Next Article in Journal
Computational Biocompatibility and Safety Evaluation of Metal-Doped PET-Carbon Quantum Dots via Multi-Target Molecular Docking and ADMET Analysis on Human Proteins
Previous Article in Journal
Influence of OH Groups of Hydroxyfullerene on the Mechanism of Its Complex Formation with the Lys-2Gly Peptide Dendrimer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co3O4/SnO2 Hybrid Nanorods as High-Capacity Anodes for Lithium-Ion Batteries

School of Chemistry, Beihang University, Beijing 100191, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Physchem 2025, 5(4), 54; https://doi.org/10.3390/physchem5040054 (registering DOI)
Submission received: 28 September 2025 / Revised: 9 November 2025 / Accepted: 8 December 2025 / Published: 10 December 2025
(This article belongs to the Section Electrochemistry)

Abstract

With the surging demand for high-performance energy storage devices, enhancing the energy density and charge-discharge efficiency of lithium-ion batteries has become an urgent need. Co3O4, with a high theoretical specific capacity of 890 mAh g−1, is regarded as a promising anode candidate. In this work, rod-like hybrid Co3O4/SnO2 composites were successfully prepared via the pyrolysis of cobalt-tin ethylene glycolate precursor. Notably, when the Co/Sn molar ratio is tuned to 3.8:1, the product evolves into nanorods. Lithium-ion batteries using Co3.8Sn1 as the anode deliver an initial specific capacity of 1588.9 mAh g−1, and retain a reversible capacity of 427.9 mAh g−1 after 500 cycles at 2 A g−1, demonstrating that Sn-doping-induced optimization of morphology and conductivity effectively enhances electrochemical performance.

Graphical Abstract

1. Introduction

With the rapid development of the pure electric vehicle market, there is a growing demand for lithium-ion batteries (LIBs) with higher energy density and improved charge-discharge efficiency [1]. However, graphite, the dominant commercial electrode material in commercial LIBs, has inherent limitations due to its relatively low theoretical specific capacity (372 mAh g−1). Furthermore, it exhibits significant polarization at high current rates, making it increasingly difficult to meet the requirements of next-generation high-performance LIBs [2]. Previous studies have demonstrated that materials with high theoretical capacities, such as transition metal oxides, show promise as novel anode materials to boost LIB performance, thus holding significant application potential.
Among transition metal oxides, Co3O4 is considered a promising anode material due to its high theoretical specific capacity (890 mAh g−1) [3]. The lithium storage mechanism of Co3O4 involves a conversion reaction, where redox reactions occur during charge and discharge processes. However, this process is accompanied by significant volume expansion [4]. Additionally, Co3O4 as a p-type semiconductor exhibits extremely low electronic conductivity at room temperature [5]. This results in rapid capacity decay and poor cycling stability in LIBs, severely hindering the commercial application of Co3O4.To address these issues, various strategies have been employed to modulate the structure and morphology of Co3O4. For instance, rod-like Co3O4 derived from metal-organic frameworks [6], capsule-shaped porous Co3O4 nanomaterials [7], and polyacrylonitrile microfibers embedded with Co3O4 nanoparticles prepared via centrifugal spinning [8] have been developed. After these structural and morphological modifications, the electrochemical performance of Co3O4 as an anode material for LIBs has been greatly improved.
Another effective approach is to fabricate nanocomposites by combining Co3O4 with metal oxides that exhibit different lithium storage mechanisms. SnO2 offers a high theoretical specific capacity (approximately 780 mAh g−1) and relatively good electrical conductivity. Unlike the lithium storage mechanism of Co3O4, SnO2 follows a two-step reaction mechanism during lithiation. The first step involves reduction into Li2O and metallic Sn, followed by the alloying reaction between Li and Sn in the second step [9]. By compositing the two, a mutual buffering effect can be achieved during the charge and discharge processes of LIBs, resulting in a synergistic enhancement of lithium storage performance [10]. For instance, Kim et al. synthesized SnO2@Co3O4 hollow nanospheres with double shells via a template-based sol-gel coating method for application as anodes in LIBs. The unique hollow structure effectively alleviates volume variation while simultaneously accelerating lithium-ion transport and enhancing cycling stability. A high reversible discharge capacity of 962 mAh g−1 was maintained after 100 cycles at a current density of 100 mA g−1. SnO2@Co3O4 hollow nanospheres demonstrates a superior performance over the SnO2 electrode, which delivers a reversible discharge capacity of only 601 mAh g−1 [11]. Xing et al. prepared SnO2-Co3O4 core-shell nanoneedle arrays on copper foil. Leveraging the synergistic effect between SnO2 and Co3O4, along with the nanoneedle array structure, the electrode demonstrated improved capacity and mitigated volume expansion [12]. In another study, Chen et al. loaded SnO2 nanoparticles onto Co3O4 micro-flowers composed of porous nanosheets, obtaining a material labeled Co@Sn-MFs. Compared to pure Co3O4 microflowers, this composite showed enhanced specific capacity and improved electrical conductivity due to the synergistic interaction between its components and its unique spatial structure. The Co@Sn-MFs electrode delivered a discharge capacity of 767.5 mAh g−1 after 100 cycles at a current density of 200 mA g−1 [13].
Based on the strategy of using ethylene glycol as a ligand to synthesize spherical cobalt-glycolate precursors, which transform into yolk-shell Co3O4 microspheres after thermal decomposition [14,15], this study introduces 0.4 mmol of Sn2+ to modulate the precursor’s morphology. Specifically, the original spherical cobalt-glycolate evolves into rod-like hybrid Co3O4/SnO2 composites (denoted as Co3.8Sn1). The resulting Co3.8Sn1 nanorods feature a unique structure with abundant voids (to accommodate volume expansion during lithiation) and enhanced electrical conductivity (attributed to the in-situ formed SnO2 component). When employed as an anode material for LIBs, these nanorods demonstrate high specific capacity and excellent cycling/stability performance. This work aims to provide a feasible morphology-regulation strategy for developing high-performance transition metal oxide-based anodes, addressing the key bottlenecks of traditional anode materials in next-generation LIBs.

2. Materials and Methods

2.1. Preparation of Co3O4 Yolk-Shell Microspheres

4 mmol of Co(CH3COO)2·4H2O were dissolved in 100 mL of ethylene glycol (EG) and stirred at 50 °C for 30 min. The mixture was then refluxed at 170 °C for 2 h. After cooling to room temperature, the light purple precipitate was collected, washed with acetone, and dried under vacuum. Subsequently, the product was transferred to a muffle furnace (Hefei Kejing Materials Technology Co., Ltd., Hefei, China) and heated to 350 °C at a heating rate of 1 °C min−1 in air, followed by calcination for 2 h. The final product consisted of Co3O4 yolk-shell microspheres.

2.2. Preparation of Co3O4/SnO2 Hybrid Nanorods

A mixture of 4 mmol of Co(CH3COO)2·4H2O and 0.4 mmol of Sn(CH3COO)2, corresponding to a Co2+/Sn2+ molar ratio of 10:1, was dissolved in 100 mL of EG. The same synthetic procedure as in Section 2.1 was employed. The resulting mixture was stirred in an oil bath at 50 °C for 30 min, forming a light purple solution. Subsequently, the solution was heated to 170 °C and refluxed for 2 h, during which the color deepened to purple. After cooling to room temperature, the purple precipitate was collected by centrifugation, thoroughly washed with acetone, and dried under vacuum. The dried precursor was then transferred to a muffle furnace and calcined at 350 °C for 2 h in air. The final product was designated as Co3.8Sn1.
The same procedure was repeated using equimolar amounts of Co2+ and Sn2+ (i.e., a 1:1 molar ratio) to prepare the corresponding sample, which was denoted as Co3O4-SnO2.

2.3. Material Characterization Methods

The crystal structure and phase composition of the as-synthesized materials were determined by X-ray diffraction (XRD) on a D8 ADVANCE diffractometer (Beijing Bruker Technology Co., Ltd., Beijing, China) using Cu-Kα radiation (λ = 0.15418 nm). The microscopic morphology and structural features were examined using a JSM7600F (NEC Electronics Corporation, Tokyo, Japan) scanning electron microscope (SEM) and a Talos F200X G2 (Thermo Fisher Scientific Inc., Waltham, MA, USA) transmission electron microscope (TEM). The elemental composition and chemical states were analyzed by X-ray photoelectron spectroscopy (XPS) on an ESCALab250Xi spectrometer with an Al-Kα excitation source. The binding energy scale of all XPS spectra was calibrated by referencing the C 1s peak of adventitious carbon to 284.8 eV. The molar ratios of the elements in the samples were determined using an Agilent 5110 (Agilent Technology Co., Ltd, Santa Clara, CA, USA) inductively coupled plasma optical emission spectrometer (ICP-OES).

2.4. Electrochemical Measurements

The electrochemical performance of the Co3O4/SnO2 hybrid nanorods was evaluated by assembling CR2025-type coin cells in an argon-filled glove box (H2O and O2 levels < 0.1 ppm; MBraun, Stratham, NH, USA). The working electrode was prepared by thoroughly mixing 70 wt% active material, 20 wt% Ketjen black, and 10 wt% carboxymethyl cellulose (CMC) binder in deionized water to form a homogeneous slurry via magnetic stirring. This slurry was then uniformly coated onto a copper foil current collector and dried at 80 °C under vacuum for 12 h. The resulting electrode had an active material loading mass of approximately 1.0–1.2 mg cm−2. The electrolyte was 1.0 M LiPF6 dissolved in a mixture of ethylene carbonate (EC), ethyl methyl carbonate (EMC), and dimethyl carbonate (DMC) (1:1:1 by volume). A lithium metal foil was used as the counter/reference electrode, and a Celgard 2400 membrane (Suzhou Sinero Technology Co., Ltd., Suzhou, China) was employed as the separator. Galvanostatic charge-discharge tests were performed on a Land CT2001A battery test system within a voltage window of 0.001–3.0 V (vs. Li/Li+) at various current densities to assess the specific capacity, cycling stability, and rate capability. Cyclic voltammetry (CV) was conducted on a VSP electrochemical workstation at a scan rate of 0.1 mV s−1. Electrochemical impedance spectroscopy (EIS) measurements were carried out over a frequency range from 100 kHz to 100 mHz to evaluate the electrode’s interfacial and charge-transfer resistance.

3. Results

3.1. Structural Characterization and Phase Analysis of Materials

The synthetic route for Co3.8Sn1 is illustrated in Scheme 1. A cobalt-tin glycolate rod-like precursor (CoSn-gly) was first synthesized via a solvothermal reaction in EG from Co(CH3COO)2·4H2O and Sn(CH3COO)2. Subsequent calcination of this precursor at 350 °C in air yielded the final material. ICP-OES measurements confirmed a Co:Sn molar ratio of 3.8:1 in the final product, consequently, the product was designated as Co3.8Sn1. The reduced molar amount of Co in the product is attributed to the loss of a portion of Co2+ ions that did not participate in the reaction.
Figure S1 presents the SEM images of the precursors for Co3O4 and Co3.8Sn1. As shown in Figure S1a,b, the Co-gly precursor consists of microspheres with a uniform diameter of approximately 300 nm. In contrast, the introduction of Sn2+ significantly alters the morphology of the precursor. Figure S1c,d reveal that the presence of Sn promotes the formation of a composite structure comprising both nanorods and nanospheres, indicating a tendency for rod-like growth. These precursors were subsequently calcined at 350 °C for 2 h. The resulting products are shown in Figure S2. After calcination, the Co-gly microspheres transform into yolk-shell Co3O4 microspheres. Meanwhile, the Co3.8Sn1 material maintains its rod-like morphology. Despite the persistence of numerous nanospheres, the rod-like morphology of the final Co3.8Sn1 material was largely retained.
The microstructure of the samples was further characterized by TEM. Figure S3 presents TEM images of the Co3O4 sample, revealing that the precursor transforms into yolk-shell microspheres after calcination (Figure S3b). The HRTEM image in Figure S3c shows lattice fringes with an interplanar spacing of 0.243 nm, corresponding to the (311) plane of Co3O4. Elemental mapping images (Figure S3d–f) confirm the uniform distribution of Co and O throughout the microspheres. As shown in Figure 1a,b, the Co3.8Sn1 composite consists of numerous nanoparticles assembled into rod-like architectures with substantial internal voids, a structure conducive to rapid ion transport. The HRTEM image in Figure 1c displays two distinct sets of lattice fringes: spacings of 0.237 nm and 0.341 nm, which are assigned to the (311) plane of Co3O4 and the (110) plane of SnO2, respectively. This, combined with the elemental mapping results in Figure 1d–f which show a homogeneous distribution of Co, Sn, and O, indicates that the Co3.8Sn1 material is composed of Co3O4 nanoparticles encapsulated by a SnO2 outer layer.
The crystal structures and phase compositions of the synthesized materials were characterized by XRD, as shown in Figure 2a. For Co3O4, the distinct diffraction peaks at 31.2°, 36.8°, 44.8°, 59.3°, and 65.2° are indexed to the (220), (311), (400), (511), and (440) planes, respectively, of a face-centered cubic spinel structure (JCPDS No. 42-1467). The XRD pattern of the Co3O4-SnO2 sample shows three prominent peaks at 26.6°, 33.8°, and 51.7°, which correspond to the (110), (101), and (211) planes of tetragonal SnO2 (JCPDS No. 41-1445), confirming its successful formation. In the case of the Co3.8Sn1 composite, diffraction peaks from both Co3O4 and SnO2 are present. However, the characteristic peaks of SnO2 are of relatively low intensity due to its lower concentration in the composite. The XRD pattern further confirms that the Co3.8Sn1 is a composite material consisting of SnO2 and Co3O4.
The elemental composition and chemical states of the synthesized samples were analyzed by XPS. As shown in Figure 2b, the high-resolution O 1s spectrum was deconvoluted into three peaks at binding energies of 529.5 eV, 531.2 eV, and 533.4 eV, which are assigned to lattice oxygen (Olat), adsorbed oxygen species (Oads), and oxygen from surface-adsorbed water or hydroxyl groups (Ow), respectively [16]. The introduction of Sn2+ caused a positive shift in the O 1s peak, indicating a decrease in the electron density around oxygen atoms due to a strong electronic interaction with SnO2. Figure 2c displays the Co 2p spectrum, characterized by two spin-orbit doublets. The main peaks at 779.4 eV and 794.5 eV are attributed to Co2p3/2 and Co 2p1/2, respectively. Each doublet could be further fitted with two components: the peaks at 779.4 eV and 794.5 eV correspond to Co2+, while those at 781.2 eV and 796.3 eV are indicative of Co3+, and the spectrum is also accompanied by satellite features confirms [17]. These results demonstrate that cobalt in Co3.8Sn1 exists in a mixed-valence state, with the coexistence of Co2+ and Co3+ [18]. Notably, the Co 2p peaks in the SnO2-containing composite Co3.8Sn1 shifts to higher binding energies compared to pure Co3O4, suggesting the formation of a p-n heterojunction at the interface between Co3O4 and SnO2, which modifies the local electron density [19]. The Sn 3d spectrum of Co3.8Sn1 is presented in Figure 2d. The peaks located at 486.3 eV and 494.7 eV are assigned to Sn 3d5/2 and Sn 3d3/2, respectively. The spin-orbit splitting of 8.4 eV is consistent with the standard value for Sn4+ in SnO2, confirming the successful incorporation of tin oxide [20].

3.2. Electrochemical Properties

The electrochemical reactions of Co3O4 and Co3.8Sn1 were investigated by CV, as shown in Figure 3a,b. A distinct difference is observed between the first-cycle CV curves and the subsequent cycles for both materials, which is attributed to irreversible reactions and structural changes during the initial cycle, leading to the formation of a solid electrolyte interphase (SEI) layer [21]. For Co3O4 (Figure 3a), the first cathodic scan exhibited a reduction peak at approximately 0.82 V, attributed to the reduction of Co3O4 to metallic Co and the formation of Li2O [22]. The subsequent anodic scan showed an oxidation peak at around 2.10 V, corresponding to the reversible oxidation of Co to Co3O4. For Co3.8Sn1 (Figure 3b), in its first cathodic scan, the peak at 0.03 V is assigned to the alloying reaction of Li+ with Sn to form LixSn alloys [23], while the prominent peak at 0.69 V corresponds to the further reduction of Co3O4 and SnO2 to metallic Co and Sn. Compared to the reduction peak of pure Co3O4 at 0.82 V, this peak is attributed to the formation of the Co3O4/SnO2 heterojunction in the Co3.8Sn1 composite, which modifies the electronic structure and consequently enhances the electrochemical performance of the material. In the anodic scan, the peaks at 0.56 V, 1.07 V, and 2.11 V are related to the de-alloying of LixSn (x represents the stoichiometric coefficient of Li in the LixSn alloy), the oxidation of Sn to SnO2, and the oxidation of Co to Co3O4, respectively [24]. The curves for the second and third cycles overlapped well, demonstrating excellent cycling stability. The diverse lithium storage mechanisms and highly reversible CV curves suggest that the Co3.8Sn1 electrode possesses high structural stability, low polarization, and the potential for high specific capacity. The relevant reaction equations are as follows:
Co3O4 + 8Li+ + 8e ↔ 3Co + 4Li2O,
SnO2 + 4Li+ + 4e ↔ Sn + 2Li2O,
Sn + xLi+ + xe ↔ LixSn (0 ≤ x ≤ 4.4).
The EIS results for Co3O4 and Co3.8Sn1 are presented in Figure 3c. The Nyquist plots for both materials consist of a depressed semicircle in the high-to-medium frequency region and an inclined line in the low-frequency region. The intercept on the Z’-axis at high frequency corresponds to the ohmic resistance (Rs) of the electrolyte. The diameter of the semicircle reflects the charge-transfer resistance (Rct) at the electrode/electrolyte interface, while the sloping line is associated with the diffusion resistance Zw of lithium ions in the electrolyte. It is evident that the Co3.8Sn1 electrode exhibits a significantly smaller semicircular diameter compared to the Co3O4 electrode, indicating a much lower Rct. This reduction can be attributed to the altered microstructure induced by Sn2+ incorporation, which increases the electrode/electrolyte contact area and facilitates ion diffusion. The synergistic effect between Co and Sn species in the Co3.8Sn1 composite accelerates charge transfer kinetics, thereby leading to improved electrochemical performance.
The rate capability of the Co3.8Sn1 anode was evaluated at various current densities, as shown in Figure 4a. The electrode delivered high average specific capacities of 2084.84, 1996.96, 1623.31, 1235.13, and 888.46 mAh g−1 at current densities of 0.1, 0.2, 0.5, 1 and 2 A g−1, respectively. Notably, even at a high current density of 2 A g−1, the capacity of Co3.8Sn1 significantly surpassed that of the pure Co3O4 anode. Furthermore, when the current density was returned to 0.2 A g−1, the capacity recovered to 1365.22 mAh g−1, demonstrating excellent reversibility and robust lithium storage performance.
The cycling performance of the Co3.8Sn1 and Co3O4 electrodes was evaluated at a current density of 0.2 A g−1, as shown in Figure 4b. The Co3.8Sn1 anode demonstrated a high initial specific capacity and exceptional stability, maintaining a maximum specific capacity of 1402.7 mAh g−1 after 150 cycles with a Coulombic efficiency of approximately 98%. In contrast, the Co3O4 anode not only started with a lower initial capacity but also suffered from a rapid capacity decay over 100 cycles. The incorporation of SnO2 effectively modifies the structure of Co3O4, mitigating the agglomeration of nanoparticle and enhancing electrical conductivity. These synergistic effects collectively contribute to the significantly improved cycling stability and enhanced capacity of the Co3.8Sn1 composite. In the long-term cycling tests conducted at various current densities, the specific capacity of the Co3.8Sn1 electrode consistently demonstrates a trend of initial decrease followed by gradual increase. This phenomenon can be attributed to the presence of transition metal Co, which may catalyze electrolyte decomposition and the formation of a gel-like film, thereby contributing additional pseudocapacitive lithium storage [9]. The galvanostatic charge-discharge profiles of the Co3.8Sn1 and Co3O4 electrodes for the 1st, 2nd, 5th, and 10th cycles are presented in Figure 4c. The Co3.8Sn1 electrode exhibited an initial discharge capacity of 2645.8 mAh g−1 and a charge capacity of 1820.4 mAh g−1, yielding a first-cycle Coulombic efficiency of 68.8%. The initial irreversible capacity loss is primarily associated with SEI formation. The highly overlapping voltage profiles from the 2nd to the 10th cycle indicate excellent cycling reversibility thereafter. In the initial discharge curve, the voltage plateau observed between 0.4–1.0 V can be attributed to the conversion of SnO2 to Sn, the reduction of Co3O4 to Co. The subsequent plateau in the range of 0.02–0.20 V corresponds to the alloying reaction between Sn and Li+ to form LixSn alloys. During the charging process, three distinct voltage plateaus emerge: the first plateau (0.4–0.7 V) is associated with the de-alloying of LixSn, while the subsequent two plateaus (0.8–1.2 V and 1.8–2.2 V) correspond to the stepwise oxidation of Sn to SnO2 and Co to Co3O4, respectively [25].
The long-term cycling performance of the Co3.8Sn1 electrode was further evaluated at a high current density of 1 A g−1, as shown in Figure 4d. Following activation at 0.2 A g−1, the cells were cycled at 1 A g−1. The corresponding galvanostatic charge-discharge profiles for Co3O4 and Co3.8Sn1 are presented in Figure 4e,f, respectively. The Co3.8Sn1 electrode delivered a superior initial specific capacity and demonstrated exceptional stability, retaining a capacity of 496.8 mAh g−1 after 500 cycles. In stark contrast, the Co3O4 electrode retained only 139.8 mAh g−1 after the same number of cycles, underscoring the significant performance enhancement afforded by the Co3.8Sn1 composite.
As shown in Figure 4g, at a current density of 2 A g−1, the Co3.8Sn1 electrode delivered a specific capacity of 427.9 mAh g−1 after 500 cycles, significantly outperforming the mere 84.5 mAh g−1 retained by the Co3O4 electrode. The Coulombic efficiency of Co3.8Sn1 remained at approximately 98.7%. Figure 4h,i present the corresponding galvanostatic charge-discharge profiles of Co3O4 and Co3.8Sn1 at 2 A g−1, respectively. With increasing cycle numbers, Co3.8Sn1 exhibited a markedly slower capacity decay and more stable voltage profiles compared to Co3O4. These results indicate that Sn doping effectively enhances the cycling stability and reversibility of the material, enabling Co3.8Sn1 to maintain a high capacity retention rate and superior electrochemical performance during long-term cycling at high current densities.

4. Conclusions

In summary, Co3O4/SnO2 hybrid nanorods were successfully synthesized via a solvothermal reaction coupled with thermal decomposition. The introduction of Sn2+ induced the transformation of spherical cobalt-glycolate precursors into rod-like architectures with abundant voids, which provided buffer space to accommodate volume expansion during lithiation/delithiation cycles. Meanwhile, the in-situ formed SnO2 component significantly enhanced the electrical conductivity of the hybrid system, addressing the poor electron transport issue of pristine Co3O4. Furthermore, the synergistic interaction between Co and Sn species optimized the lithium storage behavior, strengthening the adsorption and diffusion of Li+ while promoting reversible redox reactions between multi-valent metal species. When evaluated as an anode for LIBs, the Co3.8Sn1 electrode exhibited remarkable electrochemical performance, which delivered high reversible specific capacities of 496.8 and 427.9 mAh g−1 after 500 cycles at high current densities of 1 and 2 A g−1, respectively. These results confirm that Sn-doping-mediated microstructure engineering is an effective strategy to overcome the inherent bottlenecks of volume expansion and low conductivity of transition metal oxide anodes. More importantly, the Co3.8Sn1 hybrid nanorods, with their balanced structural stability and electrochemical activity, emerge as a promising candidate material for next-generation high-performance LIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/physchem5040054/s1, Figure S1: SEM images of (a,b) Co-gly, (c,d) Co3.8Sn1-gly; Figure S2: SEM images of (a) Co3O4, (b) Co3.8Sn1; Figure S3: (a,b) TEM images of Co3O4; (c) HRTEM image of Co3O4; (d–f) elemental mapping images of Co3O4.

Author Contributions

Conceptualization, Q.Z. and D.L.; Software, Q.Z.; Validation, Q.Z., J.Z. and L.F.; Formal analysis, Q.Z., J.Z., L.F. and Y.Z.; Investigation, Q.Z.; Data curation, J.Z.; Writing—original draft, Q.Z. and J.Z.; Writing—review & editing, D.L.; Supervision, D.L. and Y.Z.; Project administration, Y.Z.; Funding acquisition, D.L. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 52432004, U22A20141, U23A20575, 52472183, and 524B2020).

Data Availability Statement

All data are contained in this article or the Supplementary Materials.

Acknowledgments

The authors also acknowledge the facilities, and the scientific and technical assistance of the Analysis & Testing Center and the High-Performance Computing Center of Beihang University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ji, G.; He, L.; Wu, T.; Cui, G. The design of fast charging strategy for lithium-ion batteries and intelligent application: A comprehensive review. Appl. Energy 2025, 377, 124538. [Google Scholar] [CrossRef]
  2. Liu, P.; Wang, H.; Huang, T.; Li, L.; Xiong, W.; Huang, S.; Ren, X.; Ouyang, X.; Hu, J.; Zhang, Q.; et al. Cost-effective natural graphite reengineering technology for lithium-ion batteries. Chin. Chem. Lett. 2024, 35, 108330. [Google Scholar] [CrossRef]
  3. Wang, L.; Yuan, Y.F.; Zheng, Y.Q.; Zhang, X.T.; Yin, S.M.; Guo, S.Y. Capsule-like Co3O4 nanocage@ Co3O4 nanoframework/TiO2 nodes as anode material for lithium-ion batteries. Mater. Lett. 2019, 253, 5–8. [Google Scholar] [CrossRef]
  4. Gu, D.; Li, W.; Wang, F.; Bongard, H.; Spliethoff, B.; Schmidt, W.; Weidenthaler, C.; Xia, Y.; Zhao, D.; Schith, F. Controllable synthesis of mesoporous peapod-like Co3O4@carbon nanotube arrays for high-performance lithium-ion batteries. Angew. Chem. Int. Ed. 2015, 54, 7060–7064. [Google Scholar] [CrossRef]
  5. Park, J.; Shen, X.; Wang, G. Solvothermal synthesis and gas-sensing performance of Co3O4 hollow nanospheres. Sens. Actuators B Chem. 2009, 136, 494–498. [Google Scholar] [CrossRef]
  6. Chen, J.; Mu, X.; Du, M.; Lou, Y. Porous rod-shaped Co3O4 derived from Co-MOF-74 as high-performance anode materials for lithium ion batteries. Inorg. Chem. Commun. 2017, 84, 241–245. [Google Scholar] [CrossRef]
  7. Chen, J.; Zhao, N.; Tang, C.; Chen, Y. Facile synthesis of Co3O4 with porous capsule-shaped structure and its lithium storage properties as anode materials for Li-ion batteries. Ceram. Int. 2024, 50, 31567–31575. [Google Scholar] [CrossRef]
  8. Ayala, J.; Ramirez, D.; Myers, J.C.; Lodge, T.P.; Parsons, J.; Alcoutlabi, M. Performance and morphology of centrifugally spun Co3O4/C composite fibers for anode materials in lithium-ion batteries. J. Mater. Sci. 2021, 56, 16010–16027. [Google Scholar] [CrossRef]
  9. Hassan, M.F.; Guo, Z.P.; Du, G.D.; Wexler, D.; Liu, H.K. Preparation of tin nanocomposite as anode material by molten salts method and its application in lithium ion batteries. Phys. Status Solidi A 2009, 11, 2546–2550. [Google Scholar]
  10. Wang, J.; Wang, H.; Yao, T.; Liu, T.; Tian, Y.; Li, C.; Li, F.; Meng, L.; Cheng, Y. Porous N-doped carbon nanoflakes supported hybridized SnO2/Co3O4 nanocomposites as high-performance anode for lithium-ion batteries. J. Colloid Interface Sci. 2020, 560, 546–554. [Google Scholar] [CrossRef]
  11. Kim, W.S.; Hwa, Y.; Kim, H.C.; Choi, J.H.; Sohn, H.J.; Hong, S.H. SnO2@Co3O4 hollow nano-spheres for a Li-ion battery anode with extraordinary performance. Nano Res. 2014, 7, 1128–1136. [Google Scholar] [CrossRef]
  12. Xing, L.; Zhao, Y.; Zhao, J.; Nie, Y.; Deng, P.; Wang, Q.; Xue, X. Facile synthesis and lithium storage performance of SnO2-Co3O4 core-shell nanoneedle arrays on copper foil. J. Alloys Compd. 2014, 586, 28–33. [Google Scholar] [CrossRef]
  13. Chen, J.; Zhao, N.; Shi, D.; Wang, P.; Chen, Y. Superior lithium storage performance of SnO2-modified Co3O4 microflowers self-assembled with porous nanosheets as anode materials in Li-ion batteries. J. Alloys Compd. 2023, 967, 171751. [Google Scholar] [CrossRef]
  14. Feng, X.; Liu, D.; Li, W.; Jin, X.; Zhang, Z.; Zhang, Y. Catalytic activity boost of CeO2/Co3O4 hollow derived from CeCo-glycolate via yolk-shell structural evolution. Inorg. Chem. Front. 2020, 7, 421–426. [Google Scholar] [CrossRef]
  15. Zhu, J.; Fu, L.; Zuo, X.; Liu, D.; Zhang, Y. Co3O4/CuO Hybrid hollow microspheres as long-cycle-life lithium-ion battery anode. Chem. Asian J. 2025, 20, e202401184. [Google Scholar] [CrossRef] [PubMed]
  16. Bae, J.; Shin, D.; Jeong, H.; Kim, B.; Han, J.; Lee, H. Highly water-resistant La-doped Co3O4 catalyst for CO oxidation. ACS Catal. 2019, 9, 10093–10100. [Google Scholar] [CrossRef]
  17. Yang, L.; Wang, Y.; Wang, J.; Zheng, Y.; Ang, E.H.; Hu, Y.; Zhu, J. Imidazole-intercalated cobalt hydroxide enabling the Li+ desolvation/diffusion reaction and flame retardant catalytic dynamics for lithium ion batteries. Angew. Chem. Int. Ed. 2024, 63, e202402827. [Google Scholar] [CrossRef]
  18. Pawar, S.M.; Pawar, B.S.; Hou, B.; Ahmed, A.T.A.; Chavan, H.S.; Jo, Y.; Cho, S.; Kim, J.; Seo, J.; Cha, S.N.; et al. Facile electrodeposition of high-density CuCo2O4 nanosheets as a high-performance Li-ion battery anode material. J. Ind. Eng. Chem. 2019, 69, 13–17. [Google Scholar] [CrossRef]
  19. Li, X.; Sun, X.; Gao, Z.; Hu, X.; Ling, R.; Cai, S.; Zheng, C.; Hu, W. Highly reversible and fast sodium storage boosted by improved interfacial and surface charge transfer derived from the synergistic effect of heterostructures and pseudocapacitance in SnO2-based anodes. Nanoscale 2018, 10, 2301. [Google Scholar] [CrossRef]
  20. Balachandran, S.; Selvam, K.; Babu, B.; Swaminathan, M. The simple hydrothermal synthesis of Ag-ZnO-SnO2 nanochain and its multiple applications. Dalton Trans. 2013, 42, 16365–16374. [Google Scholar] [CrossRef]
  21. Guo, S.; Liu, J.; Zhang, Q.; Jin, Y.; Wang, H. Metal organic framework derived Co3O4/Co@N-C composite as high-performance anode material for lithium-ion batteries. J. Alloys Compd. 2021, 855, 157538. [Google Scholar] [CrossRef]
  22. Park, S.; Kim, J.; Kim, J.; Kang, Y. Metal-organic framework-templated hollow Co3O4 nanosphere aggregate/N-doped graphitic carbon composite powders showing excellent lithium-ion storage performances. Mater. Charact. 2017, 132, 320–329. [Google Scholar] [CrossRef]
  23. Qi, Y.; Zhang, H.; Du, N.; Zhai, C.; Yang, D. Synthesis of Co3O4@ SnO2@C core-shell nanorods with superior reversible lithium-ion storage. RSC Adv. 2012, 2, 9511–9516. [Google Scholar] [CrossRef]
  24. Huang, J.; Ma, Y.; Xie, Q.; Zheng, H.; Yang, J.; Wang, L.; Peng, D. 3D Graphene encapsulated hollow CoSnO3 nanoboxes as a high initial coulombic efficiency and lithium storage capacity anode. Small 2018, 14, 1703513. [Google Scholar] [CrossRef]
  25. Wang, Y.; Huang, Z.X.; Shi, Y.; Wong, J.I.; Ding, M.; Yang, H.Y. Designed hybrid nanostructure with catalytic effect: Beyond the theoretical capacity of SnO2 anode material for lithium ion batteries. Sci. Rep. 2015, 5, 9164. [Google Scholar] [CrossRef]
Scheme 1. Schematic synthesis of Co3.8Sn1.
Scheme 1. Schematic synthesis of Co3.8Sn1.
Physchem 05 00054 sch001
Figure 1. (a,b) TEM images of Co3.8Sn1; (c) HRTEM image of Co3.8Sn1; (df) elemental mapping images of Co3.8Sn1.
Figure 1. (a,b) TEM images of Co3.8Sn1; (c) HRTEM image of Co3.8Sn1; (df) elemental mapping images of Co3.8Sn1.
Physchem 05 00054 g001
Figure 2. (a) XRD of Co3O4, Co3.8Sn1 and Co3O4-SnO2; XPS spectra of (b) O 1s, (c) Co 2p, (d) Sn 3d.
Figure 2. (a) XRD of Co3O4, Co3.8Sn1 and Co3O4-SnO2; XPS spectra of (b) O 1s, (c) Co 2p, (d) Sn 3d.
Physchem 05 00054 g002
Figure 3. First three CV curves of (a) Co3O4 and (b) Co3.8Sn1. (c) EIS curves of Co3.8Sn1 and Co3O4.
Figure 3. First three CV curves of (a) Co3O4 and (b) Co3.8Sn1. (c) EIS curves of Co3.8Sn1 and Co3O4.
Physchem 05 00054 g003
Figure 4. (a) Rate capability and (b) cycling performance at a current density of 0.2 A g−1 of Co3.8Sn1; (c) Galvanostatic discharge-charge profiles of Co3.8Sn1 at 0.2 A g−1; (d) cycling performance at a current density of 1 A g−1 of Co3O4 and Co3.8Sn1. Galvanostatic discharge-charge profiles of 1 A g−1 of (e) Co3O4, (f) Co3.8Sn1. (g) Cycling performance at a current density of 2 A g−1 of Co3O4 and Co3.8Sn1. Galvanostatic discharge-charge profiles of 2 A g−1 of (h) Co3O4, (i) Co3.8Sn1.
Figure 4. (a) Rate capability and (b) cycling performance at a current density of 0.2 A g−1 of Co3.8Sn1; (c) Galvanostatic discharge-charge profiles of Co3.8Sn1 at 0.2 A g−1; (d) cycling performance at a current density of 1 A g−1 of Co3O4 and Co3.8Sn1. Galvanostatic discharge-charge profiles of 1 A g−1 of (e) Co3O4, (f) Co3.8Sn1. (g) Cycling performance at a current density of 2 A g−1 of Co3O4 and Co3.8Sn1. Galvanostatic discharge-charge profiles of 2 A g−1 of (h) Co3O4, (i) Co3.8Sn1.
Physchem 05 00054 g004
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

Zhang, Q.; Zhu, J.; Fu, L.; Liu, D.; Zhang, Y. Co3O4/SnO2 Hybrid Nanorods as High-Capacity Anodes for Lithium-Ion Batteries. Physchem 2025, 5, 54. https://doi.org/10.3390/physchem5040054

AMA Style

Zhang Q, Zhu J, Fu L, Liu D, Zhang Y. Co3O4/SnO2 Hybrid Nanorods as High-Capacity Anodes for Lithium-Ion Batteries. Physchem. 2025; 5(4):54. https://doi.org/10.3390/physchem5040054

Chicago/Turabian Style

Zhang, Qiyao, Jingchao Zhu, Lichao Fu, Dapeng Liu, and Yu Zhang. 2025. "Co3O4/SnO2 Hybrid Nanorods as High-Capacity Anodes for Lithium-Ion Batteries" Physchem 5, no. 4: 54. https://doi.org/10.3390/physchem5040054

APA Style

Zhang, Q., Zhu, J., Fu, L., Liu, D., & Zhang, Y. (2025). Co3O4/SnO2 Hybrid Nanorods as High-Capacity Anodes for Lithium-Ion Batteries. Physchem, 5(4), 54. https://doi.org/10.3390/physchem5040054

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop