossRef]
  9. Mahimai, B.M.; Li, E.; Pang, J.; Zhang, J.; Zhang, J. Interface engineering in conducting polymers-based supercapacitor. J. Energy Storage 2024, 96, 112598. [Google Scholar] [CrossRef]
  10. Wustoni, S.; Ohayon, D.; Hermawan, A.; Nuruddin, A.; Inal, S.; Indartono, Y.S.; Yuliarto, B. Material design and characterization of conducting polymer-based supercapacitors. Polym. Rev. 2024, 64, 192–250. [Google Scholar] [CrossRef]
  11. Gupta, K.; Jana, P.; Meikap, A.; Nath, T. Synthesis of La0.67Sr0.33MnO3 and polyaniline nanocomposite with its electrical and magneto-transport properties. J. Appl. Phys. 2010, 107, 073704. [Google Scholar] [CrossRef]
  12. Turcu, R.; Pana, O.; Nan, A.; Craciunescu, I.; Chauvet, O.; Payen, C. Polypyrrole coated magnetite nanoparticles from water based nanofluids. J. Phys. D Appl. Phys. 2008, 41, 245002. [Google Scholar] [CrossRef]
  13. Amarnath, C.A.; Ghamouss, F.; Schmaltz, B.; Autret-Lambert, C.; Roger, S.; Gervais, F.; Tran-Van, F. Polypyrrole/lanthanum strontium manganite oxide nanocomposites: Elaboration and characterization. Synth. Met. 2013, 167, 18–24. [Google Scholar] [CrossRef]
  14. Ren, X.; Wang, J.; Yin, H.; Tang, Y.; Fan, H.; Yuan, H.; Cui, S.; Huang, L. Hierarchical CoFe2O4@ PPy hollow nanocubes with enhanced microwave absorption. Appl. Surf. Sci. 2022, 575, 151752. [Google Scholar] [CrossRef]
  15. Kateb, M.; Safarian, S.; Kolahdouz, M.; Fathipour, M.; Ahamdi, V. ZnO–PEDOT core–shell nanowires: An ultrafast, high contrast and transparent electrochromic display. Sol. Energy Mater. Sol. Cells 2016, 145, 200–205. [Google Scholar] [CrossRef]
  16. Helli, M.; Sadrnezhaad, S.; Hosseini-Hosseinabad, S.; Vahdatkhah, P. Synthesis and characterization of CuO micro-flowers/PPy nanowires nanocomposites as high-capacity anode material for lithium-ion batteries. J. Appl. Electrochem. 2023, 54, 1–11. [Google Scholar] [CrossRef]
  17. Yin, Z.; Ding, Y.; Zheng, Q.; Guan, L. CuO/polypyrrole core–shell nanocomposites as anode materials for lithium-ion batteries. Electrochem. Commun. 2012, 20, 40–43. [Google Scholar] [CrossRef]
  18. Zhou, Y.; Jin, X.; Ni, J.; Zhang, S.; Yang, J.; Liu, P.; Wang, Z.; Lei, J. Evaporation induced uniform polypyrrole coating on CuO arrays for free-standing high lithium storage anode. J. Solid State Electrochem. 2019, 23, 1829–1836. [Google Scholar] [CrossRef]
  19. Yin, Z.; Fan, W.; Ding, Y.; Li, J.; Guan, L.; Zheng, Q. Shell structure control of PPy-modified CuO composite nanoleaves for lithium batteries with improved cyclic performance. ACS Sustain. Chem. Eng. 2015, 3, 507–517. [Google Scholar] [CrossRef]
  20. Qian, T.; Zhou, J.; Xu, N.; Yang, T.; Shen, X.; Liu, X.; Wu, S.; Yan, C. On-chip supercapacitors with ultrahigh volumetric performance based on electrochemically co-deposited CuO/polypyrrole nanosheet arrays. Nanotechnology 2015, 26, 425402. [Google Scholar] [CrossRef]
  21. Ates, M.; Serin, M.A.; Ekmen, I.; Ertas, Y.N. Supercapacitor behaviors of polyaniline/CuO, polypyrrole/CuO and PEDOT/CuO nanocomposites. Polym. Bull. 2015, 72, 2573–2589. [Google Scholar] [CrossRef]
  22. Xu, J.; Wang, D.; Yuan, Y.; Wei, W.; Gu, S.; Liu, R.; Wang, X.; Liu, L.; Xu, W. Polypyrrole-coated cotton fabrics for flexible supercapacitor electrodes prepared using CuO nanoparticles as template. Cellulose 2015, 22, 1355–1363. [Google Scholar] [CrossRef]
  23. Awad, M.; Nawwar, M.; Zhitomirsky, I. Synergy of Charge Storage Properties of CuO and Polypyrrole in Composite CuO-Polypyrrole Electrodes for Asymmetric Supercapacitor Devices. ACS Appl. Energy Mater. 2024, 7, 5572–5581. [Google Scholar] [CrossRef]
  24. MacDonald, M.; Zhitomirsky, I. Pseudocapacitive and Magnetic Properties of SrFe12O19–Polypyrrole Composites. J. Compos. Sci. 2024, 8, 351. [Google Scholar] [CrossRef]
  25. MacDonald, M.; Zhitomirsky, I. Capacitive Properties of Ferrimagnetic NiFe2O4-Conductive Polypyrrole Nanocomposites. J. Compos. Sci. 2024, 8, 51. [Google Scholar] [CrossRef]
  26. Dmitriev, A.I. Numerical model of a local contact of a polymer nanocomposite and its experimental validation. Facta Univ. Ser. Mech. Eng. 2021, 19, 079–089. [Google Scholar] [CrossRef]
  27. Wang, K.; Alaluf, D.; Rodrigues, G.; Preumont, A. Precision shape control of ultra-thin shells with strain actuators. J. Appl. Comput. Mech. 2020. [Google Scholar]
  28. Xia, Q.; Xia, T.; Wu, X. PPy decorated α-Fe2O3 nanosheets as flexible supercapacitor electrodes. Rare Met. 2022, 41, 1195–1201. [Google Scholar] [CrossRef]
  29. Liu, L.; Hu, X.; Zeng, H.-Y.; Yi, M.-Y.; Shen, S.-G.; Xu, S.; Cao, X.; Du, J.-Z. Preparation of NiCoFe-hydroxide/polyaniline composite for enhanced-performance supercapacitors. J. Mater. Sci. Technol. 2019, 35, 1691–1699. [Google Scholar] [CrossRef]
  30. Beknalkar, S.A.; Teli, A.M.; Shin, J.C. Current innovations and future prospects of metal oxide electrospun materials for supercapacitor technology: A review. J. Mater. Sci. Technol. 2023, 166, 208–233. [Google Scholar] [CrossRef]
  31. Zhang, Z.P.; Rong, M.Z.; Zhang, M.Q. Self-healable functional polymers and polymer-based composites. Prog. Polym. Sci. 2023, 144, 101724. [Google Scholar] [CrossRef]
  32. Yuan, R.; Yin, X.; Gu, S.; Chang, J.; Li, H. Flexible double-wall carbon foam/Cu-Co oxides based symmetric supercapacitor with ultrahigh energy density. J. Mater. Sci. Technol. 2023, 160, 109–117. [Google Scholar] [CrossRef]
  33. Birajdar, D.; Devatwal, U.; Jadhav, K. X-ray, IR and bulk magnetic properties of Cu1+xMnxFe2–2xO4 ferrite system. J. Mater. Sci. 2002, 37, 1443–1448. [Google Scholar] [CrossRef]
  34. Marinca, T.F.; Chicinaş, I.; Isnard, O. Structural and magnetic properties of the copper ferrite obtained by reactive milling and heat treatment. Ceram. Int. 2013, 39, 4179–4186. [Google Scholar] [CrossRef]
  35. Lu, J.; Zhou, B.; Zhang, X.; Zhao, X.; Liu, X.; Wu, S.; Yang, D.-P. Oyster shell-derived CuFe2O4-Hap nanocomposite for healthy houses: Bacterial and formaldehyde elimination. Chem. Eng. J. 2023, 477, 147054. [Google Scholar] [CrossRef]
  36. Ma, J.; Guo, X.; Ji, X. Amino-functionalized CuFe2O4 supported Pd nanoparticles as magnetically catalyst for H2 production from methanolysis of ammonia borane and hydrogenation of nitro aromatics. Int. J. Hydrogen Energy 2024, 51, 345–356. [Google Scholar] [CrossRef]
  37. Chen, Q.; Zhang, Y.; Xia, H.; Liu, R.; Wang, H. Fabrication of two novel amino-functionalized and starch-coated CuFe2O4-modified magnetic biochar composites and their application in removing Pb2+ and Cd2+ from wastewater. Int. J. Biol. Macromol. 2024, 258, 128973. [Google Scholar] [CrossRef]
  38. Mostafa, E.M.; Hammam, R.E. Tailored solar collector coatings: Synthesis and characterization of CuFe2O4/PANI nanocomposites. Opt. Mater. 2024, 156, 115879. [Google Scholar] [CrossRef]
  39. Guo, Y.; Chen, Y.; Hu, X.; Yao, Y.; Li, Z. Tween modified CuFe2O4 nanoparticles with enhanced supercapacitor performance. Colloids Surf. A Physicochem. Eng. Asp. 2021, 631, 127676. [Google Scholar] [CrossRef]
  40. Liang, W.; Yang, W.; Sakib, S.; Zhitomirsky, I. Magnetic CuFe2O4 nanoparticles with pseudocapacitive properties for electrical energy storage. Molecules 2022, 27, 5313. [Google Scholar] [CrossRef]
  41. Rushmittha, J.; Radhika, S.; Alrashidi, K.A.; Maheshwaran, G.; Dhinesh, S.; Sambasivam, S. Hybridization of CuFe2O4 by carbon microspheres with improved charge storage characteristics for high energy density solid-state hybrid supercapacitor. Mater. Today Sustain. 2024, 28, 100950. [Google Scholar] [CrossRef]
  42. Nikam, P.N.; Patil, S.S.; Chougale, U.M.; Bajantri, T.H.; Fulari, A.V.; Fulari, V.J. Synthesis and characterization of CuFe2O4 spinel ferrite for supercapacitor application. J. Indian Chem. Soc. 2024, 101, 101277. [Google Scholar] [CrossRef]
  43. Mbebou, M.; Polat, S.; Zengin, H. Sustainable cauliflower-patterned CuFe2O4 electrode production from chalcopyrite for supercapacitor applications. Nanomaterials 2023, 13, 1105. [Google Scholar] [CrossRef]
  44. Sikkema, R.; Zhitomirsky, I. Magnetic supercapacitors: Charge storage mechanisms, magnetocapacitance, and magnetoelectric phenomena. Appl. Phys. Rev. 2023, 10, 021307. [Google Scholar] [CrossRef]
  45. Ata, M.; Liu, Y.; Zhitomirsky, I. A review of new methods of surface chemical modification, dispersion and electrophoretic deposition of metal oxide particles. Rsc Adv. 2014, 4, 22716–22732. [Google Scholar] [CrossRef]
  46. Conway, B.E. Electrochemical Supercapacitors: Scientific Fundamentals and Technological Applications; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  47. Taberna, P.; Simon, P.; Fauvarque, J.-F. Electrochemical characteristics and impedance spectroscopy studies of carbon-carbon supercapacitors. J. Electrochem. Soc. 2003, 150, A292. [Google Scholar] [CrossRef]
  48. Al Kiey, S.A.; Ramadan, R.; El-Masry, M.M. Synthesis and characterization of mixed ternary transition metal ferrite nanoparticles comprising cobalt, copper and binary cobalt–copper for high-performance supercapacitor applications. Appl. Phys. A 2022, 128, 473. [Google Scholar] [CrossRef]
  49. Khairy, M.; Bayoumy, W.; Selima, S.; Mousa, M. Studies on characterization, magnetic and electrochemical properties of nano-size pure and mixed ternary transition metal ferrites prepared by the auto-combustion method. J. Mater. Res. 2020, 35, 2652–2663. [Google Scholar] [CrossRef]
  50. Berkowitz, A.; Schuele, W.; Flanders, P. Influence of crystallite size on the magnetic properties of acicular γ-Fe2O3 particles. J. Appl. Phys. 1968, 39, 1261–1263. [Google Scholar] [CrossRef]
  51. Zan, G.; Li, S.; Chen, P.; Dong, K.; Wu, Q.; Wu, T. Mesoporous Cubic Nanocages Assembled by Coupled Monolayers with 100% Theoretical Capacity and Robust Cycling. ACS Cent. Sci. 2024, 10, 1283–1294. [Google Scholar] [CrossRef]
  52. Li, J.; Dong, Z.; Chen, R.; Wu, Q.; Zan, G. Advanced nickel-based composite materials for supercapacitor electrodes. Ionics 2024, 30, 1833–1855. [Google Scholar] [CrossRef]
  53. Wang, Y.; Liu, Y.; Zhitomirsky, I. Surface modification of MnO2 and carbon nanotubes using organic dyes for nanotechnology of electrochemical supercapacitors. J. Mater. Chem. A 2013, 1, 12519–12526. [Google Scholar] [CrossRef]
  54. Conway, B.; Pell, W. Power limitations of supercapacitor operation associated with resistance and capacitance distribution in porous electrode devices. J. Power Sources 2002, 105, 169–181. [Google Scholar] [CrossRef]
  55. Conway, B.E.; Birss, V.; Wojtowicz, J. The role and utilization of pseudocapacitance for energy storage by supercapacitors. J. Power Sources 1997, 66, 1–14. [Google Scholar] [CrossRef]
  56. Pell, W.; Conway, B.; Marincic, N. Analysis of non-uniform charge/discharge and rate effects in porous carbon capacitors containing sub-optimal electrolyte concentrations. J. Electroanal. Chem. 2000, 491, 9–21. [Google Scholar] [CrossRef]
  57. Pell, W.G.; Conway, B.E. Voltammetry at a de Levie brush electrode as a model for electrochemical supercapacitor behaviour. J. Electroanal. Chem. 2001, 500, 121–133. [Google Scholar] [CrossRef]
  58. Wu, C.; Wang, J.; Bai, Y.; Li, X. Significant effect of cations on polypyrrole cycle stability. Solid State Ion. 2020, 346, 115216. [Google Scholar] [CrossRef]
  59. Aman, S.; Gouadria, S.; Ahmad, N.; Khan, S.A.; Manzoor, S.; Farid, H.M.T. The structural, dielectric and magnetic study of cobalt based spinel ferrites and polypyrrole composites. J. Mater. Sci. Mater. Electron. 2022, 33, 19534–19543. [Google Scholar] [CrossRef]
  60. Bartl, A.; Dunsch, L.; Schmeiβer, D.; Göpel, W.; Naarmann, H. Influence of oxygen on the paramagnetic properties of polypyrrole layers. Synth. Met. 1995, 69, 389–390. [Google Scholar] [CrossRef]
  61. Ghasemi, A.K.; Ghorbani, M.; Lashkenari, M.S.; Nasiri, N. Facile synthesize of PANI/GO/CuFe2O4 nanocomposite material with synergistic effect for superb performance supercapacitor. Electrochim. Acta 2023, 439, 141685. [Google Scholar] [CrossRef]
  62. Ishaq, S.; Moussa, M.; Kanwal, F.; Ayub, R.; Van, T.N.; Azhar, U.; Losic, D. One step strategy for reduced graphene oxide/cobalt-iron oxide/polypyrrole nanocomposite preparation for high performance supercapacitor electrodes. Electrochim. Acta 2022, 427, 140883. [Google Scholar] [CrossRef]
  63. Senthilkumar, B.; Sankar, K.V.; Sanjeeviraja, C.; Selvan, R.K. Synthesis and physico-chemical property evaluation of PANI–NiFe2O4 nanocomposite as electrodes for supercapacitors. J. Alloys Compd. 2013, 553, 350–357. [Google Scholar] [CrossRef]
Figure 1. (A) X-ray X-ray diffraction pattern of CuFe2O4, prepared by hydrothermal synthesis. Miller indices of planes correspond to JCPDS file 00-034-0425. (B) Magnetization (M) plotted against magnetic field (H) for CuFe2O4 (inset shows low-field range).
Figure 1. (A) X-ray X-ray diffraction pattern of CuFe2O4, prepared by hydrothermal synthesis. Miller indices of planes correspond to JCPDS file 00-034-0425. (B) Magnetization (M) plotted against magnetic field (H) for CuFe2O4 (inset shows low-field range).
Materials 17 05249 g001
Figure 2. (AC) TEM images of CuFe2O4 particles, prepared through hydrothermal synthesis. Spacings of 2.5 Å and 2.97 Å in the high-resolution image (C) correspond to d-spacings of crystallographic planes (211) and (112), respectively.
Figure 2. (AC) TEM images of CuFe2O4 particles, prepared through hydrothermal synthesis. Spacings of 2.5 Å and 2.97 Å in the high-resolution image (C) correspond to d-spacings of crystallographic planes (211) and (112), respectively.
Materials 17 05249 g002
Figure 3. (A) CV data at 20 mV s−1, (B) capacitances at different scan rates, (C) GCD data at 3 mA cm−2, and (D) capacitances at different current densities for different composites.
Figure 3. (A) CV data at 20 mV s−1, (B) capacitances at different scan rates, (C) GCD data at 3 mA cm−2, and (D) capacitances at different current densities for different composites.
Materials 17 05249 g003
Figure 4. (A) Nyquist plots, (B) equivalent circuit model of EIS data, and (C,D) components of AC capacitance versus frequency.
Figure 4. (A) Nyquist plots, (B) equivalent circuit model of EIS data, and (C,D) components of AC capacitance versus frequency.
Materials 17 05249 g004
Figure 5. (a) RCT, (b) Cm, and (c) CS versus PPy content in the CuFe2O4 − PPy mixture.
Figure 5. (a) RCT, (b) Cm, and (c) CS versus PPy content in the CuFe2O4 − PPy mixture.
Materials 17 05249 g005
Figure 6. Cyclic stability data at a sweep rate of 50 mV s−1 for (A) CFO50 and (B) CFO30.
Figure 6. Cyclic stability data at a sweep rate of 50 mV s−1 for (A) CFO50 and (B) CFO30.
Materials 17 05249 g006
Figure 7. SEM images of surface microstructures of (A) CFO50 and (B) CFO30.
Figure 7. SEM images of surface microstructures of (A) CFO50 and (B) CFO30.
Materials 17 05249 g007
Figure 8. M–H curves for (A) CFO50 and (B) CFO30 (insets show low-field range).
Figure 8. M–H curves for (A) CFO50 and (B) CFO30 (insets show low-field range).
Materials 17 05249 g008
Table 1. Capacitances of different composite electrodes.
Table 1. Capacitances of different composite electrodes.
ElectrodeCapacitance from CV DataCapacitance from GCD Data
1 mV s−1100 mV s−13 mA cm−240 mA cm−2
CS
F cm−2
Cm
F g−1
CS
F cm−2
Cm
F g−1
CS
F cm−2
Cm
F g−1
CS
F cm−2
Cm
F g−1
CFO 1001.2310.8521.71.1729.90.6316.14
CFO 901.539.21.129.351.4638.51.2232
CFO 702.5363.82.0752.22.666.72.2255.84
CFO 503.691.11.8747.13.91132.768
CFO 304.52117.372.2458.44.6120.13.5692.68
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

Awad, M.; Zhitomirsky, I. Magnetic CuFe2O4 Spinel–Polypyrrole Pseudocapacitive Composites for Energy Storage. Materials 2024, 17, 5249. https://doi.org/10.3390/ma17215249

AMA Style

Awad M, Zhitomirsky I. Magnetic CuFe2O4 Spinel–Polypyrrole Pseudocapacitive Composites for Energy Storage. Materials. 2024; 17(21):5249. https://doi.org/10.3390/ma17215249

Chicago/Turabian Style

Awad, Mahmoud, and Igor Zhitomirsky. 2024. "Magnetic CuFe2O4 Spinel–Polypyrrole Pseudocapacitive Composites for Energy Storage" Materials 17, no. 21: 5249. https://doi.org/10.3390/ma17215249

APA Style

Awad, M., & Zhitomirsky, I. (2024). Magnetic CuFe2O4 Spinel–Polypyrrole Pseudocapacitive Composites for Energy Storage. Materials, 17(21), 5249. https://doi.org/10.3390/ma17215249

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