Next Article in Journal
Single-Shot near Edge X-ray Fine Structure (NEXAFS) Spectroscopy Using a Laboratory Laser-Plasma Light Source
Previous Article in Journal
Spiral Bevel Gears Face Roughness Prediction Produced by CNC End Milling Centers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Cycling Stability of LiCuxMn1.95−xSi0.05O4 Cathode Material Obtained by Solid-State Method

1
School of Mechanical & Electrical Engineering, Henan Institute of Science and Technology, Xinxiang 453003, China
2
Research Branch of Advanced Materials & Green Energy, Henan Institute of Science and Technology, Xinxiang 453003, China
3
School of Chemistry and Chemical Engineering, Henan Institute of Science and Technology, Xinxiang 453003, China
*
Authors to whom correspondence should be addressed.
Materials 2018, 11(8), 1302; https://doi.org/10.3390/ma11081302
Submission received: 9 July 2018 / Revised: 20 July 2018 / Accepted: 25 July 2018 / Published: 27 July 2018
(This article belongs to the Section Energy Materials)

Abstract

:
The LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples have been obtained by a simple solid-state method. XRD and SEM characterization results indicate that the Cu-Si co-doped spinels retain the inherent structure of LiMn2O4 and possess uniform particle size distribution. Electrochemical tests show that the optimal Cu-doping amount produces an obvious improvement effect on the cycling stability of LiMn1.95Si0.05O4. When cycled at 0.5 C, the optimal LiCu0.05Mn1.90Si0.05O4 sample exhibits an initial capacity of 127.3 mAh g−1 with excellent retention of 95.7% after 200 cycles. Moreover, when the cycling rate climbs to 10 C, the LiCu0.05Mn1.90Si0.05O4 sample exhibits 82.3 mAh g−1 with satisfactory cycling performance. In particular, when cycled at 55 °C, this co-doped sample can show an outstanding retention of 94.0% after 100 cycles, whiles the LiMn1.95Si0.05O4 only exhibits low retention of 79.1%. Such impressive performance shows that the addition of copper ions in the Si-doped spinel effectively remedy the shortcomings of the single Si-doping strategy and the Cu-Si co-doped spinel can show excellent cycling stability.

1. Introduction

Lithium-ion batteries have been applied extensively in a lot of power supply fields, like in pure electrical vehicles (EVs), unmanned aerial vehicles and smartphones. As one important part of lithium-ion batteries, cathode materials have played a crucial role in terms of electrochemical performance [1,2,3,4,5,6,7]. Among the existing cathode materials, LiMn2O4 possesses major advantages and great potential for the large-scale commercial application due to the mature production technology, cheap production costs and non-pollution characteristics [8,9,10]. It is important to note, however, that this material shows poor cycling stability and elevated-temperature performance, which produces a serious negative effect on promoting the large-scale commercial application. These unsatisfactory deficiencies are mainly caused by Jahn-Teller distortion and manganese dissolution [11,12,13,14].
According to the existing literatures [15,16,17,18], the body-doping strategy can improve the cycling stability to some degree by introducing other cations in the spinel structure. The common doping ions mainly include the monovalent ion (Li+) [19,20], divalent ions (Mg2+, Zn2+, Cu2+, etc.) [21,22,23,24], and trivalent ions (Al3+, Co3+, Cr3+, etc.) [25,26,27,28]. The research results have established that doping the trivalent manganese ions with these low valence cations can markedly improve the cycling life. However, introducing these low valence cations usually produces certain negative effects on the reversible capacity due to the decrease of Mn3+ ions. Considering this, doping the manganese ions with tetravalent cations has been developed to improve the electrochemical performance of LiMn2O4 because this strategy can avoid the decrease of trivalent manganese ions and reversible capacity loss of LiMn2O4 and provide the more expanded and stable MnO6 octahedra, which is conducive to the diffusion of lithium ions [29,30,31]. In the previous work [32], the Si-doped LiMn2O4 samples have been obtained by solid-state method. When cycled at 0.5 C, the optimal sample can peak at 134.6 mAh g−1. Unfortunately, the capacity retention is only 85.1% after 100 cycles. It was obvious that the optimization degree of replacing the Mn4+ ions with tetravalent cations cannot reach the demand for large-scale application of LiMn2O4.
It has been reported that the Cu-doping strategy can make a positive contribution in enhancing the cycling stability due to the fact that the addition of copper ions in the LiMn2O4 decrease the trivalent manganese ions and cell volume of LiMn2O4, which can inhibit the Jahn-Teller effect and enhance structural stability [23]. In this work, the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples have been obtained by a simple solid-state method. The effect of copper doping content on the structures, morphologies and cycling life of the LiCuxMn1.95−xSi0.05O4 samples is discussed. The results indicate the addition of copper ions in the Si-doped spinel effectively remedy the shortcomings of the single Si-doping strategy and the Cu-Si co-doped spinel can show excellent cycling stability.

2. Materials and Methods

The LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples were synthesized by traditional high temperature solid-state reaction process using Li2CO3, electrolytic MnO2, C8H20O4Si and Cu(NO3)2 as reaction reagents. Firstly, the hydro-ball-milling technique was used to pretreat the electrolytic MnO2. Then, Li2CO3, electrolytic MnO2, Cu(NO3)2 and ethanol solution of C8H20O4Si were mixed thoroughly by hydro-ball-milling. The obtained slurries were dried at 70 °C and further ground into powder. Subsequently, this material was sintered at 450 °C for 4 h in air and then reground after natural cooling. The desired product LiCuxMn1.95−xSi0.05O4 were obtained by calcining at 825 °C for 18 h in air.
The crystal structures of the obtained LiCuxMn1.95−xSi0.05O4 samples were studied by X-ray diffraction technique (XRD, Bruker DX-1000, Karlsruhe, Germany) with Cu Kα radiation (λ = 0.15406 nm). The scanning electron microscopy (SEM, JEOL JSM-6360LV, Tokyo, Japan) analytical techniques were used to study the surface morphologies and microstructures.
The active electrode consists of the obtained LiCuxMn1.95−xSi0.05O4 samples, conductive acetylene black and polyvinylidene fluoride (Weight Ratio = 85:10:5). The anode material and diaphragm are lithium foil and Celgard 2400 polymer, respectively. The mixture of 1 M LiPF6, ethyl methyl carbonate (EMC), ethylene carbonate (EC) and dimethyl carbonate (DMC) was used as electrolyte (EMC:EC:DMC = 1:1:1). The electrochemical measurement was executed on LAND (Wuhan, China) battery testing system. The electrochemical impedance spectroscopy (EIS) were tested by CS-350 electrochemical workstation (Wuhan, China). These tests were investigated by using CR2016 coin-type cells.

3. Results and Discussion

Figure 1 shows the XRD results of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples. As shown here, the characteristic peaks of all these samples match with that of LiMn2O4 (JCPDS No. 35-0782), implying that the Cu-doping strategy have no material impact on the inherent structure of LiMn2O4 [17,33], where lithium and manganese ions occupy the tetrahedral sites (8a) and octahedral sites (16d), respectively. According to the reported research result, the (220) characteristic peak may be observed if other cations occupied the tetrahedral sites [34]. However, there is no (220) characteristic peak in Figure 1, suggesting that the copper ions successfully replaced the manganese ions in octahedral sites.
According to the reported literature [35], the intensity ratio of (311)/(400) peaks can be optimized by replacing the Mn ions with some other cation ions in the spinel structure of LiMn2O4. If this intensity ratio is in the range of 0.96–1.10, the obtained samples usually show excellent cycling stability. Table 1 lists this intensity ratio of LiCuxMn1.95-xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples. It can be seen that the Cu-doping strategy has played a positive role in optimizing this intensity ratio. The copper and silicon co-doped spinels can present a larger intensity ratio than that of the silicon co-doped spinel. Therefore, it can be inferred that the further addition of copper ions in the silicon-doped sample may greatly enhance the cycling stability.
Figure 2 shows the SEM images of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples. As shown in Figure 2a, the silicon-doped LiMn2O4 particles present less-than-ideal size distribution. For the copper and silicon co-doped LiMn2O4 samples, the introduction of some copper ions can further optimize the mean diameter and size distribution. When the copper doping content increases, the mean diameter of the LiCuxMn1.95−xSi0.05O4 (x = 0.02, 0.05, 0.08) has a decreasing tendency. It is important to note that the LiCu0.05Mn1.90Si0.05O4 particles shown in Figure 2c present the quite uniform size distribution. The above-mentioned results suggest that introducing some copper ions can effectively improve the crystallinity and optimize the size distribution, which is conducive to the enhancement of cycling stability.
Figure 3a shows the first discharge curves of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples. All these samples present characteristic discharge curves, which show two distinct voltage platforms around 4.10–4.15 V and 3.95–4.00 V, suggesting that introducing some copper ions do not change the electrochemical redox reaction mechanism and all these copper and silicon co-doped LiMn2O4 samples processes two extraction/insertion process of Li+ ions [14,32]. Figure 3b presents the cycling stability of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples. The cycling stability of these co-doped samples were remarkably enhanced as the copper doping content increased, due to the suppressed Jahn-Teller effect and stronger structural stability [23]. Note, however, that the addition of more copper ions has a great negative impact on the reversible capacity of the LiCu0.08Mn1.87Si0.05O4 sample in spite of the improvement of cycling life (Figure 3c). These unsatisfactory results are principally because introducing more copper ions can cause the reduction of Mn3+, which is unfavourable to the Mn(III)–Mn(IV) interconversion.
Figure 3d shows the long cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples. For the LiCu0.05Mn1.90Si0.05O4 sample, the reversible capacity peaked at 127.3 mAh g−1, which is slightly lower than that of the LiMn1.95Si0.05O4 sample. After 200 cycles, the LiCu0.05Mn1.90Si0.05O4 sample can still exhibit 121.8 mAh g−1 with outstanding retention of 95.7%. Unfortunately, the LiMn1.95Si0.05O4 sample shows lower capacity with worse cycling life. After 200 cycles, this sample only delivers 108.3 mAh g−1 with low retention of 81.6%. According to the reference [32], the undoped LiMn2O4 only delivers a discharge capacity of 48.3 mAh g−1 with capacity retention of 37.8% after 100 cycles, which is much lower than that of the LiSi0.05Mn1.95O4 sample. Although the silicon-doping enhance the cycling performance, the further addition of copper ions can significantly enhance the cycling stability of LiMn2O4.
Figure 4a shows the corresponding discharge curves of the representative LiCu0.05Mn1.90Si0.05O4 sample at varying rates. It can be seen that there are two voltage platforms which are obvious at 0.2 C and 0.5 C, suggesting the diffusion process of lithium ions [36]. When the rate further increases, these two potential plateaus gradually show ambiguous boundary and shifted toward lower voltage. This result has a lot to do with the polarization effect and ohmic drop [37]. Figure 4b shows the cycling stability of the LiCu0.05Mn1.90Si0.05O4 and LiMn1.95Si0.05O4 samples at varying rates. When cycled at 0.2 C, the capacities of these two samples reached up to 138.5 and 131.4 mAh g−1, respectively. However, what is important to pay attention to is the reversible capacity of the LiCu0.05Mn1.90Si0.05O4 sample, which showed much more obvious difference at high rates of 5.0 C.
To further study the cycling performance at a high rate, the LiCu0.05Mn1.90Si0.05O4 and LiMn1.95Si0.05O4 samples were tested at 10 C. For the LiCu0.05Mn1.90Si0.05O4 sample, the two characteristic voltage plateaus shown in Figure 4c become blurred to a certain extent. By contrast, the LiMn1.95Si0.05O4 presents lower voltage plateau and corresponding to this, the capacity of this material shows severe degradation. Figure 4d presents the cycling life of these two spinels at 10 C. The LiMn1.95Si0.05O4 sample shows unsatisfactory capacity retention of 85.7% with low initial capacity of 68.4 mAh g−1, while the LiCu0.05Mn1.90Si0.05O4 sample can display a higher capacity of 82.3 mAh g−1. More importantly, after 100 cycles, the corresponding retention can reach up to 94.0% with the 100th cycle with a capacity of 77.4 mAh g−1. The above discussion indicates that the introduction of copper ions has great value in the optimization of the rate capability.
Figure 5 shows the electrochemical properties of the LiCu0.05Mn1.90Si0.05O4 and LiMn1.95Si0.05O4 samples at 55 °C. As shown in Figure 5a, the LiCu0.05Mn1.90Si0.05O4 exhibits an initial capacity of 127.2 mAh g−1 at 0.5 C. After 100 cycles, this sample still maintains a high capacity of 119.6 mAh g−1 with excellent retention of 94.0%. However, the LiMn1.95Si0.05O4 sample shows much lower retention than that of the LiCu0.05Mn1.90Si0.05O4. The capacity retention of the LiMn1.95Si0.05O4 sample is only 79.1% with a lower capacity of 106.4 mAh g−1 after 100th cycle. Such low discharge capacity after 100 cycles is mostly given to the fact that the high temperature accelerates the dissolution of manganese and undermines the structural stability of LiMn2O4. Note, however, that the LiCu0.05Mn1.90Si0.05O4 sample can still show much better cycling stability although these two samples show low discharge capacity after 100 cycles. These results suggest that introducing some copper ions can be favorable for enhancing the cycling stability at high-temperature. Figure 5b shows the rate capability of these two samples at 55 °C. When cycled at low rates, the LiCu0.05Mn1.90Si0.05O4 and LiMn1.95Si0.05O4 samples show similar capacities. However, as the cycling rate increased, these two samples gradually show some difference. When cycled at 5.0 C, the LiCu0.05Mn1.90Si0.05O4 sample can show 103.4 mAh g−1 while the LiMn1.95Si0.05O4 only shows 91.7 mAh g−1. The above-mentioned results suggest that the addition of copper ions can further improve the rate capability of LiMn1.95Si0.05O4 at high-temperature.
Figure 6a,b show the EIS results of the LiCu0.05Mn1.90Si0.05O4 and LiMn1.95Si0.05O4 samples. It has been reported previously that the charge transfer resistance (R2) corresponds to the high-frequency semicircle, which has a connection with the cycling life [14,34]. Therefore, we mainly determine the R2 values to confirm the effect of introducing copper ions on the cycling stability. Table 2 lists the fitting values of R2. For the LiCu0.05Mn1.90Si0.05O4 sample, the original R2 value only reach 70.2 Ω cm2 but increase to 116.0 Ω cm2 after 200 cycles. The R2 value increase was relatively small with low growth rate of 64.5%. By contrast, the LiMn1.95Si0.05O4 shows a higher original R2 value (93.2 Ω cm2). However, this value increases to 158.1 Ω cm2 with growth rate of 69.6%. Through the above analysis, it is concluded that replacing some trivalent manganese ions with copper ions can have a constructive role in decreasing the R2 value and enhancing the diffusion of lithium ions [33,38,39].

4. Conclusions

The LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples have been obtained by a simple solid-state method. The further addition of copper ions in the LiMn2O4 can decrease the trivalent manganese ions and cell volume of LiMn2O4, which can inhibit the Jahn-Teller effect and enhance structural stability. As the optimal Cu-Si co-doped spinel, the LiCu0.05Mn1.90Si0.05O4 sample possessed even size distribution. More importantly, it showed much better cycling stability and elevated temperature performance than the Si-doped LiMn2O4 sample. When cycled at 0.5 C, the LiCu0.05Mn1.90Si0.05O4 sample exhibited 127.3 mAh g−1, which is slightly lower than that of the LiMn1.95Si0.05O4 sample. After 200 cycles, the LiCu0.05Mn1.90Si0.05O4 sample could exhibit 121.8 mAh g−1 with outstanding retention of 95.7% at 0.5 C. Moreover, this co-doped sample can show outstanding rate capability and high-temperature performance. All these results suggest that the further addition of copper ions can produce an obvious effect in enhancing the cycling stability of the silicon-doped LiMn2O4.

Author Contributions

H.Z. and J.S. conceived and designed the experiments; H.Z. and F.L. performed the experiments; all authors analyzed the data; H.Z. wrote the paper; all authors discussed the results and commented on the paper.

Acknowledgments

This research was funded by the Landmark Innovation Project of Henan Institute of Science and Technology (No. 203010916004), High-level Talents Introduction Project of Henan Institute of Science and Technology (No. 203010617011) and Key Research Project of Education Department of Henan Province (No. 19A150023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, W.D.; Song, B.H.; Manthiram, A. High-voltage positive electrode materials for lithium-ion batteries. Chem. Soc. Rev. 2017, 46, 3006–3059. [Google Scholar] [CrossRef] [PubMed]
  2. Dell’Era, A.; Pasquali, M.; Bauer, E.M.; Vecchio Ciprioti, S.; Scaramuzzo, F.A.; Lupi, C. Synthesis, characterization, and electrochemical behavior of LiMnxFe(1−x)PO4 composites obtained from phenylphosphonate-based organic-inorganic hybrids. Materials 2017, 11, 56. [Google Scholar] [CrossRef] [PubMed]
  3. Xia, H.; Xia, Q.Y.; Lin, B.H.; Zhu, J.W.; Seo, J.K.; Meng, Y.S. Self-standing porous LiMn2O4 nanowall arrays as promising cathodes for advanced 3D microbatteries and flexible lithium-ion batteries. Nano Energy 2016, 22, 475–482. [Google Scholar] [CrossRef]
  4. Yoo, K.S.; Kang, Y.H.; Im, K.R.; Kim, C.S. Surface modification of Li(Ni0.6Co0.2Mn0.2)O2 cathode materials by nano-Al2O3 to improve electrochemical performance in lithium-ion batteries. Materials 2017, 10, 1273. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, S.S.; Zhao, H.Y.; Tan, M.; Hu, Y.Z.; Shu, X.H.; Zhang, M.L.; Chen, B.; Liu, X.Q. Er-doped LiNi0.5Mn1.5O4 cathode material with enhanced cycling stability for lithium-ion batteries. Materials 2017, 10, 859. [Google Scholar] [CrossRef] [PubMed]
  6. Zhao, H.Y.; Liu, S.S.; Cai, Y.; Wang, Z.W.; Tan, M.; Liu, X.Q. A simple and mass production preferred solid-state procedure to prepare the LiSixMgxMn2−2xO4 (0 ≤ x ≤ 0.10) with enhanced cycling stability and rate capability. J. Alloys Compd. 2016, 671, 304–311. [Google Scholar] [CrossRef]
  7. Nitta, N.; Wu, F.X.; Lee, J.T.; Yushin, G. Li-ion battery materials: Present and future. Mater. Today 2015, 18, 252–264. [Google Scholar] [CrossRef]
  8. Yang, S.; Schmidt, D.O.; Khetan, A.; Schrader, F.; Jakobi, S.; Homberger, M.; Noyong, M.; Paulus, A.; Kungl, H.; Eichel, R.A.; et al. Electrochemical and electronic charge transport properties of Ni-doped LiMn2O4 spinel obtained from polyol-mediated synthesis. Materials 2018, 11, 806. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, H.L.; Li, Z.H.; Yu, S.S.; Xiao, Q.Z.; Lei, G.T.; Ding, Y.H. Carbon-encapsulated LiMn2O4 spheres prepared using a polymer microgel reactor for high-power lithium-ion batteries. J. Power Sources 2016, 301, 376–385. [Google Scholar] [CrossRef]
  10. Mughal, M.Z.; Amanieu, H.-Y.; Moscatelli, R.; Sebastiani, M. A comparison of microscale techniques for determining fracture toughness of LiMn2O4 particles. Materials 2017, 10, 403. [Google Scholar] [CrossRef] [PubMed]
  11. Zhao, H.Y.; Liu, S.S.; Wang, Z.W.; Cai, Y.; Tan, M.; Liu, X.Q. Enhanced elevated-temperature performance of LiAlxSi0.05Mg0.05Mn1.90−xO4 (0 ≤ x ≤ 0.08) cathode materials for high-performance lithium-ion batteries. Electrochim. Acta 2016, 199, 18–26. [Google Scholar] [CrossRef]
  12. Tron, A.; Park, Y.D.; Mun, J.Y. AlF3-coated LiMn2O4 as cathode material for aqueous rechargeable lithium battery with improved cycling stability. J. Power Sources 2016, 325, 360–364. [Google Scholar] [CrossRef]
  13. Zhang, L.G.; Zhang, Y.R.; Yuan, X.H. Enhanced high-temperature performances of LiMn2O4 cathode by LiMnPO4 coating. Ionics 2014, 21, 37–41. [Google Scholar] [CrossRef]
  14. Yi, X.; Wang, X.Y.; Ju, B.W.; Wei, Q.L.; Yang, X.K.; Zou, G.S.; Shu, H.B.; Hu, L. Elevated temperature cyclic performance of LiAlxMn2−xO4 microspheres synthesized via co-precipitation route. J. Alloys Compd. 2014, 604, 50–56. [Google Scholar] [CrossRef]
  15. Prabu, M.; Reddy, M.V.; Selvasekarapandian, S.; Subba Rao, G.V.; Chowdari, B.V.R. (Li, Al)-co-doped spinel, Li(Li0.1Al0.1Mn1.8)O4 as high performance cathode for lithium ion batteries. Electrochim. Acta 2013, 88, 745–755. [Google Scholar] [CrossRef]
  16. Peng, C.C.; Huang, J.J.; Guo, Y.J.; Li, Q.L.; Bai, H.L.; He, Y.H.; Su, C.W.; Guo, J.M. Electrochemical performance of spinel LiAlxMn2−xO4 prepared rapidly by glucose-assisted solid-state combustion synthesis. Vacuum 2015, 120, 121–126. [Google Scholar] [CrossRef]
  17. Fang, D.L.; Li, J.C.; Liu, X.; Huang, P.F.; Xu, T.-R.; Qian, M.C.; Zheng, C.H. Synthesis of a Co–Ni doped LiMn2O4 spinel cathode material for high-power Li-ion batteries by a sol-gel mediated solid-state route. J. Alloys Compd. 2015, 640, 82–89. [Google Scholar] [CrossRef]
  18. Zhang, H.; Xu, Y.L.; Liu, D.; Zhang, X.S.; Zhao, C.J. Structure and performance of dual-doped LiMn2O4 cathode materials prepared via microwave synthesis method. Electrochim. Acta 2014, 125, 225–231. [Google Scholar] [CrossRef]
  19. Yu, F.D.; Wang, Z.B.; Chen, F.; Wu, J.; Zhang, X.G.; Gu, D.M. Crystal structure and multicomponent effects in Li1+xMn2−x−yAlyO4 cathode materials for Li-ion batteries. J. Power Sources 2014, 262, 104–111. [Google Scholar] [CrossRef]
  20. Bianchini, M.; Suard, E.; Croguennec, L.; Masquelier, C. Li-rich Li1+xMn2−xO4 spinel electrode materials: An operando neutron diffraction study during Li+ extraction/insertion. J. Phys. Chem. C 2014, 118, 25947–25955. [Google Scholar] [CrossRef]
  21. Xiang, M.W.; Su, C.W.; Feng, L.L.; Yuan, M.L.; Guo, J.M. Rapid synthesis of high-cycling performance LiMgxMn2−xO4 (x ≤ 0.20) cathode materials by a low-temperature solid-state combustion method. Electrochim. Acta 2014, 125, 524–529. [Google Scholar] [CrossRef]
  22. Xu, W.Q.; Li, Q.L.; Guo, J.M.; Bai, H.L.; Su, C.W.; Ruan, R.S.; Peng, J.H. Electrochemical evaluation of LiZnxMn2−xO4 (x ≤ 0.10) cathode material synthesized by solution combustion method. Ceram. Int. 2016, 42, 5693–5698. [Google Scholar] [CrossRef]
  23. Huang, J.J.; Li, Q.L.; Bai, H.L.; Xu, W.Q.; He, Y.H.; Su, C.W.; Peng, J.H.; Guo, J.M. Preparation and electrochemical properties of LiCuxMn2−xO4 (x ≤ 0.10) cathode material by a low-temperature molten-salt combustion method. Int. J. Electrochem. Sci. 2015, 10, 4596–4603. [Google Scholar]
  24. Zhang, H.; Liu, D.; Zhang, X.S.; Zhao, C.J.; Xu, Y.L. Microwave synthesis of LiMg0.05Mn1.95O4 and electrochemical performance at elevated temperature for lithium-ion batteries. J. Solid State Electrochem. 2014, 18, 569–575. [Google Scholar] [CrossRef]
  25. Peng, Z.D.; Jiang, Q.L.; Du, K.; Wang, W.G.; Hu, G.R.; Liu, Y.X. Effect of Cr-sources on performance of Li1.05Cr0.04Mn1.96O4 cathode materials prepared by slurry spray drying method. J. Alloys Compd. 2010, 493, 640–644. [Google Scholar] [CrossRef]
  26. Guo, D.L.; Li, B.; Chang, Z.R.; Tang, H.W.; Xu, X.H.; Chang, K.; Shangguan, E.B.; Yuan, X.Z.; Wang, H.J. Facile synthesis of LiAl0.1Mn1.9O4 as cathode material for lithium ion batteries: Towards rate and cycling capabilities at an elevated temperature. Electrochim. Acta 2014, 134, 338–346. [Google Scholar] [CrossRef]
  27. Wang, J.L.; Li, Z.H.; Yang, J.; Tang, J.J.; Yu, J.J.; Nie, W.B.; Lei, G.T.; Xiao, Q.Z. Effect of Al-doping on the electrochemical properties of a three-dimensionally porous lithium manganese oxide for lithium-ion batteries. Electrochim. Acta 2012, 75, 115–122. [Google Scholar] [CrossRef]
  28. Balaji, S.R.K.; Mutharasu, D.; Shanmugan, S.; Subramanian, N.S.; Ramanathan, K. Influence of Sm3+ ion in structural, morphological, and electrochemical properties of LiMn2O4 synthesized by microwave calcination. Ionics 2010, 16, 351–360. [Google Scholar] [CrossRef]
  29. Iturrondobeitia, A.; Goñi, A.; Palomares, V.; Gil de Muro, I.; Lezama, L.; Rojo, T. Effect of doping LiMn2O4 spinel with a tetravalent species such as Si(iv) versus with a trivalent species such as Ga(iii). Electrochemical, magnetic and ESR study. J. Power Sources 2012, 216, 482–488. [Google Scholar] [CrossRef]
  30. Wang, M.; Yang, M.; Zhao, X.Y.; Ma, L.Q.; Shen, X.D.; Cao, G.Z. Spinel LiMn2−xSixO4 (x < 1) through Si4+ substitution as a potential cathode material for lithium-ion batteries. Sci. China Mater. 2016, 59, 558–566. [Google Scholar]
  31. Xiong, L.L.; Xu, Y.L.; Zhang, C.; Zhang, Z.W.; Li, J.B. Electrochemical properties of tetravalent Ti-doped spinel LiMn2O4. J. Solid State Electrochem. 2010, 15, 1263–1269. [Google Scholar] [CrossRef]
  32. Zhao, H.Y.; Liu, S.S.; Wang, Z.W.; Cai, Y.; Tan, M.; Liu, X.Q. LiSixMn2−xO4 (x ≤ 0.10) cathode materials with improved electrochemical properties prepared via a simple solid-state method for high-performance lithium-ion batteries. Ceram. Int. 2016, 42, 13442–13448. [Google Scholar] [CrossRef]
  33. Zhao, H.Y.; Liu, X.Q.; Cheng, C.; Li, Q.; Zhang, Z.; Wu, Y.; Chen, B.; Xiong, W.Q. Synthesis and electrochemical characterizations of spinel limn1.94MO4 (M = Mn0.06, Mg0.06, Si0.06, (Mg0.03Si0.03)) compounds as cathode materials for lithium-ion batteries. J. Power Sources 2015, 282, 118–128. [Google Scholar] [CrossRef]
  34. Xiong, L.L.; Xu, Y.L.; Tao, T.; Goodenough, J.B. Synthesis and electrochemical characterization of multi-cations doped spinel LiMn2O4 used for lithium ion batteries. J. Power Sources 2012, 199, 214–219. [Google Scholar] [CrossRef]
  35. Zhao, H.Y.; Li, F.; Liu, X.Q.; Cheng, C.; Zhang, Z.; Wu, Y.; Xiong, W.Q.; Chen, B. Effects of equimolar Mg (ii) and Si (iv) co-doping on the electrochemical properties of spinel LiMn2−xMgxSixO4 prepared by citric acid assisted sol-gel method. Electrochim. Acta 2015, 151, 263–269. [Google Scholar] [CrossRef]
  36. Zhao, H.Y.; Li, F.; Liu, X.Q.; Xiong, W.Q.; Chen, B.; Shao, H.L.; Que, D.Y.; Zhang, Z.; Wu, Y. A simple, low-cost and eco-friendly approach to synthesize single-crystalline LiMn2O4 nanorods with high electrochemical performance for lithium-ion batteries. Electrochim. Acta 2015, 166, 124–133. [Google Scholar] [CrossRef]
  37. Ding, Y.L.; Xie, J.; Cao, G.S.; Zhu, T.J.; Yu, H.M.; Zhao, X.B. Single-crystalline LiMn2O4 nanotubes synthesized via template-engaged reaction as cathodes for high-power lithium ion batteries. Adv. Funct. Mater. 2011, 21, 348–355. [Google Scholar] [CrossRef]
  38. de Beeck, J.O.; Labyedh, N.; Sepúlveda, A.; Spampinato, V.; Franquet, A.; Conard, T.; Vereecken, P.M.; Celano, U. Direct imaging and manipulation of ionic diffusion in mixed electronic-ionic conductors. Nanoscale 2018, 10, 12564–12572. [Google Scholar] [CrossRef] [PubMed]
  39. Balke, N.; Jesse, S.; Morozovska, A.N.; Eliseev, E.; Chung, D.W.; Kim, Y.; Adamczyk, L.; García, R.E.; Dudney, N.; Kalinin, S.V. Nanoscale mapping of ion diffusion in a lithium-ion battery cathode. Nature Nanotechnol. 2010, 5, 749–754. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples.
Figure 1. XRD patterns of LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples.
Materials 11 01302 g001
Figure 2. SEM images of LiCuxMn1.95-xSi0.05O4 samples: (a) x = 0; (b) x = 0.02; (c) x = 0.05; (d) x = 0.08.
Figure 2. SEM images of LiCuxMn1.95-xSi0.05O4 samples: (a) x = 0; (b) x = 0.02; (c) x = 0.05; (d) x = 0.08.
Materials 11 01302 g002
Figure 3. (a) Initial discharge curves and (b) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples; (c) Comparison plots of the initial discharge capacities and capacity retentions; (d) Long Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples.
Figure 3. (a) Initial discharge curves and (b) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples; (c) Comparison plots of the initial discharge capacities and capacity retentions; (d) Long Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples.
Materials 11 01302 g003
Figure 4. (a) Discharge curves of the representative LiCu0.05Mn1.90Si0.05O4 sample at varying rates; (b) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at varying rates; (c) Initial discharge curves and (d) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at 10 C.
Figure 4. (a) Discharge curves of the representative LiCu0.05Mn1.90Si0.05O4 sample at varying rates; (b) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at varying rates; (c) Initial discharge curves and (d) Cycling performance of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at 10 C.
Materials 11 01302 g004
Figure 5. (a) Cycling performance and (b) rate capacities of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at 55 °C.
Figure 5. (a) Cycling performance and (b) rate capacities of the LiCuxMn1.95−xSi0.05O4 (x = 0, 0.05) samples at 55 °C.
Materials 11 01302 g005
Figure 6. Nyquist plots of the LiMn1.95Si0.05O4 (a) and LiCu0.05Mn1.90Si0.05O4 (b) samples before cycling and after 200 cycles; (c) Equivalent circuit model of EIS.
Figure 6. Nyquist plots of the LiMn1.95Si0.05O4 (a) and LiCu0.05Mn1.90Si0.05O4 (b) samples before cycling and after 200 cycles; (c) Equivalent circuit model of EIS.
Materials 11 01302 g006
Table 1. Intensity ratio of (311)/(400) peaks of LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples.
Table 1. Intensity ratio of (311)/(400) peaks of LiCuxMn1.95−xSi0.05O4 (x = 0, 0.02, 0.05, 0.08) samples.
SamplesI(311)/I(400)
LiMn1.95Si0.05O40.98
LiCu0.02Mn1.93Si0.05O41.00
LiCu0.05Mn1.90Si0.05O41.01
LiCu0.08Mn1.87Si0.05O41.03
Table 2. Fitting values of the charge transfer resistance (R2) calculated from EIS.
Table 2. Fitting values of the charge transfer resistance (R2) calculated from EIS.
SamplesR2 (Ω cm2) before CyclesR2 (Ω cm2) after 200 Cycles
LiMn1.95Si0.05O493.2158.1
LiCu0.05Mn1.90Si0.05O470.5116.0

Share and Cite

MDPI and ACS Style

Zhao, H.; Li, F.; Bai, X.; Wu, T.; Wang, Z.; Li, Y.; Su, J. Enhanced Cycling Stability of LiCuxMn1.95−xSi0.05O4 Cathode Material Obtained by Solid-State Method. Materials 2018, 11, 1302. https://doi.org/10.3390/ma11081302

AMA Style

Zhao H, Li F, Bai X, Wu T, Wang Z, Li Y, Su J. Enhanced Cycling Stability of LiCuxMn1.95−xSi0.05O4 Cathode Material Obtained by Solid-State Method. Materials. 2018; 11(8):1302. https://doi.org/10.3390/ma11081302

Chicago/Turabian Style

Zhao, Hongyuan, Fang Li, Xiuzhi Bai, Tingting Wu, Zhankui Wang, Yongfeng Li, and Jianxiu Su. 2018. "Enhanced Cycling Stability of LiCuxMn1.95−xSi0.05O4 Cathode Material Obtained by Solid-State Method" Materials 11, no. 8: 1302. https://doi.org/10.3390/ma11081302

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