Next Article in Journal
Performance of Micronized Biowax Powders Replacing PTFE Fillers in Bio-Based Epoxy Resin Coatings
Previous Article in Journal
Electrodeposited CoFeNi Medium-Entropy Alloy Coating on a Copper Substrate from Chlorides Solution with Enhanced Corrosion Resistance
Previous Article in Special Issue
Hierarchical Porous Nickel Anode with Low Polarization at High Current Density
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical One-Step Synthesis of Cu2O with Tunable Oxygen Defects and Their Electrochemical Performance in Li-Ion Batteries

1
School of Materials and Energy, University of Electronic Science and Technology of China, Chengdu 611731, China
2
School of New Energy Materials and Chemistry, Leshan Normal University, Leshan 614004, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(5), 510; https://doi.org/10.3390/coatings15050510
Submission received: 25 March 2025 / Revised: 16 April 2025 / Accepted: 23 April 2025 / Published: 24 April 2025

Abstract

:
We report a facile galvanic oxidation corrosion method for the preparation of cuprite nanocrystals (Cu2O) with controllable oxygen vacancies. The Cu2O microspheres have been employed as active anode materials in lithium-ion batteries (LIBs), exhibiting excellent electrochemical performance. The effect of oxygen vacancies on the electrochemical properties was studied. The oxygen vacancy-rich Cu2O electrodes exhibited a high specific discharge capacity (1002.3 mAh g−1 at 0.1 C) and remarkable reversibility. Oxygen vacancies in Cu2O not only promote high electronic conductivity but also provide additional active sites for lithiation/delithiation, further enhancing electrochemical performance. Furthermore, the formation mechanism of Cu2O during the galvanic oxidation–corrosion process has been proposed.

1. Introduction

Lithium-ion batteries (LiBs) are widely used as electrochemical energy storage devices, but advancements in performance and cost reduction are crucial for significant progress in sectors such as electric vehicles and renewable energy storage [1,2,3,4]. The performance of these batteries primarily hinges on the characteristics of electrode materials, particularly anode materials, which have a direct impact on energy density, cycling stability, and rate capability [5,6,7]. Consequently, research has increasingly focused on developing anode materials with high specific energy, exceptional cycling stability, and environmental compatibility [8,9,10].
Cuprous oxide (Cu2O), a typical transition metal oxide semiconductor, is considered as a promising anode material due to its high theoretical capacity, abundant resources, and eco-friendly properties [11,12,13]. However, like other metal oxides, Cu2O is hindered by poor conductivity and cycling stability, which limit its potential as an anode material [14]. To overcome these challenges, various strategies have been employed, including particle size control [15,16], customization of microstructural morphology [17], and compositing Cu2O with conductive materials [18,19]. Although many researchers have reported improvements in the electrochemical performance of Cu2O through precise control of the particle size and morphology [20,21], these strategies alone cannot enhance cycling stability due to significant volume expansion and contraction during the charge and discharge processes [22].
To address these issues, researchers have proposed encapsulating nanoscale Cu2O particles within porous carbon nanotubes [23] or forming nanosheet structures by compositing Cu2O microspheres with graphene oxide [24]. These methods have been shown to effectively suppress the volume expansion of Cu2O during the charge–discharge process, resulting in significant improvements in cycling stability. Experimental results have confirmed the effectiveness of these approaches. However, the relatively complex preparation processes and associated high costs present obstacles to large-scale batch production.
Remarkably, recent studies have shown that incorporating oxygen vacancies (OVs) into metal oxide electrodes serves as a cost-effective and efficient strategy to boost the electrochemical performance of lithium-ion batteries (LiBs) [25,26]. These oxygen vacancies not only modify but also optimize the electronic structure and microstructure of the electrodes, resulting in a substantial improvement in their conductivity [27,28]. Additionally, numerous studies have highlighted that oxygen vacancies in metal oxide anodes contribute to enhanced cycling stability and high reversibility in LiBs [29,30]. Dong [31] reported that introducing numerous oxygen vacancies into SnO2 significantly enhances its electrical conductivity. The electrode, which comprises abundant oxygen vacancies in SnO2, refines its microstructure during charge–discharge cycles in LiBs, thus enhancing and stabilizing its capacity. Furthermore, He and colleagues [32] reported that Fe3O4 rich in oxygen vacancies, when used as an anode material for high-performance lithium-ion batteries, exhibits exceptional cycling stability and rate performance, achieving a reversible capacity of 921 mAh/g after 330 cycles.
Despite the promising potential offered by Cu2O nanoparticles with tunable oxygen vacancies as anode materials for lithium-ion batteries (LIBs), there has been limited reporting on this topic. In this study, we introduce a novel electrochemical method for synthesizing Cu2O nanoparticles with tunable oxygen vacancies. The experimental results indicate that the presence of oxygen vacancies significantly enhances the conductivity of Cu2O and introduces additional active sites for ion intercalation and deintercalation, leading to a substantial improvement in its electrochemical performance. Additionally, our work provides insight into the formation mechanism of Cu2O during the galvanic oxidation–corrosion process.

2. Experimental Section

2.1. The Synthesis of Cu2O Nanoparticles with Oxygen Defects

Cu2O nanoparticles, featuring oxygen vacancies, were synthesized in an electrochemical cell with a 100 g/L NaCl electrolyte solution, employing a copper sheet as the anode and a graphite plate as the cathode. A two-electrode system was used to synthesize Cu2O nanoparticles with oxygen vacancies, using a copper sheet as the working electrode, a graphite sheet as the counter electrode, and a 100 g/L NaCl solution as the electrolyte. Asymmetric square-wave alternating current (AC) voltages were applied between the two electrodes, with positive voltages of +3 V, +6 V, and +9 V, and a negative voltage of −3 V. After 40 min of electrochemical reaction, brick-red suspensions were generated. The suspensions were then subjected to repeated filtration and washing to eliminate soluble impurities. The final products were dried in a vacuum oven at 60 °C for 8 h prior to further characterization. The nanoparticles synthesized at positive voltages of 3 V, 6 V, and 9 V are designated as Cu2O-3, Cu2O-6, and Cu2O-9, respectively.

2.2. Characterization

The phase composition and crystalline structure of the synthesized samples were characterized by X-ray diffraction (XRD) with Cu Kα radiation. The morphology and microstructure of the samples were examined using high-resolution transmission electron microscopy (HRTEM). Additionally, the elemental composition and oxidation states were analyzed via X-ray photoelectron spectroscopy (XPS).

2.3. Electrochemical Measurements

CR2032 coin cells were fabricated to perform electrochemical tests on three distinct types of synthesized Cu2O nanoparticles. A homogeneous slurry was uniformly dispersed using a mixture of the synthesized Cu2O nanoparticles (80 wt%), acetylene black (10 wt%), and polyvinylidene fluoride (PVDF) binder (10 wt%) in N-methylpyrrolidone. The slurry was evenly coated onto copper foil and dried under vacuum conditions at 80 °C for 12 h. Thereafter, the coated foil was cut into circular electrode sheets, each with a diameter of 12 mm. The coin cells were assembled in an argon-filled glove box. These circular electrodes were used as the working electrode (anode), while lithium foil acted as the counter electrode. A polypropylene separator served as the separator, and a 1 M LiPF6 solution in EC/DMC (1:1 volume ratio) was used as the electrolyte. Galvanostatic charge–discharge tests were performed using a NEWWARE battery test system, with the voltage ranging from 0.01 V to 3 V versus Li/Li+ at room temperature. Additionally, cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) measurements were conducted on a CHI660E electrochemical workstation. For the CV test, the voltage range was set from 0.01 V to 3 V, with a scan rate of 0.1 mV/s. The EIS measurements were conducted over a frequency range spanning from 0.01 Hz to 100 kHz.

3. Results and Discussion

3.1. Materials Characterization

The phase composition of the samples was determined by analyzing the X-ray diffraction (XRD) patterns illustrated in Figure 1a. The XRD patterns of the three samples exhibit distinct diffraction peaks at 2θ values of 29.582°, 36.441°, 42.328°, 61.406°, 73.556°, and 77.414°, which align with the (110), (111), (200), (220), (311), and (222) crystal planes of Cu2O, as per JCPDS 78-2076 [33]. Notably, the distinct diffraction peaks at 2θ values are 35.504°, 38.735°, 48.66°, and 53.431°, which are attributed to CuO [34]. The intensity of CuO peaks increases with increasing the applied voltage. The Cu2O-9 sample clearly indicates that, under conditions of applied high positive potential, Cu2O may undergo additional oxidation to form CuO. Overall, the XRD results thereby confirm the success of the strategy used to synthesize Cu2O nanoparticles by directly applying an asymmetric square-wave alternating current (AC) voltage to copper sheets.
X-ray photoelectron spectroscopy (XPS) was utilized to investigate the surface composition of the synthesized Cu2O samples and to identify oxygen vacancies. As illustrated in Figure 1b, the signals corresponding to Cu, O, and C were clearly detected in the scanned XPS spectra of all three samples. The C signals detected in the three samples originate from carbon contamination [25]. The high-resolution scans of the O 1s XPS spectra (Figure 2c) and the Cu 2p XPS spectra (Figure 2d) further elucidate the valence states of the elements. Four clearly defined peaks are exhibited in Figure 2d, with their binding energies recorded at 934.30 eV, 942.50 eV, 954.10 eV, and 962.30 eV, respectively. The two peaks at 934.30 eV and 954.10 eV can be assigned to the binding energies of Cu 2p3/2 and Cu 2p1/2, respectively; conversely, the peaks at 942.50 eV and 962.30 eV are attributed to the satellite peaks of Cu2+ [35]. Following the deconvolution of the Cu 2p3/2 and Cu 2p1/2 peaks, both peaks were resolved into two separate components to ascertain the chemical states of Cu. The peaks at 932.78 eV and 952.96 eV are attributed to Cu+, whereas those at 934.89 eV and 955.49 eV are ascribed to Cu2+ [36]. The presence of Cu2+ in the samples implies that CuO was generated during the electrochemical synthesis process, which is consistent with the XRD characterization results.
Similarly, the XPS spectra of the O 1s region can be fitted with distinct peaks centered at 530.75 eV, 531.70 eV, and 532.99 eV, assigned to lattice oxygen (Cu-O bond), oxygen vacancies, and absorbed oxygen, respectively [37]. The relative peak areas indicate a systematic increase in the proportion of oxygen vacancies from the Cu2O-3 to Cu2O-9 samples. This observation implies that a higher applied positive potential during the electrochemical synthesis favors the generation of oxygen vacancies. XPS analysis revealed that the copper-to-oxygen atomic ratios for the samples, namely 0.58:1 (Cu2O-3), 0.65:1 (Cu2O-6), and 0.61:1 (Cu2O-9), deviate significantly from the theoretical stoichiometric ratio of 2:1 for Cu2O. This discrepancy can be ascribed to two factors: firstly, the presence of a substantial proportion of CuO in the synthesized samples; secondly, the contribution of oxygen vacancies, leading to adsorbed oxygen species, which account for the additional oxygen atoms. The XPS results further validate the existence of oxygen vacancies and provide insights into the impact of potential manipulation during the electrochemical synthesis on the Cu2O samples.
Figure 2 illustrates the microscopic morphologies of the three synthesized Cu2O samples. The particle size of Cu2O gradually increases with an increase in the applied positive voltage during synthesis, as illustrated in Figure 2a–c. At a low positive voltage of 3 V, the Cu2O-3 sample exhibits fine spherical particles with diameters of approximately 50 nm. When the voltage is increased to 6 V, the Cu2O-6 sample displays irregular fusiform particles with lengths of approximately 400 nm and widths ranging from 100 to 200 nm. Upon further increasing the voltage to 9 V, the particle size increases significantly, and the particles adopt a more irregular shape. Morphological analyses of the samples synthesized at different voltages reveal that the applied voltage has a significant impact on both the morphology and particle size of Cu2O.
The HRTEM images of the three samples, as depicted in Figure 2d–f, show distinct atomic lattice fringes. The measured lattice plane spacings of 0.246 nm, 0.301 nm, 0.213 nm, and 0.151 nm correspond to the (111), (110), (200), and (220) lattice planes of Cu2O, respectively. The selected-area electron diffraction (SAED) pattern of the Cu2O-6 sample, as shown in Figure 2g, further verifies the high crystallinity of Cu2O. Additionally, lattice plane spacings of 0.171 nm, 0.232 nm, and 0.275 nm, corresponding to the (020), (111), and (110) crystal planes of CuO, were observed in the Cu2O-9 sample, suggesting the presence of CuO. The HRTEM results are in agreement with the XRD findings, both of which confirm the formation of CuO in the Cu2O-9 sample. Based on the XRD, morphological analyses, and HRTEM results, we conclude that high voltage facilitates the oxidation of Cu2O, leading to the formation of CuO.

3.2. Electrochemical Investigation

Figure 3 presents the galvanostatic charge and discharge test curves of the Cu2O sample assembled into a half-cell for this study. Figure 3a,c and e illustrate the voltage-capacity curves for the three battery groups during the first, second, and fiftieth charge–discharge cycles, respectively. The galvanostatic charge–discharge (GCD) curves of the three samples are at 0.1 C. The voltage-capacity curves of the three groups of batteries at the 1st, 2nd, and 50th charging and discharging cycles reveal initial discharging capacities of 569.9 mAh/g, 1002.3 mAh/g, and 823.8 mAh/g, respectively. It is noteworthy that the Cu2O-6 group exhibits the highest initial charging (826.72 mAh/g) and discharging (1002.3 mAh/g) capacities. The first Coulomb efficiencies are 67.53%, 82.48%, and 75.62% for Cu2O-3, Cu2O-6, and Cu2O-9, respectively. The results show that all three materials demonstrate relatively high initial discharge capacities; however, these materials exhibit significantly low first-cycle Coulombic efficiencies, indicating substantial capacity loss. The primary factors contributing to the low Coulombic efficiency observed in the first cycle include irreversible reactions between Cu2O and Li during the charge–discharge processes; and the formation of the solid electrolyte interface (SEI) layer resulting from the inevitable decomposition of the electrolyte [38,39]. During the subsequent cycles (from the second to the fiftieth), the Coulombic efficiencies remain stable at nearly 100%. After fifty cycles, the capacities of the three groups are 346.5 mAh/g, 453.6 mAh/g, and 461.4 mAh/g, respectively, representing retention rates of 80.32%, 66.25%, and 81.77%. It is noteworthy that, despite Cu2O-6 exhibiting the highest initial capacity, its capacity decay is most pronounced after fifty cycles. When compared to previously reported Cu2O nanoparticles, the Cu2O samples with oxygen vacancies introduced through electrochemical methods demonstrate a significantly higher capacity retention ratio.
Figure 4 presents the rate performance of the three samples, which were cycled at current densities ranging from 0.1 C to 2 C, followed by a return to 0.1 C for an additional ten cycles. Specifically, the Cu2O-6 sample exhibited discharge capacities of 631 mAh/g, 565 mAh/g, 493 mAh/g, 412 mAh/g, and 295 mAh/g at current densities of 0.1 C, 0.2 C, 0.5 C, 1 C, and 2 C, respectively. Upon returning to a current density of 0.1 C, the electrode maintained a relatively high discharge capacity of 527 mAh/g, which corresponds to 83.52% of its initial capacity. Similarly, the Cu2O-3 and Cu2O-9 samples also demonstrated excellent rate capabilities, with their capacities returning to 77.37% and 82.55% of their initial values, respectively. At a high current density of 2 C, the discharge capacities of Cu2O-3, Cu2O-6, and Cu2O-9 were 296 mAh/g, 295 mAh/g, and 243 mAh/g, respectively, corresponding to 51.11%, 46.75%, and 45.59% of their respective initial capacities. These results demonstrate that cuprous oxide with oxygen vacancies maintains a relatively high capacity retention ratio under high current conditions. Notably, the Cu2O-6 sample exhibited both the highest discharge capacity and the most favorable rate performance. However, the Cu2O-6 sample experienced a relatively rapid capacity decay during high-current charge–discharge cycling, which is consistent with the results obtained from galvanostatic charge and discharge tests. Collectively, these rate test results reinforce the advantages of oxygen vacancy-rich cuprous oxide as an anode material for lithium-ion batteries, demonstrating its excellent kinetic performance.
The kinetics of the electrochemical reactions within the battery were investigated using electrochemical impedance spectroscopy (EIS), with the resulting Nyquist plots displayed in Figure 5. The Nyquist plots of the three electrodes show a semicircle in the high-frequency range and a linear segment in the low-frequency range. The semicircle in the high-frequency range is representative of the charge transfer process related to the electrochemical reaction at the electrode’s surface. Meanwhile, the linear segment in the low-frequency range signifies the diffusion process of ions moving from the solution to the electrode’s surface [40]. An equivalent circuit model, illustrated in Figure 5, was used to fit the Nyquist plot. In this model, R1 and R2 represent the solution resistance and the charge transfer resistance (Rct), respectively. CPE1 corresponds to the constant phase element arising from the double-layer capacitance between the solution and the electrode. W1 represents the Warburg impedance, which relates to the diffusion resistance of ions within the solution [41]. The fitting results of the equivalent circuit indicate that the Rct values for Cu2O-3, Cu2O-6, and Cu2O-9 are 349.2 Ω, 584.2 Ω, and 596.9 Ω, respectively. It is noteworthy that the Cu2O-3 sample exhibits the lowest Rct value, potentially due to its smaller particle size (as suggested in [42]). These findings from the EIS analysis confirm the potential to improve the conductivity of cuprous oxide by introducing oxygen vacancies.
To further investigate the influence of oxygen vacancies in Cu2O on lithium-ion storage, cyclic voltammetry tests were performed, and the outcomes are depicted in Figure 6. In the cyclic voltammetry (CV) profiles, all samples exhibited a prominent reduction peak within the voltage range of 0.7 V to 0.8 V. This peak is attributed to the formation of a solid electrolyte interphase (SEI) due to an irreversible reduction reaction between the electrolyte and Cu2O [43]. Within the voltage range of 1.40 V to 1.68 V, faint oxidation peaks were observed, primarily linked to the decomposition of the SEI film. For Cu2O-3 and Cu2O-6, consistent oxidation peaks were visible in the range of 2.45 V to 2.62 V across all three cycles, attributed to the oxidation of copper to Cu2O, as reported in previous lithium-ion battery cathode studies [11]. Conversely, for the Cu2O-9 sample, an oxidation peak was noted within this potential range during the initial scan, but subsequent scans showed no significant oxidation peak. Analysis via X-ray diffraction (XRD), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy (XPS) revealed the presence of some copper oxide in the sample, which may affect the lithium-ion storage mechanism.

3.3. Reaction Mechanism

A comprehensive investigation was conducted of the chemical composition, crystal structure, and electrochemical performance of Cu2O enriched with oxygen vacancies. Based on the analysis, the synthesis process and formation mechanism of this Cu2O variant were further explored, as depicted in Figure 7. Cu2O can be synthesized through electrochemical reactions, with oxygen vacancies introduced by adjusting the applied voltages [44]. Specifically, during the positive voltage phase, the copper sheet functions as the anode, undergoing oxidation to form Cu2+. Simultaneously, the graphite electrode serves as the cathode, where H+ ions, dissociated from water, are reduced to H2, leading to an increase in the solution’s pH. Conversely, when negative voltages are applied, the copper sheet acts as the cathode and undergoes a reduction reaction, converting Cu2+ ions to Cu+ [45]. Initially, a substantial amount of Cu2+ is produced, giving the solution a blue color. The high Cl concentration and low pH in the solution facilitate the complexation reaction between Cu2+ and Cl, forming [CuCl4]2 and resulting in a gradual color change to light green. These experimental observations align with the preceding analytical results. Initially, the solution’s color rapidly shifts from light blue to light green. As the electrochemical reaction proceeds, hydroxide ions accumulate, causing an increase in the solution’s pH. Consequently, Cu+ reacts with OH to form Cu2O particles, which remain suspended in the solution and gradually turn it brick-red. During the negative scanning process, these Cu2O particles further acquire electrons from the copper electrode, resulting in the formation of oxygen vacancies [46,47].
At a positive voltage of 3 V, the electrochemical reaction proceeds at a relatively slow rate, resulting in the formation of smaller cuprous oxide particles. This refinement leads to a substantial increase in the material’s specific surface area, exposing more active sites for reactions and reducing the transmission distance for Li+ and e. These factors collectively contribute to the lower charge transfer resistance observed in Cu2O-3. As the positive voltage increases, the reaction rate at the electrode surface accelerates, leading to the rapid growth of Cu2O particles. This rapid growth results in the formation of larger particles and the aggregation of smaller ones [48]. Furthermore, a portion of Cu2O undergoes oxidation to form CuO, accounting for the presence of a certain amount of cupric oxide in the Cu2O-9 sample.

4. Conclusions

In this study, we employed a versatile, cost-effective electrochemical process to synthesize Cu2O nanoparticles with a tunable concentration of oxygen vacancies. Three samples were successfully synthesized using symmetric square-wave alternating current (AC) voltages, and the existence of oxygen vacancies in these samples was confirmed through X-ray photoelectron spectroscopy (XPS). These samples were subsequently utilized as anode materials in lithium-ion batteries (LIBs), which demonstrated exceptional electrochemical performance. The presence of oxygen vacancies enhanced the conductivity and capacity of the LIBs, resulting in a high initial discharge capacity of 1002.3 mAh/g at a rate of 0.1 C, along with high reversibility. A mechanism for the electrochemical reaction of Cu2O with oxygen vacancies was proposed, which is in agreement with our experimental results. The significance of this research lies in its introduction of oxygen vacancies (Ovs) into Cu2O, thereby paving the way for new avenues of research in the field of lithium-ion batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/coatings15050510/s1.

Author Contributions

Conceptualization, H.T.; Methodology, Y.Z.; Investigation, F.J. and S.Z.; Data curation, L.H.; Writing—original draft, Y.Z.; Writing—review & editing, H.T. and L.H.; Visualization, W.T.; Project administration, L.H. and H.T.; Funding acquisition, W.T. and H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wu, F.; Maier, J.; Yu, Y. Guidelines and trends for next-generation rechargeable lithium and lithium-ion batteries. Chem. Soc. Rev. 2020, 49, 1569–1614. [Google Scholar] [CrossRef] [PubMed]
  2. Goodenough, J.B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef]
  3. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  4. Venkata, S.A.; Mewin, V.; Ivan, J.F.; Steven, J.H.; Vinodkumar, E. Unusual pseudocapacitive lithium-ion storage on defective Co3O4 nanosheets. Nanotechnology 2022, 33, 225403. [Google Scholar] [CrossRef]
  5. Nzereogu, P.U.; Omah, A.D.; Ezema, F.I.; Iwuoha, E.I.; Nwanya, A.C. Anode materials for lithium-ion batteries: A review. Appl. Surf. Sci. Adv. 2022, 9, 100233. [Google Scholar] [CrossRef]
  6. Chang, H.; Wu, Y.-R.; Han, X.; Yi, T.-F. Recent developments in advanced anode materials for lithium-ion batteries. Energy Mater. 2022, 1, 100003. [Google Scholar] [CrossRef]
  7. Arnaud, B.L.; Tobias, L.; Sebastian, M.; Karoline, S.; Sebastian, R.; Christian, H.; Alexander, M. Impact of Electrode Defects on Battery Cell Performance: A Review. Batter. Supercaps 2022, 5, 202200239. [Google Scholar] [CrossRef]
  8. Armand, M.; Tarascon, J.M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef]
  9. Rojaee, R.; Shahbazian-Yassar, R. Two-Dimensional Materials to Address the Lithium Battery Challenges. ACS Nano 2020, 14, 2628–2658. [Google Scholar] [CrossRef]
  10. Gao, Z.; Sun, H.; Fu, L.; Ye, F.; Zhang, Y.; Luo, W.; Huang, Y. Promises, Challenges, and Recent Progress of Inorganic Solid-State Electrolytes for All-Solid-State Lithium Batteries. Adv. Mater. 2018, 30, 1705702. [Google Scholar] [CrossRef]
  11. Liu, W.; Chen, G.; He, G.; Zhang, W. Synthesis of starfish-like Cu2O nanocrystals through γ-irradiation and their application in lithium-ion batteries. J. Nanopart. Res. 2011, 13, 2705–2713. [Google Scholar] [CrossRef]
  12. Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J.-M. Nano-sized transition-metal oxides as negative-electrode materials for lithium-ion batteries. Nature 2000, 407, 496–499. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, J.; Shen, S.; Guo, P.; Wang, M.; Wu, P.; Wang, X.; Guo, L. In-situ reduction synthesis of nano-sized Cu2O particles modifying g-C3N4 for enhanced photocatalytic hydrogen production. Appl. Catal. B Environ. 2014, 152–153, 335–341. [Google Scholar] [CrossRef]
  14. Yang, Y.; Wang, K.; Yang, Z.; Zhang, Y.; Gu, H.; Zhang, W.; Li, E.; Zhou, C. An efficient route to Cu2O nanorod array film for high-performance Li-ion batteries. Thin Solid Film. 2016, 608, 79–87. [Google Scholar] [CrossRef]
  15. Grugeon, S.; Laruelle, S.; Herrera-Urbina, R.; Dupont, L.; Poizot, P.; Tarascon, J. Particle size effects on the electrochemical performance of copper oxides toward lithium. J. Electrochem. Soc. 2001, 148, A285. [Google Scholar] [CrossRef]
  16. Fu, L.J.; Gao, J.; Zhang, T.; Cao, Q.; Yang, L.C.; Wu, Y.P.; Holze, R.; Wu, H.Q. Preparation of Cu2O particles with different morphologies and their application in lithium ion batteries. J. Power Sources 2007, 174, 1197–1200. [Google Scholar] [CrossRef]
  17. Chen, X.; Yu, C.; Guo, X.; Bi, Q.; Sajjad, M.; Ren, Y.; Zhou, X.; Liu, Z. Cu2O Nanoparticles and Multi-Branched Nanowires as Anodes for Lithium-Ion Batteries. Nano 2018, 13, 1850103. [Google Scholar] [CrossRef]
  18. Shen, X.; Chen, S.; Mu, D.; Wu, B.; Wu, F. Novel synthesis and electrochemical performance of nano-structured composite with Cu2O embedment in porous carbon as anode material for lithium ion batteries. J. Power Sources 2013, 238, 173–179. [Google Scholar] [CrossRef]
  19. Zhang, F.; Li, Y.-H.; Qi, M.-Y.; Tang, Z.-R.; Xu, Y.-J. Boosting the activity and stability of Ag-Cu2O/ZnO nanorods for photocatalytic CO2 reduction. Appl. Catal. B Environ. 2020, 268, 118380. [Google Scholar] [CrossRef]
  20. Wang, Y.; Cao, L.; Li, J.; Kou, L.; Huang, J.; Feng, Y.; Chen, S. Cu/Cu2O@Ppy nanowires as a long-life and high-capacity anode for lithium ion battery. Chem. Eng. J. 2020, 391, 123597. [Google Scholar] [CrossRef]
  21. Kang, W.; Liu, F.; Su, Y.; Wang, D.; Shen, Q. The catanionic surfactant-assisted syntheses of 26-faceted and hexapod-shaped Cu2O and their electrochemical performances. CrystEngComm 2011, 13, 4174. [Google Scholar] [CrossRef]
  22. He, Y.; Pu, Y.; Zhu, B.; Zhu, H.; Wang, C.; Tang, W.; Tang, H. Electrochemical one-step synthesis of Mn3O4 with tunable oxygen defects for high-performance aqueous zinc-ion batteries. J. Alloys Compd. 2023, 934, 167933. [Google Scholar] [CrossRef]
  23. Park, J.C.; Kim, J.; Kwon, H.; Song, H. Gram-Scale Synthesis of Cu2O Nanocubes and Subsequent Oxidation to CuO Hollow Nanostructures for Lithium-Ion Battery Anode Materials. Adv. Mater. 2009, 40, 803–807. [Google Scholar] [CrossRef]
  24. Yan, G.-C.; Li, X.-H.; Wang, Z.-X.; Guo, H.-J.; Zhang, Q.; Peng, W.-J. Synthesis of Cu2O/reduced graphene oxide composites as anode materials for lithium ion batteries. Trans. Nonferrous Met. Soc. China 2013, 23, 3691–3696. [Google Scholar] [CrossRef]
  25. He, X.; He, Y.; Wang, C.; Zhu, B.; Liu, A.; Tang, H. Oxygen vacancy-rich Fe3O4 nanoparticle synthesized via facile electrochemical method as anode material for high-performance lithium-ion batteries. J. Phys. Chem. Solids 2022, 171, 111028. [Google Scholar] [CrossRef]
  26. Xu, Y.; Jian, G.; Liu, Y.; Zhu, Y.; Zachariah, M.R.; Wang, C. Superior electrochemical performance and structure evolution of mesoporous Fe2O3 anodes for lithium-ion batteries. Nano Energy 2014, 3, 26–35. [Google Scholar] [CrossRef]
  27. Cao, K.; Jin, T.; Yang, L.; Jiao, L. Recent progress in conversion reaction metal oxide anodes for Li-ion batteries. Mater. Chem. Front. 2017, 1, 2213–2242. [Google Scholar] [CrossRef]
  28. Yang, G.; Yang, S.; Shen, J.; Cui, Y.; Sun, J.; Duan, G.; Cao, B.; Liu, Z. KOH-Induced Oxygen-Deficient VO2 for High-Rate Aqueous Zn-Ion Batteries. ACS Appl. Energy Mater. 2023, 6, 1871–1876. [Google Scholar] [CrossRef]
  29. Qiao, S.; Chen, Y.; Tang, Y.; Yuan, J.; Shen, J.; Zhang, D.; Du, Y.; Li, Z.; Yuan, D.; Tang, H.; et al. Oxygen vacancy–rich Cu2O@Cu with a hydrophobic microenvironment for highly selective C–C coupling to generate C2H4. Chem. Eng. J. 2023, 454, 140321. [Google Scholar] [CrossRef]
  30. Yang, S.; Li, F.; Fu, P.; Zhen, C.; Wu, J.; Lu, H.; Hou, C.; Sheng, Z. Nickel-doped δ-MnO2 abundant in oxygen vacancies as a cathode material for aqueous Zn-ion batteries with superior performance. Nanoscale 2025, 17, 7423–7433. [Google Scholar] [CrossRef]
  31. Zhu, M.; Ullah, I.; Zhang, X.; Long, Y.; Gao, X.; Liu, Z.; Yang, Y.; Shen, G. Phenolphthalein-Modified Multivalent Vanadium Oxide: Harnessing Oxygen Vacancies and Hydrophobicity for Improved Aqueous Zinc-Ion Battery Performance. ACS Sustain. Chem. Eng. 2024, 12, 4255–4263. [Google Scholar] [CrossRef]
  32. Dong, W.; Xu, J.; Wang, C.; Lu, Y.; Liu, X.; Wang, X.; Yuan, X.; Wang, Z.; Lin, T.; Sui, M.; et al. A Robust and Conductive Black Tin Oxide Nanostructure Makes Efficient Lithium-Ion Batteries Possible. Adv. Mater. 2017, 29, 1700136. [Google Scholar] [CrossRef]
  33. Zhang, X.; Zhang, Y.; Huang, H.; Cai, J.; Ding, K.; Lin, S. Electrochemical fabrication of shape-controlled Cu2O with spheres, octahedrons and truncated octahedrons and their electrocatalysis for ORR. New J. Chem. 2018, 42, 458–464. [Google Scholar] [CrossRef]
  34. Wang, C.; Higgins, D.; Wang, F.; Li, D.; Liu, R.; Xia, G.; Li, N.; Li, Q.; Xu, H.; Wu, G. Controlled synthesis of micro/nanostructured CuO anodes for lithium-ion batteries. Nano Energy 2014, 9, 334–344. [Google Scholar] [CrossRef]
  35. Wang, X.H.; Wang, Z.M.; Hong, Q.L.; Zhang, Z.N.; Shi, F.; Li, D.S.; Li, S.N.; Chen, Y. Oxygen-Vacancy-Rich Cu2O Hollow Nanocubes for Nitrate Electroreduction Reaction to Ammonia in a Neutral Electrolyte. Inorg. Chem. 2022, 61, 15678–15685. [Google Scholar] [CrossRef] [PubMed]
  36. Grigore, M.; Biscu, E.; Holban, A.; Gestal, M.; Grumezescu, A. Methods of Synthesis, Properties and Biomedical Applications of CuO Nanoparticles. Pharmaceuticals 2016, 9, 75. [Google Scholar] [CrossRef]
  37. Xie, M.; Lin, M.; Feng, C.; Liu, Z.; Xu, Y.; Wang, N.; Zhang, X.; Jiao, Y.; Chen, J. Coupling Zn2+ doping and rich oxygen vacancies in MnO2 nanowire toward advanced aqueous zinc-ion batteries. J. Colloid Interface Sci. 2023, 645, 400–409. [Google Scholar] [CrossRef]
  38. Xu, Y.-T.; Guo, Y.; Li, C.; Zhou, X.-Y.; Tucker, M.C.; Fu, X.-Z.; Sun, R.; Wong, C.-P. Graphene oxide nano-sheets wrapped Cu2O microspheres as improved performance anode materials for lithium ion batteries. Nano Energy 2015, 11, 38–47. [Google Scholar] [CrossRef]
  39. Cheng, H.; Shapter, J.G.; Li, Y.; Gao, G. Recent progress of advanced anode materials of lithium-ion batteries. J. Energy Chem. 2021, 57, 451–468. [Google Scholar] [CrossRef]
  40. Fu, Y.; Wei, Q.; Zhang, G.; Wang, X.; Zhang, J.; Hu, Y.; Wang, D.; Zuin, L.; Zhou, T.; Wu, Y.; et al. High-Performance Reversible Aqueous Zn-Ion Battery Based on Porous MnOx Nanorods Coated by MOF-Derived N-Doped Carbon. Adv. Energy Mater. 2018, 8, 1801445. [Google Scholar] [CrossRef]
  41. Wang, J.; Zhou, H.; Nanda, J.; Braun, P.V. Three-Dimensionally Mesostructured Fe2O3 Electrodes with Good Rate Performance and Reduced Voltage Hysteresis. Chem. Mater. 2015, 27, 150414100216000. [Google Scholar] [CrossRef]
  42. Chen, R.; Wang, Y.; Nuli, Y.; Yu, Y.; Gao, P.; Chen, Q.; Wei, L.; Hu, N.; Yang, Z.; Gao, R. Cu2O nanowires as anode materials for Li-ion rechargeable batteries. Sci. China Technol. Sci. 2014, 57, 1073–1076. [Google Scholar] [CrossRef]
  43. He, X.; Guo, H.; Zhang, X.; Liao, T.; Zhu, Y.; Tang, H.; Li, T.; Chen, J.S. Facile electrochemical fabrication of magnetic Fe3O4 for electrocatalytic synthesis of ammonia used for hydrogen storage application. Int. J. Hydrogen Energy 2021, 46, 24128–24134. [Google Scholar] [CrossRef]
  44. Zhao, W.; Fu, W.; Yang, H.; Tian, C.; Li, M.; Li, Y.; Zhang, L.; Sui, Y.; Zhou, X.; Chen, H.; et al. Electrodeposition of Cu2O films and their photoelectrochemical properties. CrystEngComm 2011, 13, 2871. [Google Scholar] [CrossRef]
  45. Hossain, M.A.; Al-Gaashani, R.; Hamoudi, H.; Al Marri, M.J.; Hussein, I.A.; Belaidi, A.; Merzougui, B.A.; Alharbi, F.H.; Tabet, N. Controlled growth of Cu2O thin films by electrodeposition approach. Mater. Sci. Semicond. Process. 2017, 63, 203–211. [Google Scholar] [CrossRef]
  46. Kim, D.; Dimitrakopoulos, G.; Yildiz, B. Controlling the Size of Au Nanoparticles on Reducible Oxides with the Electrochemical Potential. J. Am. Chem. Soc. 2022, 144, 21926–21938. [Google Scholar] [CrossRef] [PubMed]
  47. Taher, S.J.; Barzinjy, A.A.; Hamad, S.M. The Effect of Deposition Time on the Properties of Cu2O Nanocubes Using an Electrochemical Deposition Method. J. Electron. Mater. 2020, 49, 7532–7540. [Google Scholar] [CrossRef]
  48. Guo:, D.; Li, W.; Wang, D.; Meng, B.; Fang, D.; Wei, Z. High performance Cu2O film/ZnO nanowires self-powered photodetector by electrochemical deposition. Chin. Phys. B 2020, 29, 098504. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns, (b) XPS spectra, (c) O 1s XPS, and (d) Cu 2p spectra of samples Cu2O-3, Cu2O-6, and Cu2O-9 samples.
Figure 1. (a) XRD patterns, (b) XPS spectra, (c) O 1s XPS, and (d) Cu 2p spectra of samples Cu2O-3, Cu2O-6, and Cu2O-9 samples.
Coatings 15 00510 g001
Figure 2. (ac) TEM images and (df) HRTEM image of Cu2O-3, Cu2O-6, Cu2O-9, (g) SAED pattern of the Cu2O-6 sample, (h) FFT and (i) inverse FFT analysis of Cu2O-6.
Figure 2. (ac) TEM images and (df) HRTEM image of Cu2O-3, Cu2O-6, Cu2O-9, (g) SAED pattern of the Cu2O-6 sample, (h) FFT and (i) inverse FFT analysis of Cu2O-6.
Coatings 15 00510 g002
Figure 3. (a,c,e) Initial discharge–charge profiles, and (b,d,f) cycle performance of Cu2O-3, Cu2O-6, and Cu2O-9 samples, respectively.
Figure 3. (a,c,e) Initial discharge–charge profiles, and (b,d,f) cycle performance of Cu2O-3, Cu2O-6, and Cu2O-9 samples, respectively.
Coatings 15 00510 g003
Figure 4. Rate performance of Cu2O-3, Cu2O-6, and Cu2O-9 electrodes.
Figure 4. Rate performance of Cu2O-3, Cu2O-6, and Cu2O-9 electrodes.
Coatings 15 00510 g004
Figure 5. Electrochemical impedance spectra of Cu2O-3, Cu2O-6, and Cu2O-9 electrodes.
Figure 5. Electrochemical impedance spectra of Cu2O-3, Cu2O-6, and Cu2O-9 electrodes.
Coatings 15 00510 g005
Figure 6. CV curves of the synthesized Cu2O (a) Cu2O-3, (b) Cu2O-6, and (c) Cu2O-9 electrodes.
Figure 6. CV curves of the synthesized Cu2O (a) Cu2O-3, (b) Cu2O-6, and (c) Cu2O-9 electrodes.
Coatings 15 00510 g006
Figure 7. Schematic diagram of the electrochemical synthesis process of Cu2O.
Figure 7. Schematic diagram of the electrochemical synthesis process of Cu2O.
Coatings 15 00510 g007
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

Zheng, Y.; Huang, L.; Jian, F.; Zhao, S.; Tang, W.; Tang, H. Electrochemical One-Step Synthesis of Cu2O with Tunable Oxygen Defects and Their Electrochemical Performance in Li-Ion Batteries. Coatings 2025, 15, 510. https://doi.org/10.3390/coatings15050510

AMA Style

Zheng Y, Huang L, Jian F, Zhao S, Tang W, Tang H. Electrochemical One-Step Synthesis of Cu2O with Tunable Oxygen Defects and Their Electrochemical Performance in Li-Ion Batteries. Coatings. 2025; 15(5):510. https://doi.org/10.3390/coatings15050510

Chicago/Turabian Style

Zheng, Yu, Lanxiang Huang, Feiyu Jian, Shujia Zhao, Wu Tang, and Hui Tang. 2025. "Electrochemical One-Step Synthesis of Cu2O with Tunable Oxygen Defects and Their Electrochemical Performance in Li-Ion Batteries" Coatings 15, no. 5: 510. https://doi.org/10.3390/coatings15050510

APA Style

Zheng, Y., Huang, L., Jian, F., Zhao, S., Tang, W., & Tang, H. (2025). Electrochemical One-Step Synthesis of Cu2O with Tunable Oxygen Defects and Their Electrochemical Performance in Li-Ion Batteries. Coatings, 15(5), 510. https://doi.org/10.3390/coatings15050510

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