Mitigating the Surface Degradation and Voltage Decay of Li1.2Ni0

May 25, 2017 - Functional Materials Division, CSIR − Central Electrochemical Research Institute, Karaikudi 630003, India. ∥ Surface Engineering Di...
0 downloads 11 Views 8MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Mitigating the Surface Degradation and Voltage Decay of Li1.2Ni0.13Mn0.54Co0.13O2 Cathode Material through Surface Modification Using Li2ZrO3 Kunkanadu R. Prakasha,†,‡ Marappan Sathish,§ Parthasarathi Bera,∥ and Annigere S. Prakash*,† †

CSIR − Network Institutes of Solar Energy (CSIR − NISE) and ‡Academy of Scientific and Innovative Research (AcSIR), CSIR − Central Electrochemical Research Institute-Chennai Unit, CSIR Madras Complex, Taramani, Chennai 600113, India § Functional Materials Division, CSIR − Central Electrochemical Research Institute, Karaikudi 630003, India ∥ Surface Engineering Division, CSIR − National Aerospace Laboratories, Bengaluru 560017, India S Supporting Information *

ABSTRACT: In the quest to tackle the issue of surface degradation and voltage decay associated with Li-rich phases, Li-ion conductive Li 2 ZrO 3 (LZO) is coated on Li1.2Ni0.13Mn0.54Co0.13O2 (LNMC) by a simple wet chemical process. The LZO phase coated on LNMC, with a thickness of about 10 nm, provides a structural integrity and facilitates the ion pathways throughout the charge−discharge process, which results in significant improvement of the electrochemical performances. The surface-modified cathode material exhibits a reversible capacity of 225 mA h g−1 (at C/5 rate) and retains 85% of the initial capacity after 100 cycles. Whereas, the uncoated pristine sample shows a capacity of 234 mA h g−1 and retains only 57% of the initial capacity under identical conditions. Electrochemical impedance spectroscopy reveals that the LZO coating plays a vital role in stabilizing the interface between the electrode and electrolyte during cycling; thus, it alleviates material degradation and voltage fading and ameliorates the electrochemical performance.



INTRODUCTION Lithium-rich phases have been widely investigated as prospective high-capacity cathode materials for Li-ion batteries to be used in plug-in hybrid electric vehicles or electric vehicles. In particular, the lithium-rich transition metal oxide with the general formula of xLi2MnO3·(1 − x)LiMO2 (M = Mn, Co, Ni) is a significant milestone in materials design owing to its wellbalanced capacity/power/cost, which is tailored for high operating voltage in comparison with the existing state of the art commercial cathode materials, such as LiCoO2, LiMn2O4, and LiFePO4.1−5 Yet, there are vital snags which impede the commercial progress, such as (i) large irreversible capacity loss in the first cycle, (ii) inevitable phase transformation from layered to spinel phase leading to continuous decay in the discharge capacity,6−9 (iii) oxygen loss from the crystal lattice which triggers serious safety concerns and low initial Coulombic efficiency, (iv) structural instability due to Mn dissolution, (v) voltage decay,10−12 (vi) weak interfacial stability between the electrode and electrolyte, and (vii) thermal stability. Recently, several strategic approaches have been attempted to tackle the associated shortcomings of lithium-rich layered cathodes by adopting procedures such as surface coating, acid treatment, anion/cation doping,13,14 innovative synthesis method,15,16 substitution of heavy and bulky elements like Sn, Ti, Mo, and so forth.17,18 © 2017 American Chemical Society

All of these techniques have beefed up the commercialization aspects of Li-rich layered cathode materials. Among all strategies, surface modification has proven to be an effective method to surmount the shortcomings of Li-rich phases. Surface coatings with metal oxides,19−23 metal phosphates,24,25 and metal fluorides26 have thus been pursued to enhance the physical and chemical properties of lithium-rich cathode materials. Such surface modification improves the cycling stability, rate capability, and thermal stability of cathodes, as a result of suppression of oxygen-ion vacancies in the crystal lattice at the end of the first charge. Further, surface coating suppresses the undesired electrochemical reactivity of the electrolyte at the electrode/electrolyte interface by preventing direct contact between the electrode and electrolyte. However, most of these coating materials do not favor interfacial charge transfer of the electrode material; further, they act as a barrier for Li+ conduction and increase the diffusion path length of Li ion, resulting in inferior rate capability. To overcome the limitation of the impeded lithium ion diffusion through the surface barrier, lithium-ion conducting material has been introduced as a protective layer on lithiumReceived: March 30, 2017 Accepted: May 11, 2017 Published: May 25, 2017 2308

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

that there is no severe change in the host layered structure of LZO @ LNMC, except for some peaks that are enlarged in the inset of Figure 1b. The peaks marked in “*” are the characteristics of LZO corresponding to the monoclinic system with C2/c space group, which is consistent with the JCPDS card: 01-070-8744. Figure 2 shows the Rietveld refinement of XRD patterns for surface-modified and pristine material using Fullprof software. The lattice parameters (a, b, and c) of the surface-modified sample are same as those of the pristine pattern, suggesting that the surface coating with LZO does not cause any undue bulk structural changes. The structural parameters determined by Rietveld refinement for both pristine and coated samples using R3̅m and C2/m phases are listed in Tables S1 and S2, respectively. FESEM images of bare and 5 wt % LZO-coated LNMC particles are displayed in Figure 3a,b. No significant difference is observed in the morphology and grain size between pristine and LZO @ LNMC, whereas a closer observation of the surface reveals the relatively smooth grains for the sample of surfacemodified LZO. EDS has been employed to elucidate the uniform distribution and analyze the elemental composition of the surface-modified sample. The atomic ratio of Ni/Mn/Co/ Zr is 15.11:63.4:15.35:6.03, which is in concordance with that of the targeted composition of the phase, that is, 0.12:0.51:0.12:0.05, respectively. Elemental mapping in Figure 3d−h of the surface-modified sample clearly demonstrates the uniform distribution of zirconium element on the surface. HRTEM studies have been carried out to acquire further information pertaining to the interface between the LZO coating and bulk LNMC material (Figure 4a,b). Both the figures correspond to high-resolution transmission electron microscopy (HRTEM) images of the rectangle tagged interfacial area of their inset figures. The images show uniform surface layers of thickness of around 10 nm. In Figure 4a, the lattice spacings of 4.7 nm for bulk and 4.3 nm for the surface coating are in good concurrence with (003) and (110) interplanar spacing of isostructural hexagonal LNMC and tetragonal LZO, which is complemented by the XRD results. The (110) plane of LZO, oriented along the direction of (003) plane of LNMC shows perfect epitaxial alignment. Similarly, in Figure 4b, the (112) plane of LZO are oriented around 60° with respect to the (003) plane of the core LNMC lattice. This cross-wire epitaxial arrangement results in negligible interfacial lattice mismatch {(4.7 − 1.99/cos 60)/ 4.7 = 15%}. Such a slant epitaxial growth in the nanoheterostructure is reported for other systems.38,39 The lattice mismatch calculation used here is adopted from the previous works.38,39 The epitaxial growth with the preferred orientation and alignment of LZO on LNMC provides structural integrity and possesses a Li-ion diffusion pathway. X-ray photoelectron spectroscopy (XPS) of Li1s, Ni2p, Co2p, and Mn2p core levels in the pristine and coated samples have been carried out to understand the oxidation states of these elements in these materials. Ni2p, Co2p, Mn3p, Zr3d, O1s, and C1s core-level spectra are shown in Figure 5. The Li1s peak observed at 55.5 eV (given in Figure S1) in both samples corresponds to Li+ species.40 The Ni2p3/2 peak at 854.6 with the satellite peak at 860.7 eV is attributed to Ni2+ species.41 Co2p3/2,1/2 peaks at 780 and 794.9 eV, respectively, can be assigned to either Co2+ or Co3+ species. However, the presence of a weak satellite peak at 789.6 eV related to the 2p3/2 main peak, with a distance of around 9 eV, explicitly confirms the presence of Co3+ in both of the samples.42,43 Mn2p3/2,1/2 core

rich phases. This ion-conducting layer facilitates lithium-ion transportation between the cathode material and electrolyte; it also stabilizes their interface at high voltages.27,28 In recent years, Li2ZrO3 (LZO) coating has gained immense interest for its role in the stability of cathodes.29−31 In this context, Wu et al. have demonstrated a high rate and long cycle life performance of the LZO-coated LiNi0.5Co0.2Mn0.3O2 cathode.32,33 In the present article, we envisage the LZO coating to tackle specific issues associated with Li-rich layered phases, such as O2 evolution, reactivity of evolved O2 with the electrolyte, Mn dissolution, and phase transformation. The monoclinic LZO is structurally related to the layered Li1.2Ni0.13Mn0.54Co0.13O2 (LNMC). Hence, it is anticipated to enhance the structural integrity owing to an epitaxial relation between the host cathode and the Li-ion conducting surface layer. This epitaxial interface may play a beneficial role and effectively separate the bulk electrode from the electrolyte and stabilize the structure and the composition during high-voltage cycling. This article reports a systematic investigation of the role of the Li-ion conducting LZO surface layer in overcoming the issues with LNMC.



RESULTS AND DISCUSSION Figure 1a,b compare the X-ray diffraction patterns of bare LNMC and LZO surface-modified LNMC (LZO @ LNMC).

Figure 1. Powder XRD patterns recorded with Cu Kα radiation for (a) pristine LNMC, (b) 5% LZO @ LNMC. The diffraction peaks in the inset figure marked with “*” correspond to LZO.

Both the patterns are indexed as hexagonal α-NaFeO2 structures with R3̅m space group. The well-defined splitting of the reflections, (006)/(102) and (108)/(110), manifests in the crystallization of the layered structure, without formation of the spinel phase. The weak reflections appearing between 20 and 30° (2θCu) can be indexed considering a √3ahex × √3bhex superstructure in the ab 3d metal planes as a result of an ordering between the Li, Ni, Co, and Mn ions similar to the monoclinic Li2MnO3 structure (represented in the C2/m space group).35,36 The integrated intensity ratio of the (003) to (104) reflections used for quantification of cation mixing in the lattice being 1.40, which is greater than 1.2, provides evidence of the lesser degree of cation mixing in the Li layer, whereas in the case of the surface-modified pattern, the I(003)/I(104) ratio increases to 1.47.37 This indicates that the degree of cation mixing depreciates by surface modification. Further, it suggests 2309

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

Figure 2. Rietveld refinement pattern of powder XRD data for (a) LNMC, and (b) 5% LZO @ LNMC, with the experimental data (black dots), calculated pattern (red line), difference curve (blue line), and Bragg diffraction positions (yellow ticks for R3m ̅ space group and green ticks for C2/m space group).

Figure 3. FESEM images of (a) bare LNMC, (b) 5% LZO-coated LNMC and (c) EDAX profile, (d−h) elemental mapping images of 5% LZOcoated sample.

current rate in the potential range of 1.75−4.8 V are shown in Figure 6a. The charge profile is composed of a sloping region below 4.4 V, followed by a long plateau region around 4.5 V. Both of the samples show almost the same specific capacity in the sloping region, which corresponds to the oxidation of the transition metals, such as Ni2+ to Ni4+ and Co3+ to Co4+. The specific capacity of the plateau region begins at around 4.5 V, which is accompanied by the extraction of Li from the 3d metal layer and also the removal of lattice oxygen. It is to be noted here that the LZO-coated cathode exhibits a lower amplitude in the plateau region. This difference in capacity is speculated to be attributed to the slump in the oxygen loss from the lattice. The pristine Li-rich NMC exhibits an initial charge capacity of 329 mA h g−1, whereas the 5% LZO-coated LNMC shows initial charge capacity of 306 mA h g−1, part of this decrease in capacity is also attributed to the presence of 5% (in weight) electrochemically inactive LZO in the electrode material. The bare LNMC exhibits a discharge capacity of 234, with 95 mA h g−1 of irreversibility, whereas the 5% surface-modified sample exhibits an initial capacity of 225 mA h g−1, with 80 mA h g−1 irreversibility. This decrease in irreversible capacity is mainly

level peaks at 642.4 and 654.0 eV in the coated and uncoated samples are assigned to Mn4+ species, respectively.41 Zr3d5/2,3/2 peaks at 182.3 and 184.4 eV in the surface-modified sample shown in Figure 5 are associated with Zr 4+ species, respectively.44 The envelopes of O1s core-level spectra in both the samples are broad, which are deconvoluted into three peaks corresponding to the oxide-adsorbed hydroxyl and water molecules. The main peak at 530.1 eV stands for oxide species, whereas the higher binding energy peaks at 531.8 and 533.9 eV are related to OH− and H2O species, respectively.45 The LZOcoated sample shows a relatively large amount of adsorbed OH− and H2O species compared to that in the uncoated sample. Similarly, C1s spectra shown in the figure are broad in nature and can be deconvoluted into several component peaks. The sharp peak appearing at 284.5 eV corresponds to adventitious carbon species. The broad peak at 286.5 eV is ascribed to oxidized carbon species containing C−O bonds. The third peak at 289.4 eV is assigned to the surface-adsorbed carbonate group. The capacity versus voltage profile for the Li half-cells of the pristine and 5% LZO @ LNMC cathodes cycling at C/5 2310

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

Figure 4. (a, b) High-resolution transmission electron microscopy images showing epitaxial orientation of LZO on LNMC particles.

attributed to the suppression of solid-electrolyte interface (SEI) formation by the LZO surface barrier. To evaluate the effect of the LZO coating on LNMC, charge−discharge cycles have been conducted continuously at a current rate of C/5 between 1.75 and 4.8 V. Figure 6b shows the cycle number versus specific discharge capacity for pristine LNMC and LZO-coated LNMC samples. The surface-modified cathode shows a superior retention of 85% after 100 cycles. In contrast, the bare sample shows a rapid decline in the specific discharge capacity, with retention of 57% after 100 cycles under identical conditions. The rate capability test has been carried out for both the samples at different current rates. It can be seen from Figure 6c that the LZO surface-modified cathode delivers excellent discharge capacities of 225, 196, 160, 100, and 55 mA h g−1 at C/5, C/2, 1C, 2C, and 3C rates, respectively. On the other hand, the pristine material delivers 234, 170, 125, 60, and 30 mA h g−1 at C/5, C/2, 1C, 2C, and 3C rates, respectively. This large difference in rate performance is mainly attributed to the existence of fast lithium-ion diffusion pathway in the LZO surface barrier. The comparison of the discharge voltage profile as a function of the cycle numbers for pristine and LZO surface-modified LNMC is displayed in Figure 6d. There is an abrupt decline in the discharge capacity as well as voltage fading for uncoated

Figure 5. XPS spectra of pristine and LZO @ LNMC samples: Ni2p, Co2p, Mn2p, and Zr3d core levels and deconvoluted O1s and C1s core levels.

LNMC due to the surface deterioration owing to the side reactions at the electrode and electrolyte interface.46,47 On the other hand, the surface-modified cathode exhibits a fairly stable discharge capacity and comparatively lower voltage decay. The differential capacity (dQ/dV) plots of both materials are shown in Figure 6e. The comparison of the dQ/dV plots gives a better understanding of voltage fading. The peak seen in the dQ/dV plot above 4 V corresponds to reduction of Ni4+/Ni2+, the peak around 3.7 V corresponds to Co4+/Co3+, and the peak below 3.5 V is ascribed to the reduction of Mn4+ and of oxygen.48,49 The uncoated LNMC shows that the peaks below 3.5 V correspond to Mn reduction shifting toward lower potential, also a new peak starts to evolve at lower potential. This voltage hysteresis is attributed to the gradual accumulation of spinel environment in the crystal structure. The layered to spinel phase transformation is attributed to decrease in the Mn 2311

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

Figure 6. (a) First cycle charge−discharge profiles of the bare and 5% LZO @ LNMC, (b) cycling performance of bare and 5% LZO @ LNMC, (c) rate capability plot of bare and surface-modified LNMC, (d) discharge profiles at C/5 rate, and (e) corresponding differential capacity curves of the pristine and 5% LZO @ LNMC samples in the voltage range of 1.75−4.8 V.

Figure 7. Nyquist plots of (a) bare LNMC, (b) 5% LZO-coated sample and (c, d) their corresponding enlarged plots, respectively; (e, f) plots of the real part of impedance [Z] as a function of the inverse square root of the angular frequency [ω−1/2].

content in the structure due to dissolution.6,8 In contrast, the structural integrity between LZO coating and the surface of cathode impede the diffusion of Mn and hence limit the layered to spinel phase transformation. Electrochemical impedance spectroscopy (EIS) has been employed to elucidate the improved performance of the surface-modified cathode. Figure 7a,b displays the Nyquist plots

for fresh electrode and cycled electrode at various cycles of pristine and LZO-coated LNMC half-cells. The inflated high frequency region of Figure 7a,b are shown in Figure 7c,d respectively. As seen from the plots, the semicircle in the mid frequency region is merged with that of the high frequency region. This indicates the overlapping of the SEI layer resistance and charge-transfer resistance. The sloping region 2312

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

at low frequencies represents the Warburg resistance, which is attributed to Li+ diffusion through the bulk electrode. The EIS data have been simulated by Zsimpwin software using the equivalent circuit shown in the inset of Figure 7a. In the equivalent circuit, Rs represents the solution resistance, Rsl is the resistance of the SEI layer, Rct is the charge-transfer resistance, W represents the Warburg resistance, and CPE1 and CPE2 represent the double layer capacitance and capacitance of the passivation film, respectively. The values of Rs, Rsl, and Rct obtained from the simulation of EIS data are depicted in Table S3. The solution resistance (Rs) is negligible and almost same for all of the samples. Rct value of the pristine electrode before cycling is 196.5 Ω, and it drastically increases to 1396 Ω after 80 cycles. The reason for such a significant hike in the chargetransfer resistance is attributed to the undesirable reaction between electrode/electrolyte ensuing a slump in surface charge transfer, which is confirmed by increasing the SEI layer resistance (Rsl) from 324.2 to 809.7 Ω after 80 cycles. The charge-transfer resistance of the LZO-coated sample before cycling is 379.7 Ω, which is higher compared to that of the uncoated sample probably because of the presence of surfaceadsorbed species, which are revealed from the XPS and also IR studies (Figure S2). The adsorbed species are removed during the first charge, resulting in drop of the Rct value to 121.3 Ω. The Rct and Rsl values together increase as a function of cycles but the increasing trend is much lower than that of the pristine sample. The increase in Rsl and Rct in the uncoated sample correlates well with the capacity fading observed during cycling. The loss of lithium content is attributed to the formation of undesirable SEI on the surface. On the contrary, the electrochemically passive LZO surface barrier provides greater benefits by protecting the active sites on the surface of LNMC from acidic electrolyte, induced with oxygen at high voltage. Also, the Li+ conducting surface layer facilitates diffusion of lithium and hence a better rate performance. In addition, lithium-ion diffusion coefficients for both LNMC and LZO @ LNMC have also been explored from the same set of EIS data using the following equations50 DLi + = R2T 2/(2A2 n 4F 4C 2σ 2)

Figure 8. Calculated Lithium-ion diffusion coefficient (DLi+) for LNMC and LZO @ LNMC at various cycles, from electrochemical impedance spectroscopy.

10−17, and 42.97 × 10−17 for the 20th, 40th, and 80th cycle, respectively), whereas in the case of the surface-modified sample it is almost constant (12.95 × 10−16, 12.47 × 10−16, and 13.60 × 10−16 for the 20th, 40th, and 80th cycle, respectively). Incidentally, the decreasing DLi+ in LNMC coincides with the capacity degradation (Figure 6b) as well as spinel phase transformation (Figure 6e), which is observed as a function of the cycle number. In contrast, the almost constant value of DLi+ for the surface-modified sample is in accordance with the capacity retention and mitigation of phase transformation. To demonstrate the structural integrity of the surfacemodified cathode, HRTEM investigation of the electrode material has been conducted after the completion of 100 cycles (Figure 9). The microscopic results confirm that the LZO

(1)

where R is the gas constant, T is the absolute temperature, A is the area of the electrode, n is the number of electrons per molecule during charge−discharge process, F is the Faraday constant, C is the concentration of Li+ per unit cell and σ is the Warburg factor, which is obtained by linear fitting of the inclined lines at low frequencies, depicted in Figure 7e,f, and it has a relationship with Z (ω = 2πf), as shown in eq 2. Z = R sl + R ct + σω−1/2

Figure 9. High-resolution transmission electron microscopy images of LZO-coated LNMC electrode after completing 100 charge−discharge cycles.

(2)

coating is neither pulverized nor detached from the surface of the core active material (LNMC). The intactness after the prolonged cycle is attributed to the structural integrity that persists between LZO and LNMC, which is responsible for the improved performance. In summary, the strategy of LZO coating on the surface of LNMC can play a significant positive role and pave the way for commercial realization of Li-rich cathode materials as high-capacity electrodes.

The calculated diffusion coefficient values are plotted in Figure 8. Before cycling, the pristine sample shows a diffusion coefficient of 23.77 × 10−16, whereas that for the surfacemodified electrode is 22.18 × 10−16. After the first cycle, DLi+ values of both the pristine and surface-coated material are reduced to 12.90 × 10−16 and 15.45 × 10−16, respectively. This decrease in the diffusion coefficient is attributed to the structural rearrangement after charge. In contrast, the decreasing diffusivity coefficient in the surface-coated sample is comparatively low due to suppression of oxygen evolution during the first charge. Upon further cycling, the pristine sample shows that the rate of Li+ diffusion drastically decreases with the increasing cycle number (10.02 × 10−16, 58.83 ×



CONCLUSIONS A facile approach to prepare a uniform coating of a Li-ion conducting LZO surface layer over a high-capacity LNMC cathode is reported in the present study. The growth of LZO 2313

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

negative as well as reference electrode. The Whatman glass microfiber filter paper acted as a separator. The 1 M LiPF6 dissolved in a mixture of diethyl carbonate, dimethyl carbonate, and ethylene carbonate (2:1:2 ratio by volume) was used as the electrolyte. In a typical experiment, the cathode loading was 4− 6 mg per cell. The galvanostatic cycling experiments were carried out at room temperature at C/5 rate, where C represents the theoretical exchange of 1 mol of lithium per formula unit in 1 h. The cells were cycled between the potential window of 1.75 and 4.8 V using a VMP3Z biologic multichannel potentiostat/galvanostat.

with the preferred orientation and alignment on LNMC provides a new strategy to design a robust surface coating, which can possess an ion pathway and thereby enhance the initial Coulombic efficiency and rate performances. This is mainly attributed to (a) the protective role of Li-ion conductive surface barrier between the bulk electrode and electrolyte and (b) improved Li+ migration facilitating lithium diffusion pathways on the surface of the cathode. The intactness of the surface coating significantly suppresses the layered to spinel structural transformation and consequently inhibits the voltage decay. Most importantly, the dissolution of Mn from the bulk cathode into the electrolyte is also drastically mitigated. The surface modification strategy proposed here broadens the scope of further investigation on lithium-rich phases toward commercialization aspects and triggers new insights into the inhabitation of O2 evolution at high voltage and safety concerns of high capacity electrodes.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00381. Structural parameters obtained from two phase Rietveld refinement of XRD data (Tables S1 and S2); values of Rs, Rsl, and Rct obtained by simulated data of an impedance spectrum (Table S3); core level XPS spectra of Li1s for pristine and LZO @ LNMC samples (Figure S1); FTIR patterns of the pristine and LZO @ LNMC samples (Figure S2) (PDF)



EXPERIMENTAL SECTION The LZO-coated LNMC sample was prepared by a two step procedure as follows. In the first step, nanocrystalline LNMC was synthesized using a solution combustion process, as reported in our previous publication.34 The surface modification of solution combustion-synthesized LNMC with LZO was carried out as follows. Stoichiometric amounts of LiOH·H2O dissolved in methanol and Zr[O(CH2)3CH3]4 dissolved in carbon tetrachloride were mixed together, in which the nanocrystalline LNMC was added. The solution was kept for constant stirring at room temperature for 6 h. Then, a few drops of deionized water were added into the suspension to undergo hydrolysis reaction. The stirring was maintained for another 6 h, followed by gently heating at 60 °C to evaporate the complete solvent. The powder was then heated at 600 °C for 5 h. Hereafter, we refer to the sample, Li1.2Ni0.13Mn0.54Co0.13O2 as Li-rich NMC (LNMC) and Li2ZrO3-coated Li1.2Ni0.13Mn0.54Co0.13O2 as LZO @ LNMC. The phase characterization of these materials was primarily carried out by X-ray diffraction with a Bruker D8 Advance diffractometer, using a Cu Kα radiation (α1 = 1.54056 Å, α2 = 1.54439 Å). The morphology and size of the powder samples were obtained using field emission scanning electron microscope (FESEM MIRA3 LMU). HRTEM studies were carried out with a Tecnai-G2 F20 microscope. XPS of these samples were recorded with a Thermo Fisher Scientific Multilab 2000 spectrometer fitted with nonmonochromatic Al Kα radiation (1486.6 eV) as the X-ray source operated at a power of 150 W (12.5 mA, 12 kV). The reported binding energies were corrected with reference to the C1s peak at 284.6 eV. For XPS analysis, powder samples were pelletized to 8 mm diameter, which were mounted on the sample holders and placed in the load-lock chamber in ultrahigh vacuum (UHV) at 8 × 10−8 mbar for 5 h to desorb any volatile species absorbed on the surface. After 5 h, the samples were introduced one by one inside the analyzer chamber, with UHV at 5 × 10−10 mbar. All of the spectra were recorded with pass energy of 40 eV and step increment of 0.05 eV. Infrared spectra were obtained using the FTIR spectrometer, Bruker Optick GmbH, using KBr pressed disk. Electrochemical experiments were conducted on swageloktype Li half-cells constructed in an argon-filled glovebox. The positive electrode was formulated with the ball-milled mixture at an 80:20 weight ratio of the active material and Super-P Li Carbon (Timcal Belgium). The lithium metal disc was used as a



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected], [email protected]. ORCID

Marappan Sathish: 0000-0001-9094-5822 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank the Council of Scientific and Industrial Research (CSIR), India, for support of this work under the TAPSUN project. K.R. Prakasha thanks CSIR for a CSIR-UGC Fellowship. The authors thank the Central Instrumentation Facility, CSIR-CECRI, Karaikudi, and the CSIR Innovation complex facilities at CSIR Madras Complex, Chennai, for providing characterization facilities.



REFERENCES

(1) Armstrong, A. R.; Holzapfel, M.; Novák, P.; Johnson, C. S.; Kang, S. H.; Thackeray, M. M.; Bruce, P. G. Demonstrating Oxygen Loss and Associated Structural Reorganization in the Lithium Battery Cathode Li[Ni0.2Li0.2Mn0.6]O2. J. Am. Chem. Soc. 2006, 128, 8694−8698. (2) Johnson, C. S.; Li, N.; Lefief, C.; Vaughey, J. T.; Thackeray, M. M. Synthesis, Characterization and Electrochemistry of Lithium Battery Electrodes: xLi2MnO3·(1 − x)LiMn0.333Ni0.333Co0.333O2 (0 ≤ x ≤ 0.7). Chem. Mater. 2008, 20, 6095−6106. (3) Kim, M. G.; Jo, M.; Hong, Y. S.; Cho, J. Template-free synthesis of Li[Ni0.25Li0.15Mn0.6]O2 nanowires for high performance lithium battery cathode. Chem. Commun. 2009, 2, 218−220. (4) Ellis, B. L.; Lee, K. T.; Nazar, L. F. Positive Electrode Materials for Li-Ion and Li-Batteries. Chem. Mater. 2010, 22, 691−714. (5) Wei, Y.; Zheng, J. X.; Cui, S. H.; Song, X. H.; Su, Y. T.; Deng, W. J.; Wu, Z. Z.; Wang, X. W.; Wang, W. D.; Rao, M. M.; Lin, Y.; Wang, C. M.; Amine, K.; Pan, F. Kinetics Tuning of Li-ion Diffusion in Layered Li(NixMnyCoz)O2. J. Am. Chem. Soc. 2015, 137, 8364−8367. (6) Liu, J.; Liu, J.; Wang, R.; Xia, Y. Degradation and Structural Evolution of xLi2MnO3·(1−x)LiMn1/3Ni1/3Co1/3O2 during Cycling. Y. J. Electrochem. Soc. 2014, 161, A160−A167. 2314

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

aluminum isopropoxide on lithium-rich layered oxide. J. Power Sources 2015, 281, 444−454. (24) Kim, I. T.; Knight, J. C.; Celio, H.; Manthiram, A. Enhanced electrochemical performances of Li-rich layered oxides by surface modification with reduced graphene oxide/AlPO4 hybrid coating. J. Mater. Chem. A 2014, 2, 8696−8704. (25) Qiao, Q. Q.; Zhang, H. Z.; Li, G. R.; Ye, S. H.; Wangb, C. W.; Gao, X. P. Surface modification of Li-rich layered Li(Li0.17Ni0.25Mn0.58)O2 oxide with LiMnPO4 as the cathode for lithium-ion batteries. J. Mater. Chem. A 2013, 1, 5262−5268. (26) Xu, G.; Li, J.; Xue, Q.; Dai, Y.; Zhou, H.; Wang, X.; Kang, F. Elevated electrochemical performance of (NH4 ) 3 AlF 6 -coated 0.5Li2MnO3·0.5LiNi1/3Co1/3Mn1/3O2 cathode material via a novel wet coating method. Electrochim. Acta 2014, 117, 41−47. (27) Fu, Q.; Du, F.; Bian, X.; Wang, Y.; Yan, X.; Zhang, Y.; Zhu, K.; Chen, G.; Wang, C.; Wei, Y. Electrochemical performance and thermal stability of Li1.18Co0.15Ni0.15Mn0.52O2 surface coated with the ionic conductor Li3VO4. J. Mater. Chem. A 2014, 2, 7555−7562. (28) Li, X.; Liu, J.; Banis, M. N.; Lushington, A.; Li, R.; Cai, M.; Sun, X. Atomic layer deposition of solid-state electrolyte coated cathode materials with superior high-voltage cycling behaviour for lithium ion battery application. Energy Environ. Sci. 2014, 7, 768−778. (29) Zhanga, X.; Suna, S.; Wua, Q.; Wana, N.; Pana, D.; Bai, Y. Improved electrochemical and thermal performances of layered Li[Li0.2Ni0.17Co0.07Mn0.56]O2 via Li2ZrO3 surface modification. J. Power Sources 2015, 282, 378−384. (30) Zhang, J.; Li, Z.; Gao, R.; Hu, Z.; Liu, X. High Rate Capability and Excellent Thermal Stability of Li+-Conductive Li2ZrO3-Coated LiNi1/3Co1/3Mn1/3O2 via a Synchronous Lithiation Strategy. J. Phys. Chem. C 2015, 119, 20350−20356. (31) Zhang, J.; Zhang, H.; Gao, R.; Li, Z.; Hu, Z.; Liu, X. New insights into the modification mechanism of Li-rich Li1.2Mn0.6Ni0.2O2coated by Li2ZrO3. Phys. Chem. Chem. Phys. 2016, 18, 13322−13331. (32) Wu, H.; Wang, Z.; Liu, S.; Zhang, L.; Zhang, Y. Fabrication of Li+-Conductive Li2ZrO3-Based Shell Encapsulated LiNi0.5Co0.2Mn0.3O2 Microspheres as High-Rate and Long-Life Cathode Materials for LiIon Batteries. ChemElectroChem 2015, 2, 1921−1928. (33) Song, B.; Li, W.; Oh, S.-M; Manthiram, A. Long-life nickel-rich layered oxide cathodes with a uniform Li2ZrO3 surface coating for lithium-ion batteries. ACS Appl. Mater. Interfaces 2017, 9, 9718−9725. (34) Prakasha, K. R.; Prakash, A. S. A time and energy conserving solution combustion synthesis of nano Li1.2Ni0.13Mn0.54Co0.13O2 cathode material and its performance in Li-ion batteries. RSC Adv. 2015, 5, 94411−94417. (35) Weill, F.; Tran, N.; Croguennec, L.; Delmas, C. Cation ordering in the layered Li1+x(Ni0.425Mn0.425Co0.15)1 − xO2 materials (x = 0 and 0.12). J. Power Sources 2007, 172, 893−900. (36) Boulineau, A.; Croguennec, L.; Delmas, C.; Weill, F. Reinvestigation of Li2MnO3 Structure: Electron Diffraction and High Resolution TEM. Chem. Mater. 2009, 21, 4216−4222. (37) Arinkumar, T. A.; Wu, Y.; Manthiram, A. Factors Influencing the Irreversible Oxygen Loss and Reversible Capacity in Layered Li[Li1/3Mn2/3]O2−Li[M]O2 (M = Mn0.5‑yNi0.5‑yCo2y and Ni1‑yCoy) Solid Solutions. Chem. Mater. 2007, 19, 3067−3073. (38) Zhou, W.; Cheng, C.; Liu, J.; Tay, Y. Y.; Jiang, J.; Jia, X.; Zhang, J.; Gong, H.; Hng, H. H.; Yu, T.; Fan, H. J. Epitaxial Growth of Branched α -Fe2O3 /SnO2 Nano-Heterostructures with Improved Lithium-Ion Battery Performance. Adv. Funct. Mater. 2011, 21, 2439− 2445. (39) Niu, M. T.; Huang, F.; Cui, L. F.; Huang, P.; Yu, Y. L.; Wang, Y. S. Hydrothermal Synthesis, Structural Characteristics, and Enhanced Photocatalysis of SnO2/α-Fe2O3 Semiconductor Nano heterostructures. ACS Nano 2010, 4, 681−688. (40) Moulder, F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D. Handbook of X-ray Photoelectron Spectroscopy; Perkin-Elmer Corporation: Eden Prairie, 1992.

(7) Li, X.; Xu, M.; Chen, Y.; Lucht, B. L. Surface study of electrodes after long-term cycling in Li1.2Ni0.15Mn0.55Co0.1O2 graphite lithium-ion cells. J. Power Sources 2014, 248, 1077−1084. (8) Gu, M.; Belharouak, I.; Zheng, J.; Wu, H.; Xiao, J.; Genc, A.; Amine, K.; Thevuthasan, S.; Baer, D. R.; Zhang, J. G.; et al. Formation of the Spinel Phase in the Layered Composite Cathode Used in Li-Ion Batteries. ACS Nano 2013, 7, 760−767. (9) Ates, M. N.; Jia, Q.; Shah, A.; Busnaina, A.; Mukerjee, S.; Abraham, K. M. Mitigation of Layered to Spinel Conversion of a LiRich Layered Metal Oxide Cathode Material for Li-Ion Batteries. J. Electrochem. Soc. 2014, 161, A290−A301. (10) Zhu, Y.; Li, Y.; Abraham, D. P. Mitigating Performance Degradation of High-Capacity Lithium-Ion Cells with Boronate-Based Electrolyte Additives. J. Electrochem. Soc. 2014, 161, A1580−A1585. (11) Trask, S. E.; Li, Y.; Kubal, J. J.; Bettge, M.; Polzin, B. J.; Zhu, Y.; Jansen, A. N.; Abraham, D. P. From coin cells to 400 mAh pouch cells: Enhancing performance of high-capacity lithium-ion cells via modifications in electrode constitution and fabrication. J. Power Sources 2014, 259, 233−244. (12) Bettge, M.; Li, Y.; Sankaran, B.; Rago, N. D.; Spila, T.; Haasch, R. T.; Petrov, I.; Abraham, D. P. Improving high-capacity Li1.2Ni0.15Mn0.55Co0.1O2-based lithium-ion cells by modifying the positive electrode with alumina. J. Power Sources 2013, 233, 346−357. (13) Zhang, H. Z.; Qiao, Q. Q.; Li, G. R.; Gao, X. P. PO43− polyanion-doping for stabilizing Li-rich layered oxides as cathode materials for advanced lithium-ion batteries. J. Mater. Chem. A 2014, 2, 7454−7460. (14) Iftekhar, M.; Drewett, N. E.; Armstrong, A. R.; Hesp, D.; Braga, F.; Ahmed, S.; Hardwick, L. J. Characterization of Aluminum Doped Lithium-Manganese Rich Composites for Higher Rate Lithium-Ion Cathodes. J. Electrochem. Soc. 2014, 161, A2109−A2116. (15) Min, J. W.; Yim, C. J.; Im, W. B. Facile Synthesis of Electrospun Li1.2Ni0.17Co0.17Mn0.54O2 Nanofiber and Its Enhanced High-Rate Performance for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2013, 5, 7765−7769. (16) Shen, C. H.; Wang, Q.; Fu, F.; Huang, L.; Lin, Z.; Shen, S. Y.; Su, H.; Zheng, X. M.; Xu, B. B.; Li, J. T.; Sun, S. G. Facile Synthesis of The Li-Rich Layered Oxide Li1.23Ni0.09Co0.12Mn0.56O2 with Superior Lithium Storage Performance and New Insights into Structural Transformation of the Layered Oxide Material during Charge− Discharge Cycle: In Situ XRD Characterization. ACS Appl. Mater. Interfaces 2014, 6, 5516−5524. (17) Sathiya, M.; Abakumov, A. M.; Foix, D.; Rousse, G.; Ramesha, K.; Saubanere, M.; Doublet, M. L.; Vezin, H.; Laisa, C. P.; Prakash, A. S.; Gonbeau, D.; VanTendeloo, G.; Tarascon, J. M. Origin of voltage decay in high-capacity layered oxide electrodes. Nat. Mater. 2015, 14, 230−238. (18) Mohanty, D.; Li, J.; Abraham, D. P.; Huq, A.; Payzant, E. A.; Wood, D. L., III; Daniel, C. Unraveling the Voltage-Fade Mechanism in High-Energy-Density Lithium-Ion Batteries: Origin of the Tetrahedral Cations for Spinel Conversion. Chem. Mater. 2014, 26, 6272−6280. (19) Jung, Y. S.; Cavanagh, A. S.; Yan, Y.; George, S. M.; Manthiram, A. Effects of Atomic Layer Deposition of Al2O3 on the Li[Li0.20Mn0.54Ni0.13Co0.13]O2 Cathode for Lithium-Ion Batteries. J. Electrochem. Soc. 2011, 158, A1298−A1302. (20) Uzun, D.; Doğrusöz, M.; Mazman, M.; Biçer, E.; Avci, E.; Şener, T.; Kaypmaz, T. C.; Demir-Cakan, R. Effect of MnO2 coating on layered Li(Li0.1Ni0.3Mn0.5Fe0.1)O2 cathode material for Li-ion batteries. Solid State Ionics 2013, 249−250, 171−176. (21) Yuan, W.; Zhang, H. Z.; Liu, Q.; Li, G. R.; Gao, X. P. Surface modification of Li(Li0.17Ni0.2Co0.05Mn0.58)O2 with CeO2 as cathode material for Li-ion batteries. Electrochim. Acta 2014, 135, 199−207. (22) Xu, M.; Chen, Z.; Zhu, H.; Yan, X.; Li, L.; Zhao, Q. Mitigating Capacity Fade by Constructing Highly-Ordered Mesoporous Al2O3/ Polyacene Double-Shelled Architecture in Li-Rich Cathode Materials. J. Mater. Chem. A 2015, 3, 13933−13945. (23) Xu, M.; Chen, Z.; Li, L.; Zhu, H.; Zhao, Q.; Xu, L.; Peng, N.; Gong, L. Highly crystalline alumina surface coating from hydrolysis of 2315

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316

ACS Omega

Article

(41) Hemalatha, K.; Jayakumar, M.; Bera, P.; Prakash, A. S. Improved electrochemical performance of Na0.67MnO2 through Ni and Mg substitution. J. Mater. Chem. A 2015, 3, 20908−20912. (42) Ou, Y. N.; Li, G. R.; Liang, J. H.; Feng, Z. P.; Tong, Y. X. Ce1−xCoxO2−δ Nanorods Grown by Electrochemical Deposition and Their Magnetic Properties. J. Phys. Chem. C 2010, 114, 13509−13514. (43) Petitto, S. C.; Langell, M. A. Surface composition and structure of Co3O4 (110) and the effect of impurity segregation. J. Vac. Sci. Technol., A 2004, 22, 1690−1696. (44) Baidya, T.; Bera, P. Investigation of support effect on CO adsorption and CO + O2 reaction over Ce1‑x‑yMxCuyO2‑δ (M = Zr, Hf and Th) catalysts by in situ DRIFTS. Catal., Struct. React. 2015, 1, 110−119. (45) Bera, P.; Seenivasan, H.; Rajam, K. S.; Shivakumara, C.; Parida, S. K. Characterization and micro hardness of Co−W coatings electrodeposited at different pH using gluconate bath: A comparative study. Surf. Interface Anal. 2013, 45, 1026−1036. (46) Zhang, J.; Wang, J.; Yang, J.; NuLi, Y. Artificial Interface Deriving from Sacrificial Tris(trimethylsilyl)phosphate Additive for Lithium Rich Cathode Materials. Electrochim. Acta 2014, 117, 99−104. (47) Martha, S. K.; Nanda, J.; Veith, G. M.; Dudney, N. J. Surface Studies of High Voltage Lithium Rich Composition: Li1.2Mn0.525Ni0.175Co0.1O2. J. Power Sources 2012, 216, 179−186. (48) Hy, S.; Felix, F.; Rick, J.; Su, W. N.; Hwang, B. J. Direct In situ Observation of Li2O Evolution on Li-Rich High-Capacity Cathode Material, Li[NixLi(1−2x)/3Mn(2‑x)/3]O2 (0 ≤ x ≤ 0.5). J. Am. Chem. Soc. 2014, 136, 999−1007. (49) Lu, Z. H.; Dahn, J. R. Understanding the Anomalous Capacity of Li/Li[NixLi(1/3−2x/3)Mn(2/3‑x/3]O2 Cells Using In Situ X-ray Diffraction and Electrochemical Studies. J. Electrochem. Soc. 2002, 149, A815−A822. (50) Li, Y.; Wu, C.; Bai, Y.; Liu, L.; Wang, H.; Wu, F.; Zhang, N.; Zou, Y. Hierarchical Mesoporous Lithium-Rich Li[Li0.2Ni0.2Mn0.6]O2 Cathode Material Synthesized via Ice Templating for Lithium Ion Battery. ACS Appl. Mater. Interfaces 2016, 8, 18832−18840.

2316

DOI: 10.1021/acsomega.7b00381 ACS Omega 2017, 2, 2308−2316