Next Article in Journal
Optical Characterization of Thin Films by Surface Plasmon Resonance Spectroscopy Using an Acousto-Optic Tunable Filter
Previous Article in Journal
Numerical Studies of the Impact of Electromagnetic Field of Radiation on Valine
Previous Article in Special Issue
Lithium-Ion Batteries under Low-Temperature Environment: Challenges and Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon-Coated CuNb13O33 as A New Anode Material for Lithium Storage

Institute of Materials for Energy and Environment, School of Materials Science and Engineering, Qingdao University, Qingdao 266071, China
*
Authors to whom correspondence should be addressed.
Submission received: 7 January 2023 / Revised: 28 January 2023 / Accepted: 31 January 2023 / Published: 22 February 2023
(This article belongs to the Special Issue Preparation, Characterization and Mechanism of Electrode Materials)

Abstract

:
Niobates are very promising anode materials for Li+-storage rooted in their good safety and high capacities. However, the exploration of niobate anode materials is still insufficient. In this work, we explore ~1 wt% carbon-coated CuNb13O33 microparticles (C-CuNb13O33) with a stable shear ReO3 structure as a new anode material to store Li+. C-CuNb13O33 delivers a safe operation potential (~1.54 V), high reversible capacity of 244 mAh g−1, and high initial-cycle Coulombic efficiency of 90.4% at 0.1C. Its fast Li+ transport is systematically confirmed through galvanostatic intermittent titration technique and cyclic voltammetry, which reveal an ultra-high average Li+ diffusion coefficient (~5 × 10–11 cm2 s−1), significantly contributing to its excellent rate capability with capacity retention of 69.4%/59.9% at 10C/20C relative to 0.5C. An in-situ XRD test is performed to analyze crystal-structural evolutions of C-CuNb13O33 during lithiation/delithiation, demonstrating its intercalation-type Li+-storage mechanism with small unit-cell-volume variations, which results in its capacity retention of 86.2%/92.3% at 10C/20C after 3000 cycles. These comprehensively good electrochemical properties indicate that C-CuNb13O33 is a practical anode material for high-performance energy-storage applications.

1. Introduction

Compared with the traditional nickel-iron, nickel-metal, and lead-acid batteries, lithium-ion batteries (LIBs) are much more popular power sources for consumable electronics and electric vehicles due to the higher energy density, higher power density, and low self-discharge [1,2,3,4,5,6,7,8]. To satisfy the fast development need from electric vehicles, the exploration of higher-performance electrode materials is highly necessary [9]. At present, the most popular anode material is based on intercalation-type graphite due to its low cost and high practical capacity of 330–360 mAh g−1 [10,11]. However, it suffers a safety issue of lithium-dendrite formation when fast discharged/charged at its low potential plateau (<0.1 V) [12]. Intercalation-type Li4Ti5O12, which is the second most popular anode material, delivers a safe operation potential (~1.55 V), successfully avoiding the above safety issue [13]. Unfortunately, its insufficient practical capacity (~170 mAh g−1) limits its wide applications [14]. Hence, it is desirable to develop new anode materials with both large practical capacities and safe performance [15].
Recently, niobates with a high niobium valance of +5 have been served as high-performance anode materials [16,17,18,19]. The active Nb4+/Nb5+ and Nb3+/Nb4+ redox couples result in larger practical capacities than Li4Ti5O12 and safer operation potentials than graphite. Additionally, the large anion/cation ratios in niobates result in open crystal structures, enabling fast Li+ transport [20]. So far, a few niobates have been extensively studied as anode materials for Li+ storage. For instance, TiNb2O7 was firstly used as an anode material by Goodenough et al. in 2011, which showed a high reversible capacity of 285 mAh g−1 (0.1C) [21]. The Ti2Nb10O29 microparticles were synthesized by Wu et al., which exhibited a high reversible capacity of 247 mAh g−1 (0.1C) and high rate capability (130 mAh g−1 at 10C) [22]. The Ni2Nb34O87 microparticles were reported by Lv et al., which delivered high rate capability (capacity retention of 57.5% at 10C relative to 0.5C at 25 °C, 64.0% at 2C relative to 0.5C at −10 °C, and 65.3% at 10C relative to 0.5C at 60 °C) [23]. The Cu2Nb34O87 microparticles were synthesized by Yang et al., which presented a high reversible capacity of 343 mAh g−1 (0.1C), high rate capability (184 mAh g−1 at 10C) and good cyclability (88.5% capacity retention after 1000 cycles at 10C) [19]. However, the challenge remains to explore more niobate anode materials for LIBs.
Here, we explore CuNb13O33 with a stable ReO3 crystal structure as a new niobate anode material. To improve the electrical conduction among the CuNb13O33 microparticles, a carbon-coating strategy is employed. The carbon-coated CuNb13O33 (C-CuNb13O33) exhibits fast Li+ diffusivity with 5.01 × 10–11 cm2 s−1, significantly contributing to its outstanding rate capability (capacity percentage of 69.4%/59.9% at 10C/20C relative to 0.5C). Its maximum unit-cell-volume change is only 7.61%, leading to its superior cyclability (86.2%/92.3% capacity retention after 3000 cycles at 10C/20C). Furthermore, C-CuNb13O33 shows a high practical capacity (244 mAh g−1), high initial-cycle Coulombic efficiency (90.4%), and safe operation potential (~1.54 V) at 0.1C. These comprehensively good electrochemical properties of C-CuNb13O33 indicates that it is an ideal anode material of the LIBs for electric vehicles.

2. Materials and Methods

CuNb13O33 was prepared using solid-state reaction. 1 mmol Cu2O (Macklin, 97.0%) and 13 mmol Nb2O5 (AD 8758, Companhia Brasileira de Metalurgia e Mineração (CBMM)) were milled in a high-energy ball miller (SPEX 8000M, Metuchen, NJ, USA) for 1 h. After sintering the milled mixture at 900 °C for 4 h in Ar, CuNb13O33 microparticles were obtained. Then, 0.56 mmol CuNb13O33 and 0.35 mmol lactose (Macklin, 98.0%) were mixed in water under stirring at 80 °C until the mixture was fully dried. The obtained solid was heated at 700 °C for 2 h in Ar for carbonization, forming carbon-coated CuNb13O33 (C-CuNb13O33).
The X-ray diffraction (XRD) data of C-CuNb13O33 was recorded by an X-ray diffractometer (Rigaku Ultima IV, Tokyo, Japan) with Cu-Kα radiation. The in-situ XRD test was performed by a specially-designed electrochemical cell module (Scistar LIB-XRD) equipped with a low-X-ray-absorption Be window [24]. The lattice-paraments were determined by Rietveld refinements of the powder and in-situ XRD patterns conducted on the General Structure Analysis System (GSAS) software (Revision 1253) with the EXPGUI interface [25]. The particle size, microstructure, and selected area electron diffraction (SAED) pattern were recorded by a field emission scanning electron microscopy (FESEM, JEOL JSM-7800F, Tokyo, Japan) equipped with energy-dispersive X-ray spectroscopy (EDX, OXFORD X-Max, Oxford, UK), and high-resolution transmission electron microscopy (HRTEM, JEOL JEM-2100, Tokyo, Japan). The weight percentage of carbon in C-CuNb13O33 was determined by a thermogravimetry analyzer (TGA, TGA 2, Mettler Toledo, Zurich, Switzerland). The elemental valence states of C-CuNb13O33 were examined by X-ray photoelectron spectroscopy (XPS, PHI5000 Versaprobe III, Mausaki, Japan).
The CR2032-type coin cells, which were fabricated in an Ar-filled glove box, were used to assess the electrochemical properties of C-CuNb13O33. C-CuNb13O33 (70 wt%), Super-P® conductive carbon (20 wt%), and polyvinylidene fluoride (10 wt%) were mixed in N-methylpyrrolidone. After stirred for 8 h, the uniform slurry was casted on a Cu current collector, which was dried in a vacuum oven at 110 °C for 10 h. It was cut to circular electrodes with a diameter of 12 mm, obtaining the electrodes with active-material loadings of ~1.0 mg cm−2. The electrolyte was consisted of 1 M LiPF6 in an ethylene carbonate/diethylene carbonate/dimethyl carbonate mixed solvent (1:1:1 in volume). Glass fibers (Whatman GF/D-1823) were used as separators. The Li-metal foils were served as both counter and reference electrodes. Galvanostatic charge–discharge (GCD) tests were conducted on an automatic battery testing system (CT-3008, Neware, Shenzhen, China) at room temperature. An electrochemical workstation (Gamry Interface 1010E, Philadelphia, PA, USA) was employed to record the cyclic voltammogram (CV) profiles. All the electrochemical properties were examined within 1.0–3.0 V.

3. Results and Discussion

3.1. Physico–Chemical Characterizations

The Rietveld-refined XRD data (Figure 1a) indicates that CuNb13O33 has a monoclinic lattice with C2/m space group. The lattice parameters of CuNb13O33 are Rietveld-refined to be a = 22.49305(113) Å, b = 3.82579(17) Å, c = 15.40986(76) Å, β = 91.336(4)°, and V = 1325.714(140) Å3 (Table S1), and its fraction atomic parameters are listed in Table S2 (Supplementary Materials). The crystal structure of CuNb13O33 (Figure 1b) is constructed by NbO6 octahedra spreading on different shear ac-planes [26]. The NbO6 octahedra are connected by sharing corners and edges in each layer. However, the arrangement is interrupted by Cu+ in a regular way. Each Cu+ is connected to two O2-ions and occupies the 2c site. Moreover, the interlayers share octahedron edges. This special octahedron connection results in a stable A–B–A layered structure. The resulting tunnels in this crystal structure are beneficial for Li+ transport and storage. Figure S1 shows the XPS spectra of Nb and Cu elements in C-CuNb13O33. The Nb-3d spectrum consists of a Nb-3d5/2 and Nb-3d3/2 doublet respectively at 207.8 and 210.5 eV (Figure S1a) [24], and the Cu-2p spectrum comprises a Cu-2p3/2 and Cu-2p1/2 doublet respectively at 932.5 and 952.5 eV (Figure S1b) [27], indicating that the valences states of Nb and Cu elements are respectively +5 and +1, as expected.
The FESEM image (Figure 1c) of C-CuNb13O33 exhibits that its particle size distributes in a range of 1–5 μm. Its HRTEM image (Figure 1d) indicates that the d-spacing of 0.377 nm is indexed to the (110) crystallographic plane and that the carbon layer has a thickness of ~3 nm. The carbon content of C-CuNb13O33 is calculated to be ~1 wt% from its TGA result (Figure S2). The regular diffraction spots from CuNb13O33 are shown in the SAED pattern (Figure 1e), matching well with its (200), (20 1 ¯ ), and (40 1 ¯ ) crystallographic planes, which confirms its monoclinic structure with C2/m space group. The EDX mapping images (Figure S3) show uniform C, Cu, Nb, and O elements distributions with only tiny Cu precipitation.

3.2. Li+-Storage Properties

To study the redox mechanism of C-CuNb13O33, a CV experiment is performed on the C-CuNb13O33/Li half-cell (Figure 2a). The initial cycle is different from the following one, which can be attributed to the irreversible polarization and the formation of thin SEI films [28]. However, the CV profiles show good repeatability after the initial cycle. The second cycle exist four obvious peak pairs respectively located at 1.63/1.73, 1.55/1.67, 1.30/1.44 and 1.16/1.26 V. The first and second pairs could be ascribed to the Nb4+/Nb5+ redox reaction, while the third and fourth pairs could correspond to the Nb3+/Nb4+ redox reaction [29].
Figure 2b shows the GCD profiles of the C-CuNb13O33/Li half-cell within 1.0–3.0 V at 0.1C. C-CuNb13O33 exhibits a large first-cycle discharge/charge capacity (270/244 mAh g−1) and high Coulombic efficiency (i.e., the charge capacity divided by the discharge capacity is 90.4%). The average operation potential during lithiation/delithiation is ~1.54 V, which is similar to the popular Li4Ti5O12 and indicates the high safety performance of C-CuNb13O33. With increasing the current rate, C-CuNb13O33 is capable of retaining large reversible capacities of 222, 209, 195, 174, 154, and 133 mAh g−1 for ten cycles each at 0.5C, 1C, 2C, 5C, 10C, and 20C, respectively (Figure 2c,d), revealing its excellent rate capability with capacity percentage of 69.4%/59.9% at 10C/20C relative to 0.5C. Meanwhile, the capacity has no obvious decay when the rate returns from 20C to 0.5C. Long-term discharge/charge tests are further performed at 10C and 20C, revealing superior cyclability with 86.2% and 92.3% retention after 3000 cycles, respectively (Figure 2e). In addition, the comprehensive properties of C-CuNb13O33 are better than not only those of CuNb13O33 (Figure S4) but also those of most intercalation-type anode materials previously reported (Table S3).

3.3. Electrochemical Kinetics

To study the Li+ transport kinetics of C-CuNb13O33, the Li+ apparent diffusion coefficients (DLi) during lithiation/delithiation are determined based on both the GITT and CV techniques. Figure 3a shows the GITT profiles of the C-CuNb13O33/Li half-cell during the initial cycle at 0.1C. The D Li GITT value can be calculated by using Equation (1), which is rooted from Fick’s second law [30]:
D Li GITT = 4 π m a V m M a S 2 Δ E s τ d E τ / d τ 2   τ     L 2 D Li GITT
where ma, Ma, Vm, S, and L present the mass, molar mass, molar volume, electrode surface area, and electrode thickness of C-CuNb13O33, respectively; τ is the time during which a constant current is applied; ΔEs means the variation of the equilibrium potential, and ΔEτ means the variation of potential during the current pulse, which can be gained from the GITT profiles (Figure 3b). Since a linear relationship is achieved between the potential E and τ0.5 during each titration (Figure 3c), Equation (1) can be simplified as Equation (2):
D Li GITT = 4 π τ m b V m M b S 2 Δ E s Δ E τ 2   τ     L 2 D Li GITT
Based on Equation (2), the D Li GITT value of C-CuNb13O33 is determined to be 4.07 × 10−11 cm2 s−1/5.94 × 10−11 cm2 s−1 during lithiation/delithiation (Figure 3d).
The DLi values of C-CuNb13O33 during lithiation/delithiation are also determined from its CV data at different sweep rates (Figure 3e). It is found that the Ip of the intensive cathodic/anodic reaction is in proportional to v0.5 (Figure 3f), which exhibits linear semi-infinite diffusion during lithiation/delithiation. Hence, the D Li CV values can be determined based on the Randles–Servick equation Equation (3) [30]:
I P = 0.4463 nFAC nFvD Li CV RT 0.5
where n, A, and C mean the number of charge transfer, electrode area, and molar concentration of Li+ in solid, respectively, and Ip, F, v, R, and T represent the peak current, Faraday constant, sweep rate, temperature, and gas constant, respectively. Equation (3) can be simplified as Equation (4) due to the fact that the electrochemical system is performed at 25 °C.
I P = 268,600 n 1.5 AD Li CV 0.5 Cv 0.5
Based on Equation (4), the D Li CV value of C-CuNb13O33 is determined to be 2.85 × 10−11 cm2 s−1/4.55 × 10−11 during lithiation/delithiation, which matches with the GITT results. It should be emphasized that the average DLi values of C-CuNb13O33 surpass those of the previously-reported niobates (Table S4). This fast Li+ diffusivity of C-CuNb13O33 is beneficial for achieving its excellent rate capability.
To analyze the capacitive behavior of C-CuNb13O33, the electrochemical kinetics obtained from the CV data (Figure 3e) is further investigated through the relationship between Ip and ν [31,32]:
I p = av b
where a and b are changeable parameters. b = 1 indicates 100% capacitive behavior, while b = 0.5 indicates 100% diffusion-controlled behavior [31]. To calculate the b value, Equation (5) can be transformed into Equation (6):
log I p = log   a + b   log   v
The b values of C-CuNb13O33 for the cathodic and anodic peaks are respectively determined to be 0.867 and 0.950 (Figure 3g). These large b values indicate that the electrochemical process is dominated by capacitive control.
For further understanding the electrochemical kinetics of C-CuNb13O33, the capacitive contribution (jv) and the diffusion-controlled contribution (kv1/2) are determined based on Equation (7) [32]:
I = jv + kv 1 / 2
where I represents the detected current at a fixed potential, and j and k are changeable parameters. Equation (7) can be transformed to Equation (8). As a result, the j value is the slope obtained through the linear relationship between the I/v1/2 and v1/2, and thus the capacitive contribution ratio can be easily obtained.
I / v 1 / 2 = jv 1 / 2 + k
The large capacitive contribution ratios of C-CuNb13O33 indicate dominant capacitive contributions at different sweep rates (Figure 3h), matching with the b values. It is worth noting that the capacitive contribution ratio of C-CuNb13O33 reaches 87.9% at 1.1 mV s−1 (Figure 3i). This significant capacitive behavior is also beneficial for the rate capability.

3.4. Crystal Structure Evolutions

An in-situ XRD test is performed to clarify crystal-structural evolutions of CuNb13O33 during lithiation/delithiation. Figure 4a exhibits the initial-three-cycle original and contour in-situ XRD patterns with GCD profiles. During the initial lithiation, all the peaks exhibit decreased intensities because the Li+ insertion undoubtedly decreases the lattice order [33]. The (110), (51 1 ¯ ), (313), (51 2 ¯ ), and (114) peaks shift toward smaller angles. The (204) and (60 2 ¯ ) peaks shift toward smaller angles until ~1.6 V and then slowly shifts towards larger angles until ~1.4 V. The (204) peak shifts toward smaller angles and the (60 2 ¯ ) peak continues shifting toward larger angles until ~1.0 V. Although these peak evolutions are complex, they almost fully recover their initial positions and intensities during the following delithiation. The subsequent peak evolutions, however, are almost similar to those in the initial cycle, indicating the excellent electrochemical stability of C-CuNb13O33. The lithiation/delithiation mechanism in CuNb13O33 can be described based on Equation (9):
CuNb13O33 + xe + xLi+ ↔ LixCuNb13O33 (0 ≤ x ≤ 26)
Figure 4b illustrates the lattice-parameter changes of C-CuNb13O33 gained by Rietveld-refining the in-situ XRD patterns. During the Li+ insertion, the a and c-value changes follow a sequence of increase → decrease → slight increase. The total a- and c-value changes are only −1.49 and −0.13%, respectively. The b value gradually increases with a maximum change of 8.93%. The obvious increase of the b value suggests that Li+ stores in the (010) crystallographic plane during lithiation. The β value first increases and then decreases with a total change of −0.33%. Consequently, the V value increases by 7.61%. During the Li+ extraction, all the lattice-parameter changes are very reversible to those during the Li+ insertion, which matches well with the peak evolutions. These changes of the lattice parameters are in good agreement with the ex-situ HRTEM result, which exhibits that the (110) d-spacing increases from 0.377 nm (pristine state, Figure 4c) to 0.411 nm (discharge to 1.0 V, Figure 4d) and returns to 0.377 nm (charge to 3.0 V, Figure 4e).

4. Conclusions

~1 wt% carbon-coated CuNb13O33 microparticles (C-CuNb13O33) with a stable shear ReO3 structure are explored as a new anode material with comprehensively good Li+ storage properties. This new material owns the Nb4+/Nb5+ and Nb3+/Nb4+ redox couples, enabling a safe operation potential (~1.54 V), high practical capacity (244 mAh g−1), and a high initial-cycle Coulombic efficiency (90.4%) at 0.1C. Its very fast Li+ transport (5.01 × 10−11 cm2 s−1) and significant capacitive behavior result in its excellent rate capability with capacity percentages of 69.4% (10C relative to 0.5C) and 59.9% (20C relative to 0.5C). Additionally, it exhibits superior cyclability with capacity retention of 86.2%/92.3% at 10C/20C after 3000 cycles due to its maximum unit-cell-volume change of only 7.61%. Therefore, C-CuNb13O33 is an ideal anode material for Li+-storage.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ma16051818/s1, Figure S1: XPS spectra of Nb and Cu elements in C-CuNb13O33; Figure S2: TGA curves of CuNb13O33 and C-CuNb13O33.; Figure S3: EDX mapping images of C-CuNb13O33.; Figure S4: Electrochemical properties of CuNb13O33/Li half-cell; Table S1: Details of Rietveld-refinement and crystal data of CuNb13O33; Table S2: Fractional atomic parameters of CuNb13O33 with C2/m; Table S3: Comparisons of electrochemical properties of C-CuNb13O33 with those of intercalation-type anode materials previously reported; Table S4: Comparisons of apparent Li+ diffusion coefficient (DLi) of C-CuNb13O33 with that of previously-reported niobates at 25 °C. References [19,23,24,28,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49] are cited in the supplementary materials.

Author Contributions

Conceptualization, C.L.; methodology, J.G. and C.L.; software, X.L.; validation, S.L., W.W., Y.O. and S.G.; formal analysis, J.G. and C.L.; investigation, J.G. and C.L.; resources, C.L.; data curation, X.L.; writing—original draft preparation, J.G.; writing—review and editing, C.L.; visualization, X.L.; supervision, C.L.; project administration, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, L.; Liu, T.; Wu, T.; Lu, J. Strain-retardant coherent perovskite phase stabilized Ni-rich cathode. Nature 2022, 611, 61–67. [Google Scholar] [CrossRef]
  2. Wang, C.; Liu, T.; Yang, X.; Ge, S.; Stanley, N.V.; Rountree, E.S.; Leng, Y.; McCarthy, B.D. Fast charging of energy-dense lithium-ion batteries. Nature 2022, 611, 485–490. [Google Scholar] [CrossRef]
  3. Neumann, J.; Petranikova, M.; Meeus, M.; Gamarra, J.D.; Younesi, R.; Winter, M.; Nowak, S. Recycling of lithium-ion batteries-current state of the art, circular economy, and next generation recycling. Adv. Energy Mater. 2022, 12, 2102917. [Google Scholar] [CrossRef]
  4. 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]
  5. Zhang, W.; Seo, D.H.; Chen, T.; Wu, L.; Topsakal, M.; Zhu, Y.; Lu, D.; Ceder, G.; Wang, F. Kinetic pathways of ionic transport in fast-charging lithium titanate. Science 2020, 367, 1030–1034. [Google Scholar] [CrossRef]
  6. Liu, H.; Zhu, Z.; Yan, Q.; Yu, S.; He, X.; Chen, Y.; Zhang, R.; Ma, L.; Liu, T.; Li, M.; et al. A disordered rock salt anode for fast-charging lithium-ion batteries. Nature 2020, 585, 63–67. [Google Scholar] [CrossRef]
  7. Griffith, K.J.; Wiaderek, K.M.; Cibin, G.; Marbella, L.E.; Grey, C.P. Niobium tungsten oxides for high-rate lithium-ion energy storage. Nature 2018, 559, 556–563. [Google Scholar] [CrossRef] [Green Version]
  8. Bi, Y.; Tao, J.; Wu, Y.; Li, L.; Xu, Y.; Hu, E.; Wu, B.; Hu, J.; Wang, C.; Zhang, J.; et al. Reversible planar gliding and microcracking in a single-crystalline Ni-rich cathode. Science 2020, 370, 1313–1317. [Google Scholar] [CrossRef]
  9. Schmuch, R.; Wagner, R.; Hörpel, G.; Placke, T.; Winter, M. Performance and cost of materials for lithium-based rechargeable automotive batteries. Nat. Energy 2018, 3, 267–278. [Google Scholar] [CrossRef]
  10. Hassoun, J.; Bonaccorso, F.; Agostini, M.; Angelucci, M.; Betti, M.G.; Cingolani, R.; Gemmi, M.; Mariani, C.; Panero, S.; Pellegrini, V.; et al. An advanced lithium-ion battery based on a graphene anode and a lithium iron phosphate cathode. Nano Lett. 2014, 14, 4901–4906. [Google Scholar] [CrossRef] [Green Version]
  11. Wu, Y.; Rahm, E.; Holze, R. Carbon anode materials for lithium ion batteries. J. Power Source 2003, 114, 228–236. [Google Scholar] [CrossRef]
  12. Sivakkumar, S.R.; Nerkar, J.Y.; Pandolfo, A.G. Rate capability of graphite materials as negative electrodes in lithium-ion capacitors. Electrochim. Acta 2010, 55, 3330–3335. [Google Scholar] [CrossRef]
  13. Shen, L.; Zhang, X.; Uchaker, E.; Yuan, C.; Cao, G. Li4Ti5O12 nanoparticles embedded in a mesoporous carbon matrix as a superior anode material for high rate lithium ion batteries. Adv. Energy Mater. 2012, 2, 691–698. [Google Scholar] [CrossRef]
  14. Chen, Z.; Belharouak, I.; Sun, Y.K.; Amine, K. Titanium-based anode materials for safe lithium-ion batteries. Adv. Funct. Mater. 2013, 23, 959–969. [Google Scholar] [CrossRef]
  15. Li, H.; Wang, Z.; Chen, L.; Huang, X. Research on advanced materials for Li-ion batteries. Adv. Mater. 2009, 21, 4593–4607. [Google Scholar] [CrossRef]
  16. Griffith, K.J.; Harada, Y.; Egusa, S.; Ribas, R.M.; Monteiro, R.S.; Von Dreele, R.B.; Cheetham, A.K.; Cava, R.J.; Grey, C.P.; Goodenough, J.B. Titanium niobium oxide: From discovery to application in fast-charging lithium-ion batteries. Chem. Mater. 2021, 33, 4–18. [Google Scholar] [CrossRef]
  17. Ding, H.; Song, Z.; Zhang, H.; Zhang, H.; Li, X. Niobium-based oxide anodes toward fast and safe energy storage: A review. Mater. Today Nano 2020, 11, 100082. [Google Scholar] [CrossRef]
  18. Zhang, X.; Peng, N.; Liu, T.; Zheng, R.; Xia, M.; Yu, H.; Chen, S.; Shui, M.; Shu, J. Review on niobium-based chalcogenides for electrochemical energy storage devices: Application and progress. Nano Energy 2019, 65, 104049. [Google Scholar] [CrossRef]
  19. Yang, L.; Zhu, X.; Li, X.; Zhao, X.; Pei, K.; You, W.; Li, X.; Chen, Y.; Lin, C.; Che, R. Conductive copper niobate: Superior Li+-storage capability and novel Li+-transport mechanism. Adv. Energy Mater. 2019, 9, 1902174. [Google Scholar] [CrossRef]
  20. Hu, L.; Luo, L.; Tang, L.; Lin, C.; Li, R.; Chen, Y. Ti2Nb2xO4+5x anode materials for lithium-ion batteries: A comprehensive review. J. Mater. Chem. A 2018, 6, 9799–9815. [Google Scholar] [CrossRef]
  21. Han, J.; Huang, Y.; Goodenough, J.B. New anode framework for rechargeable lithium batteries. Chem. Mater. 2011, 23, 2027–2029. [Google Scholar] [CrossRef]
  22. Wu, X.; Miao, J.; Han, W.; Hu, Y.; Chen, D.; Lee, J.; Kim, J.; Chen, L. Investigation on Ti2Nb10O29 anode material for lithium-ion batteries. Electrochem. Commun. 2012, 25, 39–42. [Google Scholar] [CrossRef]
  23. Lv, C.; Lin, C.; Zhao, X. Rational design and synthesis of nickel niobium oxide with high-rate capability and cycling stability in a wide temperature range. Adv. Energy Mater. 2021, 12, 2102550. [Google Scholar] [CrossRef]
  24. Wang, W.; Zhang, Q.; Jiang, T.; Li, S.; Gao, J.; Liu, X.; Lin, C. Conductive LaCeNb6O18 with a very open A-site-cation-deficient perovskite structure: A fast- and stable-charging Li+-storage anode compound in a wide temperature range. Adv. Energy Mater. 2022, 12, 2200656. [Google Scholar] [CrossRef]
  25. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef] [Green Version]
  26. Cotton, F.A.; Sandor, R.B. Synthesis and crystal structure of CuNb13O33 and phase analysis of the Cu-NbO2 system. Eur. J. Solid State Inorg. Chem. 1988, 25, 637–644. [Google Scholar]
  27. Su, R.; Li, Q.; Chen, Y.; Gao, B.; Yue, Q.; Zhou, W. One-step synthesis of Cu2O@carbon nanocapsules composites using sodium alginate as template and characterization of their visible light photocatalytic properties. J. Clean. Prod. 2019, 209, 20–29. [Google Scholar] [CrossRef]
  28. Qian, S.; Yu, H.; Yan, L.; Zhu, H.; Cheng, X.; Xie, Y.; Long, N.; Shui, M.; Shu, J. High-rate long-life pored nanoribbon VNb9O25 built by interconnected ultrafine nanoparticles as anode for lithium-ion batteries. ACS Appl. Mater. Interfaces 2017, 9, 30608–30616. [Google Scholar] [CrossRef]
  29. Chen, Z.; Cheng, X.; Ye, W.; Zheng, R.; Zhu, H.; Yu, H.; Long, N.; Shui, M.; Shu, J. AgNb13O33: A new anode material with high energy storage performance. Chem. Eng. J. 2019, 366, 246–253. [Google Scholar] [CrossRef]
  30. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; Wiley: New York, NY, USA, 2001. [Google Scholar]
  31. Augustyn, V.; Come, J.; Lowe, M.A.; Kim, J.W.; Taberna, P.L.; Tolbert, S.H.; Abruna, H.D.; Simon, P.; Dunn, B. High-rate electrochemical energy storage through Li+ intercalation pseudocapacitance. Nat. Mater. 2013, 12, 518–522. [Google Scholar] [CrossRef]
  32. Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive oxide materials for high-rate electrochemical energy storage. Energy Environ. Sci. 2014, 7, 1597–1614. [Google Scholar] [CrossRef] [Green Version]
  33. Liang, G.; Yang, L.; Han, Q.; Chen, G.; Lin, C.; Chen, Y.; Luo, L.; Liu, X.; Li, Y.; Che, R. Conductive Li3.08Cr0.02Si0.09V0.9O4 anode material: Novel “zero-strain” characteristic and superior electrochemical Li+ storage. Adv. Energy Mater. 2020, 10, 1904267. [Google Scholar] [CrossRef]
  34. Yang, C.; Lin, C.; Lin, S.; Chen, Y.; Li, J. Cu0.02Ti0.94Nb2.04O7: An advanced anode material for lithium-ion batteries of electric vehicles. J. Power Source 2016, 328, 336–344. [Google Scholar] [CrossRef]
  35. Yang, C.; Yu, S.; Ma, Y.; Lin, C.; Xu, Z.; Zhao, H.; Wu, S.; Zheng, P.; Zhu, Z.-Z.; Li, J.; et al. Cr3+ and Nb5+ co-doped Ti2Nb10O29 materials for high-performance lithium-ion storage. J. Power Source 2017, 360, 470–479. [Google Scholar] [CrossRef]
  36. Yang, C.; Deng, S.; Lin, C.; Lin, S.; Chen, Y.; Li, J.; Wu, H. Porous TiNb24O62 microspheres as high-performance anode materials for lithium-ion batteries of electric vehicles. Nanoscale 2016, 8, 18792–18799. [Google Scholar] [CrossRef]
  37. Zhu, X.; Xu, J.; Luo, Y.; Fu, Q.; Liang, G.; Luo, L.; Chen, Y.; Lin, C.; Zhao, X. MoNb12O33 as a new anode material for high-capacity, safe, rapid and durable Li+ storage: Structural characteristics, electrochemical properties and working mechanisms. J. Mater. Chem. A 2019, 7, 6522–6532. [Google Scholar] [CrossRef]
  38. Lou, X.; Xu, Z.; Luo, Z.; Lin, C.; Yang, C.; Zhao, H.; Zheng, P.; Li, J.; Wang, N.; Chen, Y.; et al. Exploration of Cr0.2Fe0.8Nb11O29 as an advanced anode material for lithium-ion batteries of electric vehicles. Electrochimica Acta 2017, 245, 482–488. [Google Scholar] [CrossRef]
  39. Lou, X.; Fu, Q.; Xu, J.; Liu, X.; Lin, C.; Han, J.; Luo, Y.; Chen, Y.; Fan, X.; Li, J. GaNb11O29 nanowebs as high-performance anode materials for lithium-ion batteries. ACS Appl. Nano Mater. 2018, 1, 183–190. [Google Scholar] [CrossRef]
  40. Lou, X.; Li, R.; Zhu, X.; Luo, L.; Chen, Y.; Lin, C.; Li, H.; Zhao, X. New anode material for lithium-ion batteries: Aluminum niobate (AlNb11O29). ACS Appl. Mater. Interfaces 2019, 11, 6089–6096. [Google Scholar] [CrossRef]
  41. Zhu, X.; Fu, Q.; Tang, L.; Lin, C.; Xu, J.; Liang, G.; Li, R.; Luo, L.; Chen, Y. Mg2Nb34O87 porous microspheres for use in high-energy, safe, fast-charging, and stable lithium-ion batteries. ACS Appl. Mater. Interfaces 2018, 10, 23711–23720. [Google Scholar] [CrossRef]
  42. Fu, Q.; Liu, X.; Hou, J.; Pu, Y.; Lin, C.; Yang, L.; Zhu, X.; Hu, L.; Lin, S.; Luo, L.; et al. Highly conductive CrNb11O29 nanorods for use in high-energy, safe, fast-charging and stable lithium-ion batteries. J. Power Source 2018, 397, 231–239. [Google Scholar] [CrossRef]
  43. Li, R.; Liang, G.; Zhu, X.; Fu, Q.; Chen, Y.; Luo, L.; Lin, C. Mo3Nb14O44: A new Li+ container for high-performance electrochemical energy storage. Energy Environ. Mater. 2021, 4, 65–71. [Google Scholar] [CrossRef]
  44. Fu, Q.; Li, R.; Zhu, X.; Liang, G.; Luo, L.; Chen, Y.; Lin, C.; Zhao, X. Design, synthesis and lithium-ion storage capability of Al0.5Nb24.5O62. J. Mater. Chem. A 2019, 7, 19862–19871. [Google Scholar] [CrossRef]
  45. Lin, C.; Wang, G.; Lin, S.; Li, J.; Lu, L. TiNb6O17: A New Electrode Material for Lithium-Ion Batteries. Chem. Commun. 2015, 51, 8970–8973. [Google Scholar] [CrossRef]
  46. Lin, C.; Yu, S.; Zhao, H.; Wu, S.; Wang, G.; Yu, L.; Li, Y.; Zhu, Z.Z.; Li, J.; Lin, S. Defective Ti2Nb10O27.1: An advanced anode material for lithium-ion batteries. Sci. Rep. 2015, 5, 17836. [Google Scholar] [CrossRef] [Green Version]
  47. Li, R.; Qin, Y.; Liu, X.; Yang, L.; Lin, C.; Xia, R.; Lin, S.; Chen, Y.; Li, J. Conductive Nb25O62 and Nb12O29 anode materials for use in high-performance lithium-ion storage. Electrochimica Acta 2018, 266, 202–211. [Google Scholar] [CrossRef]
  48. Hu, L.; Lu, R.; Tang, L.; Xia, R.; Lin, C.; Luo, Z.; Chen, Y.; Li, J. TiCr0.5Nb10.5O29/CNTs nanocomposite as an advanced anode material for high-performance Li+-ion storage. J. Alloys Compd. 2018, 732, 116–123. [Google Scholar] [CrossRef]
  49. Yang, C.; Yu, S.; Lin, C.; Lv, F.; Wu, S.; Yang, Y.; Wang, W.; Zhu, Z.; Li, J.; Wang, N.; et al. Cr0.5Nb24.5O62 nanowires with high electronic conductivity for high-rate and long-life lithium-ion storage. ACS Nano 2017, 11, 4217–4224. [Google Scholar] [CrossRef]
Figure 1. Physico–chemical characterizations of C-CuNb13O33. (a) XRD pattern of C-CuNb13O33 with Rietveld-refinement (main diffraction peaks are labeled). (b) Crystal structure of CuNb13O33 (C2/m). (c) FESEM image. (d) HRTEM image. (e) SAED pattern.
Figure 1. Physico–chemical characterizations of C-CuNb13O33. (a) XRD pattern of C-CuNb13O33 with Rietveld-refinement (main diffraction peaks are labeled). (b) Crystal structure of CuNb13O33 (C2/m). (c) FESEM image. (d) HRTEM image. (e) SAED pattern.
Materials 16 01818 g001
Figure 2. Electrochemical properties of C-CuNb13O33/Li half-cell. (a) CV curves at 0.2 mV s−1. (b) Initial four-cycle GCD profiles at 0.1C. (c) GCD profiles at different current rates, (d) rate capability, (e) Cyclability over 3000 cycles at 10C and 20C. 1C = 378 mAh g−1.
Figure 2. Electrochemical properties of C-CuNb13O33/Li half-cell. (a) CV curves at 0.2 mV s−1. (b) Initial four-cycle GCD profiles at 0.1C. (c) GCD profiles at different current rates, (d) rate capability, (e) Cyclability over 3000 cycles at 10C and 20C. 1C = 378 mAh g−1.
Materials 16 01818 g002
Figure 3. Electrochemical kinetics of C-CuNb13O33/Li half-cell. (a) GITT lithiation/delithiation profile. (b) Potential (E) and time (t) profile of single step in GITT test. (c) Relationship between E and square root of titration duration (τ0.5) during a typical titration. (d) Variations of D Li GITT during lithiation/delithiation. (e) CV curves at variable sweep rates. (f) Relationship between peak current (Ip) and square root of sweep rate (ν0.5) for intensive cathodic/anodic peaks. (g) Calculations of b values through the relationship between Ip and ν. (h) Pseudocapacitive contribution ratios at different sweep rates. (i) Pseudocapacitive contribution ratio at 1.1 mV s−1.
Figure 3. Electrochemical kinetics of C-CuNb13O33/Li half-cell. (a) GITT lithiation/delithiation profile. (b) Potential (E) and time (t) profile of single step in GITT test. (c) Relationship between E and square root of titration duration (τ0.5) during a typical titration. (d) Variations of D Li GITT during lithiation/delithiation. (e) CV curves at variable sweep rates. (f) Relationship between peak current (Ip) and square root of sweep rate (ν0.5) for intensive cathodic/anodic peaks. (g) Calculations of b values through the relationship between Ip and ν. (h) Pseudocapacitive contribution ratios at different sweep rates. (i) Pseudocapacitive contribution ratio at 1.1 mV s−1.
Materials 16 01818 g003
Figure 4. Crystal structure evolutions of C-CuNb13O33. (a) Original in-situ XRD patterns with GCD profiles (0.4C) and contour in-situ XRD patterns of C-CuNb13O33/Li in-situ cell (first three cycles). (b) Lattice-parameter variations of C-CuNb13O33 (initial three cycles). Ex-situ HRTEM characterization of (c) original, (d) lithiated (1.0 V), and (e) delithiated (3.0 V) C-CuNb13O33 samples.
Figure 4. Crystal structure evolutions of C-CuNb13O33. (a) Original in-situ XRD patterns with GCD profiles (0.4C) and contour in-situ XRD patterns of C-CuNb13O33/Li in-situ cell (first three cycles). (b) Lattice-parameter variations of C-CuNb13O33 (initial three cycles). Ex-situ HRTEM characterization of (c) original, (d) lithiated (1.0 V), and (e) delithiated (3.0 V) C-CuNb13O33 samples.
Materials 16 01818 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

Gao, J.; Li, S.; Wang, W.; Ou, Y.; Gao, S.; Liu, X.; Lin, C. Carbon-Coated CuNb13O33 as A New Anode Material for Lithium Storage. Materials 2023, 16, 1818. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16051818

AMA Style

Gao J, Li S, Wang W, Ou Y, Gao S, Liu X, Lin C. Carbon-Coated CuNb13O33 as A New Anode Material for Lithium Storage. Materials. 2023; 16(5):1818. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16051818

Chicago/Turabian Style

Gao, Jiazhe, Songjie Li, Wenze Wang, Yinjun Ou, Shangfu Gao, Xuehua Liu, and Chunfu Lin. 2023. "Carbon-Coated CuNb13O33 as A New Anode Material for Lithium Storage" Materials 16, no. 5: 1818. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16051818

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