Yanzhi Wang*ab and
Jiantao Tanga
aHebei Key Laboratory of Applied Chemistry, College of Environmental and Chemical Engineering, Yanshan University, Qinhuangdao, Hebei 066004, China. E-mail: hhwyz@ysu.edu.cn; Fax: +86 335 8061569; Tel: +86 335 8061569
bState Key Laboratory of Metastable Material Science and Technology, Yanshan University, Qinhuangdao, 066004, China
First published on 3rd July 2018
To improve the cycling stability and dynamic properties of layered oxide cathodes for sodium-ion batteries, surface modified P2–Na0.67Co0.25Mn0.75O2 with different levels of CeO2 was successfully synthesized by the solid-state method. X-ray photoelectron spectra, X-ray diffraction and Raman spectra show that the P2-structure and the oxidation state of cobalt and manganese of the pristine oxide are not affected by CeO2 surface modification, and a small amount of Ce4+ ions have been reduced to Ce3+ ions, and a few Ce ions have entered the crystal lattice of the P2-oxide surface during modification with CeO2. In a voltage range of 2.0–4.0 V at a current density of 20 mA g−1, 2.00 wt% CeO2-modified Na0.67Co0.25Mn0.75O2 delivers a maximum discharge capacity of 135.93 mA h g−1, and the capacity retentions are 91.96% and 83.38% after 50 and 100 cycles, respectively. However, the pristine oxide presents a low discharge capacity of 116.14 mA h g−1, and very low retentions of 39.83% and 25.96% after 50 and 100 cycles, respectively. It is suggested that the CeO2 modification enhances not only the maximum discharge capacity, but also the electric conductivity and the sodium ion diffusivity, resulting in a significant enhancement of the cycling stability and the kinetic characteristics of the P2-type oxide cathode.
Surface modification with the inert/active compound has been proven to be an effective approach in improving the cyclic stability and thermal stability of various electrode materials at high cut-off potentials by preventing the direct contact between active particles and electrolyte, and decreasing the side reaction and elemental dissolution during the cycling.2–4,21–35 At present, several kinds of inorganic materials have been used for surface modification on electrode materials, mainly including metal oxides,4,21–28 metal fluorides,28,29 carbon material,18,32 and other materials.33–35 For instance, Na2/3[Ni1/3Mn2/3]O2 was synthesized by solid state reaction, and delivered an initial capacity of 160 mA h g−1 in a voltage range of 2.5–4.3 V at a current density of 0.5C, and Al2O3 surface modified Na2/3[Ni1/3Mn2/3]O2 delivered a similar initial capacity of 160 mA h g−1, but the capacity retention rate increased from 28.0% to 73.2% after 300 cycles.23 Na0.5Ni0.33Mn0.67O2 presented initial discharge capacity around 119 mA h g−1 within 2.20–4.25 V with the final capacity retention was 70.75% until 100th cycle, and the MgO-coated Na0.5Ni0.33Mn0.67O2 delivered an initial capacity increased to 131 mA h g−1 in the voltage range of 2.0–4.5 V, and the increased capacity retention was 96.45% after 70 cycles.4 However, most of the materials used for modification are insulators, possessing less transportation capability for electrons or ions in the modification layer. It was reported that CeO2 could accelerate electron transfer between CeO2 oxide and the supported active material, and supplies a good conductive connection between the active particles and decreases the charge transfer resistance and electrode polarization, and then improves the electrochemical performance of electrode materials.28,36–38 For example, the sample Li4Ti5O12 presented capacity ∼80 mA h g−1 at 10C rate with the capacity retention of only 40% after 50 cycles, the sample Ce and CeO2 modified Li4Ti5O12 with Ti/Ce = 4.85:0.15 delivered a capacity of 161 mA h g−1 at same rate with the capacity retention of 100% after 50 cycles.36 At 5C charge–discharge rate, the pristine Li5Cr7Ti6O25 only exhibited an initial charge capacity of ∼94.7 mA h g−1, and the capacity only maintained 87.4 mA h g−1 after 100 cycles. However, Li5Cr7Ti6O25@CeO2 (3 wt%) presented an initial charge capacity of 107.5 mA h g−1, and the capacity also reached 100.5 mA h g−1 even after 100 cycles.37 In addition, compared with the pristine LiNi0.5Mn1.5O4, the cathode material LiNi0.5Mn1.5O4–CeO2 (3 wt%) exhibited outstanding discharge capacity, cycling stability and rate capability.28 D. Arumugam et al. investigated nano-CeO2-coated LiMn2O4 cathode for rechargeable lithium ion batteries, and suggested that surface modification was an effective way to improve the chemical stability of the electrode and their cyclability and rate capability during long-term cycles.39 M. Michalska et al. demonstrated that surface modification with 1 wt% CeO2 improved cycle stability of LiMn2O4, and CeO2 acted as both conductive additive, increasing the electrical conductivity of the material, and as a medium, preventing a material from degradation during cycles, and the increase in cycle life may be due to CeO2 limiting the dissolution of manganese into electrolyte through reduced contact area.40 Some researchers studied that the structures, and cyclic stability and rate performance of CeO2-coated Li[Ni0.5Co0.2Mn0.3]O2 samples, and revealed that the enhanced electrochemical characteristics could be attributed to the decrease of the interfacial polarization and stabilization of the oxide structure by CeO2 coating.41,42 Y. Wu, et al. modified (1 − z) Li[Li1/3Mn2/3]O2 − (z) Li[Mn0.5−yNi0.5−yCo2y]O2 layered solid solutions with 3 wt% Al2O3, CeO2, ZrO2, SiO2, ZnO, respectively, and demonstrated that surface modification may be an effective way to decrease the irreversible capacity and maximize the discharge capacity for the layered oxide cathodes.43 W. Yuan et al. modified the Li-rich layered oxide Li(Li0.17Ni0.2Co0.05Mn0.58)O2 with hexagonal α-NaFeO2 structure by using CeO2 nanocrystallites, and the CeO2-modified layered oxide presented higher discharge capacity, lower charge transfer resistance, and larger initial coulombic efficiency than the pristine oxide.44 However, we have not found the modification of P2 oxide with CeO2.
Herein, we have synthesized Na0.67Co0.25Mn0.75O2 oxide using a solid-state method, and the oxide was subsequently surface-modified with CeO2 by the same method. Microstructure analyses indicate that the P2-type structure of the pristine oxide has not been changed by CeO2 surface modification, and a small amount of Ce4+ ions have been reduced to Ce3+ ions, and which might have entered the crystal lattice of the P2-oxide during the synthesis. 2.00 wt% CeO2-modified Na0.67Co0.25Mn0.75O2 exhibits a maximum discharge capacity of 135.93 mA h g−1 with the retention of 91.96% and 83.38% after 50 and 100 cycles in a voltage range of 2.0–4.0 V at a current density of 20 mA g−1. Besides, the kinetic characteristics of P2-type oxide cathode have been improved by CeO2 modification.
Fig. 1 XPS patterns of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites, (a) Mn 2p, (b) Co 2p, (c) Ce 3d. |
Fig. 2a indicates the XRD patterns of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites. As can be seen, each sample presents the strongest (002) peak of a layered P2-type oxide, and all the main diffraction peaks can be indexed into a hexagonal P63/mmc (P2-type) phase (JCPDS no. 54-0894).9–11 Besides the main P2-type phase, additional diffraction peaks are also observed in the XRD patterns of the CeO2-modified NCMO composites, which attributes to CeO2, and the intensity of the additional peak increases with. the increasing CeO2 amount. The lattice parameters of NCMO and CeO2-modified NCMO composites by the Rietveld refinement are tabulated in Table 1. It presents that the lattice parameters of the main phases for CeO2-modified NCMO composites are larger than that of the pristine NCMO, which confirms that a few Ce ions have entered the composite crystal lattice after surface modification with CeO2 because of Ce ion with a larger size than that of Mn ion. This is consistent with the result from Ce 3d XPS spectra. The similar phenomenon also appeared in some literatures,38,46 for example, Li4Ti5O12 was doped with various dopping level Ce3+ accompanied by the occurrence of CeO2 surface modification.38 The dopping of Ce4+ and the combination with in situ generated CeO2 in Li4Ti5O12 were favorable for improving the lithium insertion/extraction kinetics of Li4Ti5O12.36 It was reported that a few Ce4+ ions entered the crystal lattice of Li5Cr7Ti6O25 and then increased the lattice parameter during CeO2 coating, and Li5Cr7Ti6O25@3wt.%CeO2 electrode showed a better electrochemical activity than that of the pristine Li5Cr7Ti6O25.37 In general, the a-axis expansion will lead to the better stability of P2-type material, and the c-axis expansion will lead to enlarged path and increased rate of Na+ migration in the material, which illustrates that both the stability and dynamics of NCMO cathode would be enhanced by Ce ion dopping during CeO2 surface modification. In addition, it can also be seen that the lengths of a-axis and c-axis for NCMO-2.00 wt% CeO2 material are the largest among the four samples, respectively, which indicates that this composite would exhibit the best cyclability and dynamics as a cathode of SIBs among the four oxides.
Sample | a (Å) | b (Å) | c (Å) | RP | RWP |
---|---|---|---|---|---|
NCMO | 2.848 | 2.848 | 11.166 | 9.743 | 11.342 |
NCMO-1.00wt% CeO2 | 2.865 | 2.865 | 11.242 | 9.792 | 11.324 |
NCMO-2.00wt% CeO2 | 2.879 | 2.879 | 11.324 | 9.412 | 11.317 |
NCMO-3.00wt% CeO2 | 2.866 | 2.866 | 11.271 | 9.691 | 11.237 |
To further study the structural effect of CeO2 surface modification on NCMO sample, Raman spectra have been measured. Fig. 2b exhibits the Raman spectra of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites. It presents that the two peaks between 400 cm−1 and 600 cm−1, corresponding to the active E2g modes due to Na and O vibrations, and a shoulder peak at around 637 cm−1, which can be assigned to A1g mode due to O vibrations.16,47 These peaks further demonstrate that the layered P2-type structure has not been changed after CeO2 surface modification, which is highly matched to the above XRD results.
Fig. 3 shows FESEM images of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites. It is clear that all samples have similar flake shapes, in which the surface of NCMO sample is smooth, and there are some particles in the surface of CeO2-modified NCMO composites. Fig. 4 indicates EDS elementary mapping images of NCMO-x wt% CeO2 (x = 0, 2.00) composites. The elements of Na, Co, Mn, O and Ce are uniformly distributed in the samples. TEM images of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites are shown in Fig. 5. As can be seen, the surface of NCMO sample in the three composites has been successfully modified by CeO2, which can effectively prevent the aggregation between the composites particles. To clarify the effects of the CeO2 surface modification on the kinetic properties of NCMO electrode, the electrochemical impedance spectra of the four samples were performed on fresh cells at the frequency range from 100 kHz to 10 mHz. The Nyquist plots of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) samples are given in Fig. 6a. The EIS plots are composed of two well-separated semicircles at high and intermediated frequency region and an oblique line in the low-frequency region. In the equivalent circuit shown in the inset of Fig. 6a, the electrolyte resistance (Rs) is mainly related to the high-frequency intercept at the Zreal axis, the SEI film resistance (Rf) originates from the semicircle in the high frequency region, and the charge-transfer resistance (Rct) is related to the second semicircle in the middle frequency, which reflects the intercalation kinetics of the electrodes, and Warburg impedance (W) is connected with the slope in the low-frequency region, reflecting the intra-electrode mass diffusion.15,16 The EIS analysis results of the four samples are listed in Table 2. As shown, the small change in Rs values between 7.1 and 17.2 could have no significant effect on the dynamic characteristics of electrodes. Before increasing CeO2 content up to 2.00 wt%, the Rct values for NCMO-x wt% CeO2 electrodes decrease notably from 1589.0 Ω (x = 0) to 149.2 Ω (x = 2.00), and then increase to 562.7 Ω (x = 3.00) as x further increases. NCMO-2.00 wt% CeO2 composite attains the best conductivity. It is suggested that the appropriate CeO2 surface modification is beneficial for the improvement of the charge transportation between the NCMO electrode and the electrolyte interface, and then enhances conductivity of the active materials during cycling.28 The reasons may be as follows: crystal lattice defects may be induced by Ce ions entering NCMO lattice, which is beneficial to the charge transfer inside NCMO lattice. Besides, CeO2 with a high electrical conductivity enhances the total electrical conductivities of the composite electrodes.36 Compared to Rct change, the values of Rf for the four electrodes have the similar rules, which suggests that the electrolyte decomposition, disadvantageous SEI film, and the Mn3+-ion dissolution during the cycling process can be suppressed by appropriate CeO2 modification.28 This will be beneficial to enhance the cycling stability of the composites. In addition, the sodium ion diffusion coefficients (D) can be calculated from the oblique line in the low-frequency region according to the eqn (1):16,36
D = R2T2/(2A2n4F4C2σ2) | (1) |
Fig. 3 FESEM images of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites, (a) x = 0, (b) x = 1.00, (c) x = 2.00, (d) x = 3.00. |
Fig. 4 EDS elementary mapping images of NCMO-x wt% CeO2 (x = 0, 2.00) composites, (a) x = 0, (b) x = 2.00. |
Fig. 5 TEM images of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites, (a) x = 0, (b) x = 1.00, (c) x = 2.00, (d) x = 3.00. |
Fig. 6 (a) EIS Nyquist plots of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) composites and the equivalent circuit in inset, and (b) their linear fitting of the Zreal versus ω−1/2. |
Cathode | Impedance parameters (Ω) | Na+ diffusion coefficient (×10−13, cm2 s−1) | ||
---|---|---|---|---|
Rs | Rf | Rct | ||
NCMO | 13.1 | 1484.0 | 1589.0 | 2.68 |
NCMO-1.00 wt% CeO2 | 11.2 | 1363.0 | 1225.0 | 8.21 |
NCMO-2.00 wt% CeO2 | 7.1 | 853.7 | 149.2 | 8.77 |
NCMO-3.0 wt% CeO2 | 17.2 | 902.6 | 562.7 | 8.30 |
Here, R, T, A, n, F and C are fixed values, which corresponds to the gas constant, the absolute temperature, the surface area of the electrode, the number of electrons per reaction, the Faraday constant, and the concentration of Na+ in the material, respectively. Warburg factor σ can be related to Zreal as the eqn (2):16,28
Zreal = Rs + Rct + σω−1/2 | (2) |
In this equation, ω is the angular frequency, and σ value can be obtained from the slope of the fitting line in the Fig. 6b. The calculated results are also summarized in the Table 2. The apparent Na+ diffusion coefficients in NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) electrodes significantly increase from 2.68 × 10−13 cm2 s−1 to 8.77 × 10−13 cm2 s−1 as x increases up to 2.00, and then decrease to 8.30 × 10−13 cm2 s−1 with further increasing x to 3.00. The above composite electrodes have higher apparent sodium-ion diffusion coefficients than that of the pristine one, indicating that the former has higher electrochemical activity during cycling. These results assuredly indicate that the sodium-ion transfer ability in NCMO electrode can be effectively improved by CeO2 modification.36,37 Apparently, the high Na+ diffusion coefficient for the CeO2 modified NCMO electrode should be attributed to the stable structure and wider track of Na+ migration. Among the four samples, NCMO-2.00 wt% CeO2 has the highest sodium-ion diffusion, which is consistent with the longest length of c-axis. From the above findings, the amount of CeO2 must be properly optimized to improve sodium-ion migration and enhance rate performance. Moreover, NCMO-1.00 wt% CeO2 and NCMO-3.00 wt% CeO2 composites show shorter c-axis than that of the pristine NCMO, whereas the former shows bigger apparent Na+ diffusion coefficients than that of the pristine one. This can be explained by the effect of Ce ions entering into the lattice of NCMO material, which has induced the lattice defects, and then provides more paths to facilitate the sodium ions to diffuse in the composites. However, too much CeO2 modification on the NCMO particle surface may suppress the migration of sodium ions to the interface,37 which is unfavorable for the improvement of the kinetic characteristics.
Fig. 7 shows the galvanostatic charge and discharge curves of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) cathodes at a current density of 20 mA g−1. It can be seen that all the discharge curves resemble slopping curves. This suggests that the principal mechanism of the charge/discharge process is based on a solid-solution process, which is evidence of the structural stability of the materials. It is evident that all the charge curves show similar shapes, two well-defined regions in the each curve can be observed, an oblique plateau at about 2.4 V, which is attributed to the Mn4+/Mn3+ and (or) Co3+/Co2+ redox couples, and another a visible short-flat voltage at around 3.5 V, corresponding to Co4+/Co3+ redox couples.13,14,16,17 This is in accordance with the redox peaks in the above CV profile. The above phenomenon indicates that the principal mechanism in the sodiation/desodiation process for the material has not been changed by CeO2 surface modification. In addition, surface modification of CeO2 remarkably decreased the initial charge specific capacity because more vacancies have probably been induced by CeO2 modification to provide sufficient active sites for sodium ions insertion in the subsequent discharge process.44 The maximum discharge capacity of NCMO-x wt% CeO2 cathode enhances from 116.14 mA h g−1 to 135.93 mA h g−1 as CeO2 content increases up to 2.00 wt%, and then decreases to 122.88 mA h g−1 with increasing CeO2 level to 3.00 wt%, which indicates that moderate Ce ions dopping and CeO2 modification can enhance the discharge capacity of NCMO at a low rate.28,37 We suggested that Ce ions dopping is more effective than CeO2 modifying for increasing the maximum discharge capacity. The cathodes NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) attain their maximum discharge capacity at 1, 2, 5 and 4 cycles, respectively, which indicates that CeO2-modified NCMO cathodes need a few activation numbers. This may be due to the fact that CeO2 has no electrochemical activity, and electrolyte decomposition and the formation of an SEI (solid-electrolyte interface) layer, causing irreversible capacity loss during the initial cycles, were suppressed on the CeO2-modified NCMO surface.39 As is known to all, the electrode materials synthesized by the solid-state method have a lower homogeneity than that of the materials prepared by a soft chemical method. However, the composites NCMO-x wt% CeO2 cathodes also show high discharge capacities. This indicates that moderate CeO2 modification can enhance the discharge capacity of NCMO material.28,36
Fig. 7 The galvanostatic charge and discharge curves of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) cathodes, (a) x = 0, (b) x = 1.00, (c) x = 2.00, (d) x = 3.00. |
Fig. 8a shows CV curves of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) electrodes. It reveals that there are the similar overall profiles with multiple peaks within 2.0–4.0 V, which confirms the high similarity of redox reaction mechanism in the process of Na+ intercalation/deintercalation. These peaks can be divided into two groups, the peaks at low voltages (<3.0 V vs. Na+/Na) can be associated with the redox process of Mn4+/Mn3+ and (or) Co3+/Co2+, and the peaks between 3.0 to 4.0 V can be can be attributed to Co4+/Co3+ redox couple.16,17 Obviously, the area of CV curve for the composite NCMO-2.0 wt% CeO2 is the largest, which implies that the composite would have the highest specific capacity among the four samples. This is highly matched to the above charge and discharge curves.
Fig. 8 (a) Cyclic voltammogram curves, (b) cycling stability, (c) the coulombic efficiencies, and (d) the rate performance of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) cathodes. |
Fig. 8b and c present the cycling performance and the corresponding coulombic efficiencies of the four cathodes at a current density of 20 mA g−1, respectively. As shown, the capacity retentions of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) cathodes after 50 cycles are 34.29%, 79.05%, 91.96% and 77.46%, respectively, and after 100 cycles, the retentions of the four cathodes are accordingly 25.96%, 68.03%, 83.30% and 68.36%, respectively, in which NCMO-2.0 wt% CeO2 cathode exhibits the best cyclic stability. The above results confirm that the cycling stability of NCMO cathode can be evidently enhanced by CeO2 surface modification. The reasons for the improvement are as follows: first of all, the a-axis expansion leads to the better cycling stability of NCMO cathode by Ce ions dopping during CeO2 surface modification. Secondly, Mn3+ in NCMO cathode from the charge compensation easily causes Jahn–Taller distortion during cycles,15–19 which may be suppressed by proportionate CeO2 modification. Thirdly, Mn3+ is easy to dissolve into electrolyte during charge and discharge of the NCMO material, which is probably suppressed by optimal CeO2 modification.28,36 Fourthly, as mentioned above, the SEI film resistance for NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) electrodes decreases from 1484.0 Ω (x = 0) to 853.7 Ω (x = 2.00) and then increases to 902.6 Ω (x = 3.00) with increasing CeO2 content, and reaches the lowest values at 2.00 wt% CeO2, which is another evidence for improving the cyclic stability of the pristine NCMO material by CeO2 surface modification. Moreover, the coulombic efficiencies are close to 100% except for the initial cycles for the CeO2-modified NCMO materials, showing good cycling stability. We suggested that CeO2 modifying is more effective than Ce ions dopping for improving cycle stability.
Fig. 8d exhibits that the rate performance of NCMO-x wt% CeO2 (x = 0, 1.00, 2.00 and 3.00) cathodes. As can be seen, the rate performances of CeO2-modified NCMO cathodes significantly outperform that of the pristine NCMO. For example, at a current density of 500 mA g−1, the mean discharge capacity of NCMO-2.00 wt% CeO2 composite is about 7 times as much as that of the pristine oxide. In addition, NCMO-2.00 wt% CeO2 cathode still delivers a high mean discharge capacity of 100.64 mA h g−1 at a current density of 200 mA g−1, and reducing the current density from 500 mA g−1 to 20 mA g−1, the average discharge capacity of the each CeO2-modified NCMO cathode is close to its maximum discharge capacity. These results imply that CeO2 surface modification is conducive to the reversible intercalation and deintercalation of Na+ in NCMO material. This is probably attributed to the protection of CeO2 particles on the surface of NCMO, suppressing the structural damage at high current density,36,48 and the c-axis expansion, leading to enlarged path and increased rate of Na+ migration in the material, and the decrease of the charge-transfer resistance and the increase of the apparent Na+ diffusion coefficients, improving the electrode dynamics during cycling. In addition, it is reported the internal ions adsorption on the CeO2 surface may bring about the space-charge effect,49 increasing the positive ion vacancy concentration at the CeO2 interface, which results in the wonderful electrical contact, and the good transmission efficiency of electrons and Na+ between the NCMO and CeO2, and then obviously enhances the desodiation and sodiation capacity and cycling stability at high current density because of CeO2 modification.37,38 We suggested that Ce ions dopping as important as CeO2 modifying for improving the rate performance.
This journal is © The Royal Society of Chemistry 2018 |