Next Article in Journal
A Yield Stress and Work Hardening Model of Al-Mg-Si Alloy Considering the Strengthening Effect of β” and β’ Precipitates
Next Article in Special Issue
Point-Contact Spectroscopy in Bulk Samples of Electron-Doped Cuprate Superconductors
Previous Article in Journal
Modification of Epoxides with Metallic Fillers—Mechanical Properties after Ageing in Aqueous Environments
Previous Article in Special Issue
Axion Field Influence on Josephson Junction Quasipotential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rough and Porous Micropebbles of CeCu2Si2 for Energy Storage Applications

1
Department of Physics “E.R. Caianiello”, University of Salerno, Via Giovanni Paolo II, 132, 84084 Fisciano, Italy
2
NANO_MATES Research Centre, University of Salerno, Via Giovanni Paolo II, 132, 84084 Fisciano, Italy
3
CNR-SPIN, c/o University of Salerno, Via Giovanni Paolo II, 132, 84084 Fisciano, Italy
*
Author to whom correspondence should be addressed.
Materials 2023, 16(22), 7182; https://doi.org/10.3390/ma16227182
Submission received: 17 September 2023 / Revised: 13 November 2023 / Accepted: 14 November 2023 / Published: 16 November 2023
(This article belongs to the Special Issue Physics and Application of Superconductivity)

Abstract

:
Supercapacitors have attracted considerable attention due to their advantages, including being lightweight and having rapid charge–discharge, a good rate capability, and high cyclic stability. Electrodes are one of the most important factors influencing the performance of supercapacitors. Herein, a three-dimensional network of rough and porous micropebbles of CeCu2Si2 has been prepared using a one-step procedure and tested for the first time as a supercapacitor electrode. The synthesized material was extensively characterized in a three-electrode configuration using different electrochemical techniques, such as cyclic voltammetry (CV), galvanostatic charge and discharge (GCD) tests, and electrochemical impedance spectroscopy (EIS). CeCu2Si2 shows rather high mass-capacitance values: 278 F/g at 1 A/g and 295 F/g at 10 mV/s. Moreover, the material exhibits remarkable long-term stability: 98% of the initial capacitance was retained after 20,000 cycles at 10 A/g and the Coulombic efficiency remains equal to 100% at the end of the cycles.

1. Introduction

In the present day, the escalating global demand for electrical energy has sparked a significant demand for energy storage solutions capable of delivering top-notch performance. Indeed, this technology plays a crucial role in numerous stand-alone applications. On the other hand, within the context of distributed renewable energy systems, energy storage holds a pivotal position, not only serving stand-alone systems but also facilitating the integration of these renewable sources into the electrical grid. In this context, energy storage takes on a central role in the effort to combine forthcoming sustainable energy supply with the established norms of technical services and products. Numerous energy storage technologies have been developed, including electrochemical energy storage, fluid-based storage, mechanical systems, and electromagnetic solutions. These diverse energy storage methods coexist due to their unique characteristics, which makes them suitable for various specific applications [1].
In the field of electrochemical energy storage (EES), supercapacitors have drawn increasing attention in addressing the emerging energy demand, primarily because of their plethora of advantages over conventional capacitors and batteries. These advantages encompass superior power densities, exceptional long-cycle lifespans, fast charge–discharge characteristics, adaptability to a wide range of temperatures, as well as improved eco-friendliness [2,3,4].
Supercapacitors, as an energy storage solution, can be conveniently classified into three primary categories: electrical double-layer capacitors (EDLCs), redox capacitors, and pseudocapacitors. Among these, EDLCs operate on a distinct energy storage principle based on the formation of a dual layer of ions in close proximity to the electrode surfaces, a phenomenon commonly referred to as the Helmholtz electrical double layer. Owing to this specific mechanism, electrical double-layer capacitors can deliver exceptional power densities, relying on the rapid kinetics involved in physical ion adsorption and desorption. However, it is important to note that EDLCs typically exhibit lower capacitances and energy density when compared to the other categories. In contrast, redox capacitors are based on an alternative energy storage mechanism that revolves around redox reactions occurring at the interfaces between the electrodes and the electrolyte. These redox reactions can also occur within the bulk of the electrode material, endowing redox capacitors with higher capacitance compared to their EDLC counterparts. Nevertheless, the charge–discharge process in redox capacitors, which results from redox reactions, is notably slower than the physical mechanism of EDLCs. Consequently, redox capacitors are associated with reduced power densities along with higher capacitances. Ultimately, pseudocapacitors emerge as a category that bridges the gap between the previously discussed groups. The charge storage mechanism in pseudocapacitors is characterized by rapid redox reactions at the electrode–electrolyte interface so that the electrochemical signature of a pseudocapacitor electrode is close to the one of EDLSCs. This unique mechanism, often referred to as “pseudocapacitance”, confers upon pseudocapacitors the dual benefit of ensuring elevated power and energy densities [5,6].
On the other hand, lanthanide (Ln) elements have been extensively investigated across various application domains, including the display industry, magnets, bio-industry, pyro-processing technology in nuclear power plants, and energy/environment fields [7,8,9,10,11,12,13,14,15,16]. Indeed, these elements exhibit unique physicochemical properties due to their unpaired 4f electronic configurations. The closely matched energy levels of the 4f and 5d orbitals enable efficient electron transfer between these orbitals, which in turn promotes redox processes. Additionally, Ln elements can effectively store a higher number of charges due to their multiple oxidation states and the capacity to undergo reversible redox reactions. These features can contribute to improving charge–discharge rates as well as capacitance, making these materials suitable as supercapacitor electrodes [17,18]. Among alloy systems including Ln elements, CeCu2Si2 has drawn substantial attention due to its distinctive superconducting properties, with research extending back to the previous century. In 1979, indeed, a groundbreaking discovery arose within the domain of condensed matter physics since superconductivity was observed for the first time in the magnetic material CeCu2Si2. This marked a significant milestone as it introduced the scientific community to an entirely new category of superconductors, often referred to as unconventional non-phonon-mediated superconductors. Notably, in these unconventional superconductors, there is a prevailing suspicion that the mechanism responsible for the formation of Cooper pairs, which is responsible for superconductivity, is intrinsically connected to magnetism. In the context of intermetallic compounds, the electronic degrees of freedom, crucial for facilitating superconductivity, exhibit a direct and intricate relationship with the magnetic moments originating from the partially filled f shell of the atoms involved [19].
On the other hand, as a result of the collaborative contribution of various constituent elements, which can provide interesting features and fast redox reactions due to a synergistic effect of electronic conduction and enhanced oxidation states, CeCu2Si2 material emerges as an intriguing and innovative prospect for exploration as a supercapacitor electrode offering promising possibilities in this regard.
In particular, silicon (Si) has recently drawn increasing attention in the energy storage field due to its low working potential and high theoretical capacity (around 4200 mA h g−1). Moreover, silicon is the second most abundant element on Earth, which means there is a stable and sustainable supply. This, combined with its low cost and environmentally safe properties, places silicon in a favorable position as a viable choice for supercapacitor electrode materials [20,21,22]. To enhance the electrical conductivity and the specific capacitance of the electrode, the combination of several elements with Si to form a ternary material can be an effective method. On the other hand, copper (Cu) is one of the most suitable and effective candidates because of its low cost, good electrical conductivity, and environmentally friendly nature [23,24]. Additionally, density functional theory (DFT) calculations proved that the introduction of Cu atoms into another element’s lattice could enhance the electron transport rate, therefore increasing the overall electrochemical activity, as a consequence of the introduction of new energy levels near the Fermi level, originated from the 3D orbitals of Cu [25,26]. Cerium (Ce), another element considered for integration with silicon, possesses high electrical conductivity [27,28] and a rich energy-level structure [29]. These qualities enhance the inherent activity of the electrode material and improve its electrochemical performance. Indeed, Ce exhibits enhanced pseudocapacitive features owing to the fast oxidation and reduction of the Ce3+/Ce4+ ions [30]. Overall, combining Cu and Ce metals in an electrode enables one to obtain an increased number of active sites and several metal ion valence states, resulting in an additional synergistic effect of electronic conduction [31].
Furthermore, apart from composition, morphology stands as another critical element that demands careful consideration when aiming to design an efficient supercapacitor electrode. Indeed, porosity [32] and rough surfaces [33] increase wettability and the available interface area between electrodes and electrolytes.
In the following study, a rough-surface and porous ternary CeCu2Si2-based material, prepared via a one-step synthetic procedure, has been tested for the first time as a supercapacitor electrode. The synthesized material was first characterized from a morphological point of view. Eventually, its electrochemical performance was investigated via cyclic voltammetry, galvanostatic charge–discharge tests, and electrochemical impedance spectroscopy carried out in a three-electrode-setup-based electrolytic cell filled with a H2SO4 aqueous solution.

2. Materials and Methods

2.1. CeCu2Si2 Preparation

CeCu2Si2 was synthesized starting from elements with high purities (>99.9%, Sigma-Aldrich, St. Louis, MO, USA) in a molten flux of the different elements. The reaction mixture was obtained by loading the Ce, Cu, and Si elements in the optimized ratio of 1:13:11.44 into a 5 mL alumina crucible, which was then covered by a second inverted crucible filled with silica wool. Such a system was sealed under vacuum in a quartz ampoule and heated to 800 °C at 50 °C/h. Afterward, it was maintained at 800 °C for 7 h, further heated to 1200 °C at 50 °C/h, kept at 1200 °C for 12 h, and eventually cooled down at 2 °C/h to 900 °C. At this temperature, the ampoule was quickly removed from the furnace and inverted; this way, the silica wool acted as a sieve for separating most of the liquid flux from the solid material. Afterward, the ampoule was inserted into a centrifuge with the purpose of further separating the remaining liquid flux and recovering the powder.

2.2. Characterization

First of all, the samples were characterized by size, morphology, composition, porosity, and structure. For these purposes, the electron microscope, which is one of the most powerful analytical tools available for characterizing microscopic particulate material, allowing the examination of particle morphology and surface details, was used. In particular, Scanning Electron Microscopy (SEM) used in tandem with Energy-Dispersive X-ray Microanalysis (EDX) is a combination of instrumentation that is capable of providing detailed information on the size and composition at individual particle levels, with the strength of gathering data in a time-efficient manner. SEM images were obtained with a TESCAN-VEGA LMH (230 V) microscope (TESCAN, Brno, Czech Republic), equipped with an EDX probe. The X-ray diffraction analysis (XRD) of the sample was obtained through a Bruker D8 X-ray diffractometer (Bruker Corporation, Billerica, MA, USA), with monochromatic CuKa (λ = 0.1542 nm) radiation, operating at 40 kV and 30 mA. The diffraction pattern was collected at 25 °C and over an angular range of 10° to 40°. The BET method was used to measure the physisorption of nitrogen, passing between particles and into pores, cracks, and surface roughness, to derive the value of the “surface area”. The N2 adsorption–desorption isotherm was obtained at −196 °C. In particular, the surface area was evaluated using Brunauer–Emmett–Teller (BET) with a Sorptometer 1042 instrument (Costech International, Milan, Italy). Before the analysis, the sample was pre-treated at 150 °C for 60 min under a He flow. The Barrett–Joyner–Halenda (BJH) method was used to determine the pore-size distribution of the material.
Cyclic voltammetry, galvanostatic charge and discharge tests, and electrochemical impedance spectroscopy experiments were carried out using an Autolab PGSTAT302N potentiostat (Metrohm, Herisau, Switzerland) in a three-electrode-setup-based electrolytic cell filled with a 1 M H2SO4 solution. Saturated calomel, graphite, and a loadable glassy carbon ring (3 mm diameter) were adopted, respectively, as reference, counter, and working electrodes [34].
In preparation for the electrochemical tests, 8 mg of the material was mixed into 160 μL of a 5 wt% Nafion solution, 900 μL of 2-propanol, and 100 μL of water to create a stable and conductive homogenous ink. This ink was then partially drop-casted onto the working electrode and dried under an ambient environment to attain a uniform catalyst film with a mass loading of about 2 mg/cm2.
The EIS spectra were taken at open circuit voltage (OCV) using a 5 mV amplitude at frequencies ranging from 200 kHz to 10 MHz.
The mass capacitance (Csp, F/g) of the electrode has been calculated from galvanostatic charge–discharge curves according to the following equation:
C s p = i · t m · V
where i is the GCD current, ∆t is the discharging time, ∆V is the potential window, and m is the loading mass of the electrode material.
Moreover, capacitance retention at the nth cycle has been evaluated according to the following equation:
C a p a c i t a n c e   r e t e n t i o n   ( n   c y c l e , % ) = C s p   ( n   c y c l e ) C s p · 100
where Csp (n cycle) is the mass capacitance calculated at the nth cycle (with n = 2, 3… 20,000), while Csp is the mass capacitance calculated at the 1st cycle.
Eventually, Coulombic efficiency has been evaluated according to the following equation:
C o u l o m b i c   e f f i c i e n c y   ( n   c y c l e ,   % ) = C s p , d i s c h a r g e   ( n   c y c l e ) C s p ,     c h a r g e   ( n   c y c l e ) · 100
where Csp, discharge (n cycle) is defined as the measured discharge mass capacity at the nth cycle, whereas Csp, charge (n cycle) is the measured charge mass capacity at the same nth cycle, calculated from (1) considering ∆t as the charging time [35,36,37].
Energy density (E, Wh/kg) and power density (P, W/kg) have been evaluated according to the following formulas:
E = 1 2 C s p ·   V 2 1000 3600  
P = E t  
where Csp is the mass capacitance calculated in (1), ∆t is the discharging time, and ∆V is the potential window from the galvanostatic charge–discharge curves.

3. Results and Discussion

3.1. Morphological Characterization

The morphological features of the CeCu2Si2 were investigated through SEM analysis. The results of SEM analysis on hard surfaces are reported in Figure 1. As can be seen from Figure 1, the material is in the form of a three-dimensional network of micropebbles, strictly connected to each other, showing a rough and porous surface. Indeed, the BET surface area for CeCu2Si2, obtained via the nitrogen adsorption–desorption isotherm (Figure 2) is equal to 12 m2/g, and the total pore volume and the micropore volume are 0.14 cm3/g and 0.001 cm3/g, respectively. The multi-modal pore size distribution (BJH pore distribution) is centered at 26.5 nm (inter-particle surface porosity) and 16 nm (intraparticle surface porosity).
Moreover, EDX maps were collected to verify the homogeneity and chemical composition of the prepared sample. As displayed in Figure 3, all constituent elements (Ce, Si, and Cu) are distributed homogeneously within the sample. The real chemical formula CeCu2Si2 was also confirmed.
The X-ray diffraction pattern in Figure 4, showing the typical peaks of the tetragonal crystal structure of CeCu2Si2, confirms the nature of the sample [38]. The intensities of the diffraction lines are in agreement with what would be expected for a randomly oriented polycrystalline material. In particular, the peaks indexes can be referred to as the ones of the ThCr2Si2 structure type, with the Ce, Cu, and Si atoms located, respectively, on the 2a (0,0,0), 4d (0,0.5,0.25), and 4e (0,0,z) Wyckoff sites of the space group I4/mmm [39,40].

3.2. Electrochemical Characterization

The capacitive behavior of the CeCu2Si2 electrode was first investigated in a three-electrode-setup-based electrolytic cell filled with a 1 M H2SO4 aqueous electrolyte solution via cyclic voltammetry. The electrochemical curves were recorded in the potential range −0.4÷0.7 V at several scan rates (5, 10, 20, 50, 100 and 200 mV/s). As can be seen in Figure 5a, the material shows an approximately quasi rectangular current–voltage response without detectable redox peaks, suggesting pseudocapacitive behavior. The first definition of pseudocapacitance was proposed by B.E. Conway in [41]. In particular, pseudocapacitance was suggested to arise at the electrode surfaces where materials undergo fast and reversible redox reactions, which is reflected in an electrochemical signature of (quasi)-rectangular CV curves similar to EDLC. The absence of visible peaks here can be ascribed to fast surface redox reactions, suggesting that the charge transport is taking place at high rates. The lack, even at lower scan rates, may be representative of the equitable contribution to the capacitance of both the electrostatic and the redox storage mechanisms [42,43,44,45]. On the other hand, regarding redox pseudocapacitance, a contribution can be expected to be made by synergy between the Ce and Cu elements, consisting of a higher tendency to reduction and oxidation, respectively, of the Ce4+/Ce3+ and Cu+/Cu2+ redox couples as well as an improved charge transfer between them [31], as will be further explained in Section 3.3. Moreover, the CV curves of the bare glassy carbon electrode at the same current densities (see Figure 5b), exhibiting currents of only a few µA, enabled us to rule out a significant contribution of the bare electrode to the material capacitance.
GCD tests were also performed in the current density range of 1 ÷ 10 A/g, and the results are reported in Figure 6; they show quasi-linear charge and discharge curves. There is no obvious IR drop even at the high current density of 10 A/g, probably due to the high electrode conductivity, enhanced by the intimate connection between the elements in the material.
From the galvanostatic discharge curve at 1 A/g, a mass-capacitance of 278 F/g was calculated according to (1). The mass capacitance, as expected, decreases at increasing current density, due to the reduced utilization of active material caused by the shorter time taken for the electrolyte ions to diffuse into the electrode channels. However, it still remains high (~180 F/g) at the high discharge current density of 10 A/g (see Figure 7a), with a retention of about 63% within this current range. Moreover, the sample exhibits a capacitance retention of about 89%, (see Figure 7b), with a mass capacitance of 295 F/g at a scan rate of 10 mV/s decreasing to only 262 F/g at 100 mV/s. Even in this case, the mass capacitance value steadily decreases with the increase in the scan rate since the electrolyte ions have less and less time for diffusion into the material as the scan rate increases. At a higher scan rate, a faster redox reaction happens, which reduces the electrolyte ion diffusion into the active material and consequently lowers the capacitance [46]. To further investigate the electrochemical processes, the relationship between current density and scan rate was investigated. Typically, the above-mentioned relationship is commonly expressed in the following manner:
i = a v b
where i is the current density, v is the scan rate, and a and b are appropriate values. In this specific case, the value of b is 0.72, suggesting that the electrochemical process is not completely unconditioned by diffusion paths [47].
On the other hand, even when the material power density, calculated according to (5), reaches 47 kW·kg−1, its energy density, obtained according to (4), maintains a value of 72 Wh·kg−1.
Furthermore, the sample shows high durability in terms of capacitance retention, evaluated according to (2): in 1 M H2SO4 at a high current density of 10 A/g, the sample retains 98% of the initial capacitance over 20,000 cycles (Figure 7c). The Coulombic efficiency of the sample, which measures the electron transfer efficiency during the charge–discharge process, was also calculated according to (3), being nearly 100%, which can be attributed to the high charge–discharge rates of the material due to Ce and Cu, and it maintains its value, which is almost unchanged until the 20,000th cycle (Figure 7d). Eventually, electrochemical impedance spectroscopy was carried out to determine the resistivity, as well as the charge-transfer and ion diffusion behavior of the CeCu2Si2 material. Figure 7e illustrates the Nyquist plot exhibiting a semi-circle in the high-frequency region and a quasi-vertical line feature at a lower frequency related to the pseudo-capacitive contribution. In the middle-frequency region, an initial inclined portion of the curve (about 45°) can be attributed to the Warburg impedance, related to the resistance to the ion diffusion/transport from the electrolyte to the electrode [48]. In the high-frequency region, the intercept on the Z′ axis indicates the equivalent series resistance (ESR), i.e., the combination of the internal resistance of the electrode, the ohmic resistance of the electrolyte, as well as the contact resistance between the current collector and electrode. For the CeCu2Si2 material, an ESR of 0.42 Ω has been calculated, which indicates a rather small internal resistance. The nature of the material, its wettability ensured by the rough surface structure, as well as the path for ion diffusion due to mesopores, might be responsible for the interesting results, especially the non-trivial stability observed for the reported non-nanometric size material. These results encourage the use of this non-precious and cheaper material, non-conventionally considered and unexplored for supercapacitor application.

3.3. Comparison with Literature and Discussion

In Table 1, the capacitive performance of CeCu2Si2 is compared with the ones of some of the best-performing Cu, Ce, and/or Si-containing electrode materials for supercapacitor applications in the literature.
From the comparison, it is evident that the obtained material ranks among the best-performing ones, even when compared with finer-size materials and materials produced with less scalable processes. The CeCu2Si2 electrode exhibits a rather high specific capacitance at low current density, and, above all, it retains capacitance for a very high number of cycles. This can be attributed to several factors: the increase in the overall electrical conductivity due to the intimate electronic connection among the components; the rough surface and mesoporosity which guarantee both wettability with the electrolytic solution and ion diffusion path within the particles; very small internal resistance to charge transfer, which further extends the electrode durability.
Regarding Ce and Cu, they can exhibit enhanced pseudocapacitive features, owing to the fast reduction and oxidation, respectively, of the Ce4+/Ce3+ and the Cu+/Cu2+ redox couples. In particular [31], the synergistic effect between these two elements could have an important influence on the reaction mechanism. Indeed, having different standard reduction potentials (E° = 1.44 V for Ce4+/Ce3+; E° = 0.15 V for Cu+/Cu2+), Ce4+ is a stronger oxidant (7), whereas Cu2+ is a stronger reductant (9), with a contextual reduction of Ce3+ to Ce4+ and oxidation of Cu2+ to Cu, enabling fast charge transfer (8).
C e 4 + + e C e 3 +
C e 3 + + C u 2 + C e 4 + + C u +
C u + C u 2 + + e

4. Conclusions

In summary, rough-surface and porous-structure-based micropebbles of CeCu2Si2, a material typically known for its unique superconducting properties, have been successfully prepared with the aim of a simple one-step synthesis process and tested for the first time as a supercapacitor electrode. The X-ray diffraction analysis, which revealed the characteristic peaks associated with the tetragonal crystal structure of CeCu2Si2, confirmed the nature of the sample. Scanning electron microscope images at different magnifications, Brunauer–Emmett–Teller, and Barrett–Joyner–Halenda further elucidated the material morphology, indicating that CeCu2Si2 is in the form of a three-dimensional network of interconnected micropebbles, with a rough and mesoporous surface. On the other hand, the corresponding EDX maps confirmed that all the constituent elements (Ce, Si, and Cu) are distributed homogeneously within the material structure.
Afterward, the synthesized material was broadly characterized by different electrochemical techniques, and the capacitance performance was evaluated. Cyclic voltammetry and galvanostatic charge–discharge tests provided nearly rectangular and quasi-linear curves without plateaus, respectively, suggesting a pseudocapacitor behavior with an equitable contribution to the capacitance of both the redox and the electrostatic storage mechanisms. The good capacitive performance of CeCu2Si2 was confirmed by its rather high mass capacitance values (278 F/g at 1 A/g and 295 F/g at 10 mV/s), which remain high as both current density and scan rate increase. Furthermore, long-term cycling tests were performed, proving that the material retains its capacitance for a very high number of cycles. This behavior can be attributed to several factors: the increase in the overall electrical conductivity due to the intimate electronic connection among the components; the rough surface and mesoporosity which guarantee both wettability with the electrolytic solution and ion diffusion path within the particles; very small internal resistance to charge transfer which further extends the electrode durability. Moreover, the reported study demonstrates the potential to explore well-established materials in the domain of condensed matter physics, from both a chemical and structural perspective, and also in new emerging applications, such as in the EES field, where they can be found as really promising materials.

Author Contributions

Conceptualization, M.S., D.S. and E.P.; methodology, C.C. (Claudia Cirillo); investigation, C.L.; resources, M.S.; data curation, R.A. and M.I.; writing—original draft preparation, D.S.; writing—review and editing, D.S. and M.S.; supervision, M.S., C.A., A.N. and C.C. (Carla Cirillo); project administration, M.S., A.N. and C.C. (Carla Cirillo); funding acquisition, M.S. and C.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fernão Pires, V.; Romero-Cadaval, E.; Vinnikov, D.; Roasto, I.; Martins, J.F. Power converter interfaces for electrochemical energy storage systems—A review. Energy Convers. Manag. 2014, 86, 453–475. [Google Scholar] [CrossRef]
  2. Guo, M.; Liu, Y.; Cheng, F.; Cheng, C.; Mia, Y.; Gao, F.; Yu, J. Inactive Al3+-doped La(CoCrFeMnNiAlx)1/(5+x)O3 high-entropy perovskite oxides as high performance supercapacitor electrodes. J. Adv. Ceram. 2022, 11, 742–753. [Google Scholar] [CrossRef]
  3. Han, C.; Xu, X.; Mu, H.; Tian, Q.; Li, Q.; Liu, Y.; Zhang, X.; Zhao, Z.; Su, X. Construction of hierarchical sea urchin-like manganese substituted nickel cobaltite@tricobalt tetraoxide core-shell microspheres on nickel foam as binder-free electrodes for high performance supercapacitors. J. Colloid. Interface Sci. 2021, 596, 89–99. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, W.; Wang, J.; Gao, S.; Zhang, H.; Wang, H.; Li, Q. Photo-assisted charging of carbon fiber paper-supported CeO2/MnO2 heterojunction and its long-lasting capacitance enhancement in dark. J. Adv. Ceram. 2022, 11, 1735–1750. [Google Scholar] [CrossRef]
  5. Wang, G.; Zhang, L.; Zhangb, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef]
  6. Brousse, T.; Bélanger, D.; Long, J.W. To Be or Not To Be Pseudocapacitive? J. Electrochem. Soc. 2015, 162, A5185. [Google Scholar] [CrossRef]
  7. Ricci, P.C. Assessment of crystalline materials for solid state lighting applications: Beyond the rare earth elements. Crystals 2020, 10, 559. [Google Scholar] [CrossRef]
  8. Monteiro, J.H.S.K. Recent Advances in Luminescence Imaging of Biological Systems Using Lanthanide (III) Luminescent Complexes. Molecules 2020, 25, 2089. [Google Scholar] [CrossRef]
  9. Gontcharenko, V.E.; Kiskin, M.A.; Dolzhenko, V.D.; Korshunov, V.M.; Taydakov, I.V.; Belousov, Y.A. Mono- and Mixed Metal Complexes of Eu3+, Gd3+, and Tb3+ with a Diketone, Bearing Pyrazole Moiety and CHF2-Group: Structure, Color Tuning, and Kinetics of Energy Transfer between Lanthanide Ions. Molecules 2021, 26, 2655. [Google Scholar] [CrossRef]
  10. Tigaa, R.A.; Ortega, R.E.; Lin, X.; Strouse, G.F. A Versatile Tripodal Ligand for Sensitizing Lanthanide (LnIII) Ions and Color Tuning. Chemistry 2021, 3, 138–145. [Google Scholar] [CrossRef]
  11. Kaya, E.E.; Kaya, O.; Stopic, S.; Gürmen, S.; Friedrich, B. NdFeB Magnets Recycling Process: An Alternative Method to Produce Mixed Rare Earth Oxide from Scrap NdFeB Magnets. Metals 2021, 11, 716. [Google Scholar] [CrossRef]
  12. Choi, E.Y.; Jeong, S.M. Electrochemical processing of spent nuclear fuels: An overview of oxide reduction in pyroprocessing technology. Prog. Nat. Sci. 2015, 25, 572–582. [Google Scholar] [CrossRef]
  13. Khan, M.M.; Ansari, S.A.; Pradhan, D.; Han, D.H.; Lee, J.; Cho, M.H. Defect-induced band gap narrowed CeO2 nanostructures for visible light activities. Ind. Eng. Chem. Res. 2014, 53, 9754–9763. [Google Scholar] [CrossRef]
  14. Vita, A. Catalytic Applications of CeO2-Based Materials. Catalysts 2020, 10, 576. [Google Scholar] [CrossRef]
  15. Zou, F.R.; Wang, S.N.; Wang, F.F.; Liu, D.; Li, Y. Synthesis of Lanthanide Functionalized Carbon Quantum Dots for Chemical Sensing and Photocatalytic Application. Catalysts 2020, 10, 833. [Google Scholar] [CrossRef]
  16. Kim, D. Recent Developments in Lanthanide-Doped Alkaline Earth Aluminate Phosphors with Enhanced and Long-Persistent Luminescence. Nanomaterials 2021, 11, 723. [Google Scholar] [CrossRef]
  17. Arunachalam, S.; Kirubasankar, B.; Pan, D.; Liu, H.; Yan, C.; Guo, Z.; Angaiah, S. Research progress in rare earths and their composites based electrode materials for supercapacitors. Green. Energy Environ. 2020, 5, 259–273. [Google Scholar] [CrossRef]
  18. Malavekar, D.B.; Magdum, V.V.; Khot, S.D.; Kim, J.H.; Lokhande, C.D. Doping of rare earth elements: Towards enhancing the electrochemical performance of pseudocapacitive materials. J. Alloys Compd. 2023, 960, 170601. [Google Scholar] [CrossRef]
  19. Steglich, F.; Aarts, J.; Bredl, C.D.; Lieke, W.; Meschede, D.; Franz, W.; Schäfer, H. Superconductivity in the Presence of Strong Pauli Paramagnetism: CeCu2Si2. Phys. Rev. Lett. 1979, 43, 1892. [Google Scholar] [CrossRef]
  20. He, D.; Xing, S.; Sun, B.; Cai, H.; Suo, H.; Zhao, C. Design and Construction of Three-Dimensional Flower-like CuO Hierarchical Nanostructures on Copper Foam for High Performance Supercapacitor. Electrochim. Acta 2016, 210, 639–645. [Google Scholar] [CrossRef]
  21. Yuan, M.; Gui, X.; Liu, Y.; Pang, H. Si-based materials derived from biomass: Synthesis and applications in electrochemical energy storage. J. Mater. Chem. A 2019, 7, 22123–22147. [Google Scholar] [CrossRef]
  22. Wu, Z.; Zhang, X.; Jin, X.; Li, T.; Ge, J.; Li, Z. A Review on Cutting Edge Technologies of Silicon-Based Supercapacitors. J. Nanomater. 2021, 2021, 6650131. [Google Scholar] [CrossRef]
  23. Liu, Y.; Huang, H.; Peng, X. Highly Enhanced Capacitance of CuO Nanosheets by Formation of CuO/SWCNT Networks through Electrostatic Interaction. Electrochim. Acta 2013, 104, 289–294. [Google Scholar] [CrossRef]
  24. Romanitan, C.; Varasteanu, P.; Mihalache, I.; Culita, D.; Somacescu, S.; Pascu, R.; Tanasa, E.; Eremia, S.A.V.; Boldeiu, A.; Simion, M.; et al. High-performance solid state supercapacitors assembling graphene interconnected networks in porous silicon electrode by electrochemical methods using 2,6-dihydroxynaphthalen. Sci. Rep. 2018, 8, 9654. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Z.; Cheng, Y.; Li, H.; Xu, Z.; Sun, S.; Yin, S. Construction of Cu-Doped Ni−Co-Based Electrodes for High-Performance Supercapacitor Applications. ACS Appl. Energy Mater. 2022, 5, 6642–6653. [Google Scholar] [CrossRef]
  26. Liu, S.; Kang, L.; Hu, J.; Jung, E.; Zhang, J.; Jun, S.C.; Yamauchi, Y. Unlocking the Potential of Oxygen-Deficient Copper-Doped Co3O4 Nanocrystals Confined in Carbon as an Advanced Electrode for Flexible Solid-State Supercapacitors. ACS Energy Lett. 2021, 6, 3011–3019. [Google Scholar] [CrossRef]
  27. Maheswari, N.; Muralidharan, G. Supercapacitor Behavior of Cerium Oxide Nanoparticles in Neutral Aqueous Electrolytes. Energy Fuels 2015, 29, 8246–8253. [Google Scholar] [CrossRef]
  28. Kalubarme, R.S.; Kim, Y.H.; Park, C.J. One Step Hydrothermal Synthesis of a Carbon Nanotube/Cerium Oxide Nanocomposite and its Electrochemical Properties. Nanotechnology 2013, 24, 365401. [Google Scholar] [CrossRef]
  29. Cui, S.; Wang, F.; Sun, K.; Wang, X.; Hu, Q.; Peng, H.; Ma, G.; Lei, Z. High-Performance Hybrid Supercapacitors Based on Ce-Doped NiMoO4 Nanosheets and Fe3O4@Bi2O3 Nanoarrays. J. Phys. Chem. C 2021, 125, 18129c18140. [Google Scholar] [CrossRef]
  30. Saranya, P.E.; Selladurai, S. Mesoporous 3D network Ce-doped NiO nanoflakes as high performance electrodes for supercapacitor applications. New J. Chem. 2019, 43, 7441. [Google Scholar] [CrossRef]
  31. Shi, J.W.; Wang, Y.; Duan, R.; Gao, C.; Wang, B.; He, C.; Niu, C. The synergistic effects between Ce and Cu in CuyCe1−yW5Ox catalysts for enhanced NH3-SCR of NOx and SO2 tolerance. Catal. Sci. Technol. 2019, 9, 718–730. [Google Scholar] [CrossRef]
  32. Rauda, I.E.; Augustyn, V.; Dunn, B.; Tolbert, S.H. Enhancing pseudocapacitive charge storage in polymer templated mesoporous materials. Acc. Chem. Res. 2013, 46, 1113–1124. [Google Scholar] [CrossRef] [PubMed]
  33. Wei, J.; Li, Y.; Dai, D.; Zhang, F.; Zou, H.; Yang, X.; Ji, Y.; Li, B.; Wei, X. Surface Roughness: A Crucial Factor To Robust Electric Double Layer Capacitors. ACS Appl. Mater. Interfaces 2020, 12, 5786–5792. [Google Scholar] [CrossRef] [PubMed]
  34. Sarno, M.; Ponticorvo, E.; Scarpa, D. Controlled PtIr nanoalloy as an electro-oxidation platform for methanol reaction and ammonia detection. Nanotechnology 2019, 30, 394004. [Google Scholar] [CrossRef] [PubMed]
  35. Sarno, M.; Ponticorvo, E.; Scarpa, D. Ru and Os based new electrode for electrochemical flow supercapacitors. Chem. Eng. J. 2019, 3771, 120050. [Google Scholar] [CrossRef]
  36. Shi, X.; Yu, J.; Huang, J.; Chen, B.; Fang, L.; Shao, L.; Sun, Z. Metal-organic framework derived high-content N, P and O-codoped Co/C composites as electrode materials for high performance supercapacitors. J. Power Sources 2020, 467, 228304. [Google Scholar] [CrossRef]
  37. Tornheim, A.; O’Hanlon, D.C. What do Coulombic efficiency and capacity retention truly measure? A deep dive into cyclable lithium inventory, limitation type, and redox side-reactions. J. Electrochem. Soc. 2020, 167, 110520. [Google Scholar] [CrossRef]
  38. Holter, G.; Adrian, H. Thin film preparation, superconductivity and transport properties of the heavy-fermion system CeCu2Si2. Solid. State Commun. 1986, 58, 45–49. [Google Scholar] [CrossRef]
  39. Stockert, O.; Arndt, J.; Faulhaber, E.; Geibel, C.; Jeevan, H.S.; Kirchner, S.; Loewenhaupt, M.; Schmalzl, K.; Schmidt, W.; Si, Q. Magnetically driven superconductivity in CeCu2Si2. Nat. Phys. 2010, 7, 119–124. [Google Scholar] [CrossRef]
  40. Shatruk, M. ThCr2Si2 Structure Type: The “Perovskite” of Intermetallics. J. Solid. State Chem. 2019, 272, 198–209. [Google Scholar] [CrossRef]
  41. Conway, B.E. Electrochemical Supercapacitors: Scientific Fundamentals and Technological Applications; Kluwer Academic/Plenum Publishers: New York, NY, USA, 1999. [Google Scholar]
  42. Ahuja, P.; Sahu, V.; Ujjain, S.K.; Sharma, R.K.; Singh, G. Performance evaluation of Asymmetric Supercapacitor based on Cobalt manganite modified graphene nanoribbons. Electrochim. Acta 2014, 146, 429–436. [Google Scholar] [CrossRef]
  43. Ujjain, S.K.; Singh, G.; Sharma, R.K. Co3O4@Reduced Graphene Oxide Nanoribbon for high performance Asymmetric Supercapacitor. Electrochim. Acta 2015, 169, 276–282. [Google Scholar] [CrossRef]
  44. Ahuja, P.; Ujjain, S.K.; Kanojia, R. Electrochemical Behaviour of Manganese & Ruthenium mixed Oxide@ Reduced Graphene Oxide Nanoribbon Composite in Symmetric and Asymmetric Supercapacitor. Appl. Surf. Sci. 2017, 427, 102–111. [Google Scholar]
  45. Rakhi, R.B.; Chen, W.; Cha, D.; Alshareef, H.N. Substrate Dependent Self-Organization of Mesoporous Cobalt Oxide Nanowires with Remarkable Pseudocapacitance. Nano Lett. 2012, 12, 2559–2567. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, J.; Liu, W.; Niu, H.; Cheng, K.; Ye, K.; Zhu, K.; Wang, G.; Cao, D.; Yan, J. Ultrahigh energy density battery-type asymmetric supercapacitors: NiMoO4 nanorod-decorated graphene and graphene/Fe2O3 quantum dots. Nano Res. 2018, 11, 4744–4758. [Google Scholar] [CrossRef]
  47. Liu, W.; Niu, H.; Yang, J.; Cheng, K.; Ye, K.; Zhu, K.; Wang, G.; Cao, D.; Yan, J. Ternary Transition Metal Sulfides Embedded in Graphene Nanosheets as Both the Anode and Cathode for High-Performance Asymmetric Supercapacitors. Chem. Mater. 2018, 30, 1055–1068. [Google Scholar] [CrossRef]
  48. Masarapu, C.; Zeng, H.F.; Hung, K.H.; Wei, B. Effect of Temperature on the Capacitance of Carbon Nanotube Supercapacitors. ACS Nano 2009, 3, 2199–2206. [Google Scholar] [CrossRef]
  49. Chen, K.; Pan, W.; Xue, D. Phase Transformation of Ce3+-Doped MnO2 for Pseudocapacitive Electrode Materials. J. Phys. Chem. C 2016, 120, 20077–20081. [Google Scholar] [CrossRef]
  50. Su, X.; Yu, L.; Cheng, G.; Zhang, H.; Sun, M.; Zhang, L.; Zhang, J. Controllable hydrothermal synthesis of Cu-doped δ-MnO2 films with different morphologies for energy storage and conversion using supercapacitors. Appl. Energy 2014, 134, 439–445. [Google Scholar] [CrossRef]
  51. Lavanya, T.; Ramaprabhu, S. Copper nanoparticles incorporated porous carbon nanofibers as a freestanding binder-free electrode for symmetric supercapacitor with enhanced electrochemical performance. Mater. Res. Express 2019, 6, 105005. [Google Scholar] [CrossRef]
  52. Oh, I.; Kim, M.; Kim, J. Carbon-coated Si/MnO2 nanoneedle composites with optimum carbon layer activation for supercapacitor applications. Chem. Eng. J. 2015, 273, 82–91. [Google Scholar] [CrossRef]
  53. Liu, Q.; Nayfeh, M.H.; Yau, S.T. Supercapacitor electrodes based on polyaniline–silicon nanoparticle composite. J. Power Sources 2010, 195, 3956–3959. [Google Scholar] [CrossRef]
Figure 1. SEM images of CeCu2Si2.
Figure 1. SEM images of CeCu2Si2.
Materials 16 07182 g001
Figure 2. Nitrogen adsorption isotherms at 77 K of CeCu2Si2.
Figure 2. Nitrogen adsorption isotherms at 77 K of CeCu2Si2.
Materials 16 07182 g002
Figure 3. EDX analysis on CeCu2Si2 with an SEM image and the corresponding EDX maps for Ce, Si, and Cu.
Figure 3. EDX analysis on CeCu2Si2 with an SEM image and the corresponding EDX maps for Ce, Si, and Cu.
Materials 16 07182 g003
Figure 4. X-ray diffraction pattern of CeCu2Si2.
Figure 4. X-ray diffraction pattern of CeCu2Si2.
Materials 16 07182 g004
Figure 5. (a) CV curves of CeCu2Si2 at different scan rates. (b) CV curves of the bare glassy carbon electrode at different scan rates. In the inset, a magnification of the same graph, with the current values expressed in µA, is reported.
Figure 5. (a) CV curves of CeCu2Si2 at different scan rates. (b) CV curves of the bare glassy carbon electrode at different scan rates. In the inset, a magnification of the same graph, with the current values expressed in µA, is reported.
Materials 16 07182 g005
Figure 6. Galvanostatic charge–discharge curves of CeCu2Si2 at different current densities.
Figure 6. Galvanostatic charge–discharge curves of CeCu2Si2 at different current densities.
Materials 16 07182 g006
Figure 7. (a) Mass capacitance vs. current density, (b) mass capacitances of CeCu2Si2 at different scan rates, (c) capacitance retention of CeCu2Si2 as a function of cycle number, (d) Coulombic efficiency as a function of cycle number, and (e) Nyquist plot for CeCu2Si2.
Figure 7. (a) Mass capacitance vs. current density, (b) mass capacitances of CeCu2Si2 at different scan rates, (c) capacitance retention of CeCu2Si2 as a function of cycle number, (d) Coulombic efficiency as a function of cycle number, and (e) Nyquist plot for CeCu2Si2.
Materials 16 07182 g007
Table 1. Supercapacitive performance of the CeCu2Si material and some of the best-performing Cu, Ce, and/or Si-containing electrode materials for supercapacitor applications in the literature.
Table 1. Supercapacitive performance of the CeCu2Si material and some of the best-performing Cu, Ce, and/or Si-containing electrode materials for supercapacitor applications in the literature.
Cu, Ce, Si-Containing Electrode Material (with Average Size)CV Potential Window (V)GCD Current Density (A/g)Mass Capacitance (F/gsample)Capacitance RetentionRef.
Ce-doped NiO nanoflakes (27–30 nm)0 ÷ 0.45 vs. Ag/AgCl11775retains about 93% after 2000 cycles at 5 A/g[30]
CeO2 nanoparticles (14 nm)0 ÷ 0.8 vs. Ag/AgCl2457retains about 82% after 2000 cycles at 10 A/g[27]
Ce-doped MnO2 nanorods (10–20 nm)0 ÷ 0.8 vs. SCE1101.1retains about 99.5% after 1000 cycles at 5 A/g[49]
Cu-doped-MnO2 nanosheets0 ÷ 1 vs. SCE1296retains about 79% after 1000 cycles at 2 A/g[50]
Ni–Co–Cu oxide nanorods (10–40 nm)−0.1 ÷ 0.65 vs. Hg/HgO36.54 F/cm2the device retains about 40% after 2000 cycles at 2 A/g[25]
Cu nanoparticles/PCNFs0 ÷ 11333.5retains about 95.8% after 10,000 cycles at 3 A/g[51]
Si/MnO2 nanoneedle (20–40 nm)−0.2 ÷ 0.8 vs. Ag/AgCl1240.1retains about 85.2% after 2000 cycles at 1 A/g[52]
PANI-Si nanoparticles (1–2.8 nm)−0.2 ÷ 1 vs. Ag/AgCl5 mA/cm2470retains 78% after 1500 cycles[53]
CeCu2Si2 rough micro pebbles−0.4 ÷ 0.7 vs. SCE1278retains about 98% after 20,000 cycles at 10 A/gThis work
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

Scarpa, D.; Cirillo, C.; Luciano, C.; Nigro, A.; Adami, R.; Cirillo, C.; Attanasio, C.; Iuliano, M.; Ponticorvo, E.; Sarno, M. Rough and Porous Micropebbles of CeCu2Si2 for Energy Storage Applications. Materials 2023, 16, 7182. https://doi.org/10.3390/ma16227182

AMA Style

Scarpa D, Cirillo C, Luciano C, Nigro A, Adami R, Cirillo C, Attanasio C, Iuliano M, Ponticorvo E, Sarno M. Rough and Porous Micropebbles of CeCu2Si2 for Energy Storage Applications. Materials. 2023; 16(22):7182. https://doi.org/10.3390/ma16227182

Chicago/Turabian Style

Scarpa, Davide, Claudia Cirillo, Christopher Luciano, Angela Nigro, Renata Adami, Carla Cirillo, Carmine Attanasio, Mariagrazia Iuliano, Eleonora Ponticorvo, and Maria Sarno. 2023. "Rough and Porous Micropebbles of CeCu2Si2 for Energy Storage Applications" Materials 16, no. 22: 7182. https://doi.org/10.3390/ma16227182

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