3D Reticular Li1.2Ni0.2Mn0.6O2 Cathode Material for Lithium-Ion

Dec 27, 2016 - The synthesized 3D reticular Li1.2Ni0.2Mn0.6O2 microparticles consisted of two interlaced. 3D nanonetworks and a mesopore channel syste...
0 downloads 0 Views 4MB Size
Research Article www.acsami.org

3D Reticular Li1.2Ni0.2Mn0.6O2 Cathode Material for Lithium-Ion Batteries Li Li,§,†,‡ Lecai Wang,§,† Xiaoxiao Zhang,† Qing Xue,† Lei Wei,† Feng Wu,†,‡ and Renjie Chen*,§,†,‡ †

ACS Appl. Mater. Interfaces 2017.9:1516-1523. Downloaded from pubs.acs.org by UNIV OF LOUISIANA AT LAFAYETTE on 10/09/18. For personal use only.

School of Materials Science & Engineering, Beijing Key Laboratory of Environmental Science and Engineering, Beijing Institute of Technology, Beijing 100081, China ‡ Collaborative Innovation Center of Electric Vehicles in Beijing, Beijing 100081, China ABSTRACT: In this study, a hard-templating route was developed to synthesize a 3D reticular Li1.2Ni0.2Mn0.6O2 cathode material using ordered mesoporous silica as the hard template. The synthesized 3D reticular Li1.2Ni0.2Mn0.6O2 microparticles consisted of two interlaced 3D nanonetworks and a mesopore channel system. When used as the cathode material in a lithium-ion battery, the as-synthesized 3D reticular Li1.2Ni0.2Mn0.6O2 exhibited remarkably enhanced electrochemical performance, namely, superior rate capability and better cycling stability than those of its bulk counterpart. Specifically, a high discharge capacity of 195.6 mA h g−1 at 1 C with 95.6% capacity retention after 50 cycles was achieved with the 3D reticular Li1.2Ni0.2Mn0.6O2. A high discharge capacity of 135.7 mA h g−1 even at a high current of 1000 mA g−1 was also obtained. This excellent electrochemical performance of the 3D reticular Li1.2Ni0.2Mn0.6O2 is attributed to its designed structure, which provided nanoscale lithium pathways, large specific surface area, good thermal and mechanical stability, and easy access to the material center. KEYWORDS: 3D reticular, Li1.2Ni0.2Mn0.6O2, cathode, lithium-ion battery, hierarchy

1. INTRODUCTION Cathode material is one of the major components of modern lithium-ion batteries and directly influences the performance and price of the final battery pack. The high cost of current state-of-the-art commercialized lithium-ion batteries is still the main barrier to the storage of energy on the electrical grid to improve grid efficiency and stability while enabling integration of intermittent renewable energy technologies (such as wind and solar) into the baseload supply.1 Furthermore, user experiences of electric vehicles and portable electronics are rarely free from battery life anxiety. Recently, the lithium-rich cathode material Li1.2Ni0.2Mn0.6O2 has attracted much attention for its high specific capacity (ca. 250 mA h g−1), a promising factor for long battery life.2−13 It is believed that this layered oxide belongs to the xLi2MnO3·(1− x)LiMO2 (M = Mn, Co, or Ni) family, which is composed of two layered α-NaFeO2-type structures.12 One of the most interesting features of these lithium-rich materials is that, except for the active LiMO2 (M = Mn, Co, or Ni) phase, they can be charged to a high potential (>4.5 V vs Li), resulting in the electrochemical activation of the Li2MnO3 component and a high capacity. Notably, Li1.2Ni0.2Mn0.6O2 is also relatively cheap and has a low environmental toxicity owing to the absence of cobalt. Along with its intrinsic high capacity, however, Li1.2Ni0.2Mn0.6O2 suffers from poor rate capacity and modest cycle stability, mainly because of structural rearrangement at its surface during cycling.12 Specifically, lithium extraction of © 2016 American Chemical Society

lithium-rich materials is accompanied by the oxidation of the O2p band (Li2O) during the initial charge, followed by the irreversible migration of a significant amount of transition-metal ions into the lithium sites.6 Many strategies, such as element doping and surface coating, have been adopted to improve the electrochemical performance of lithium-rich layered oxides.3,4,14,15 These methods can mitigate the poor electrochemical performance fairly. The recently reported gradient surface Na+ doping method showed an insight by enhancing the kinetics.16 Meanwhile, reducing the particle size to nanoscale levels has also been proven to be an effective way to improve the rate capability of layered oxides.13,17,18 Use of nanosized cathode materials shortens lithium-ion diffusion pathways and increases the size of the active surfaces, mainly the (010) lattice planes in the case of αNaFeO2 structure layered oxides such as Li1.2Ni0.2Mn0.6O2; therefore, electrode kinetic issues can be circumvented by switching to nanomaterials. This is an approach with high potential: scientists have been successfully controlling and expanding the properties of materials through nanotechnology in many fields for decades.19−24 Research also indicates that nanomaterials are usually thermodynamically unstable and tend to agglomerate, which increases the resistance of the final Received: October 18, 2016 Accepted: December 27, 2016 Published: December 27, 2016 1516

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

Scheme 1. Brief Illustration of the Designed Route for Synthesizing 3D Reticular Li1.2Ni0.2Mn0.6O2 and the Morphological Effects of the As-Synthesized Material

several times with water and ethanol, the sample was dried at 60 °C and labeled as r-NMO. To prepare 3D reticular Li1.2Ni0.2Mn0.6O2, 0.2 g of r-NMO was mixed with excess LiOH·H2O (7%) in 20 mL of ethanol. The mixture was dried at room temperature with stirring and then heated to 800 °C at 2 °C min−1 and calcined for 5 h. The final product was labeled as rLNMO. Bulk Li1.2Ni0.2Mn0.6O2 (b-LNMO) was prepared according to the literature with a change only in the final calcination conditions to 800 °C for 5 h.29 Ultrasound-treated r-LNMO (ur-LNMO) was also prepared for a thermal stability test. 2.2. Characterization. Crystallographic analysis of the prepared samples was carried out on an X-ray diffractometer (XRD; Rigaku Ultima IV-185) with a Cu Kα radiation source. Field emission scanning electron microscopy (FESEM; FEI, Quanta 200f), specific surface area and pore size analysis (ASIQM000-1-MP, Thermo Fisher Scientific Inc.), transmission electron microscopy (TEM; JEM-2100f), and small-angle X-ray scattering (SAXS; Rigaku SmartLab 2) were used to characterize the morphology of the prepared samples. Elemental mappings were carried out with an energy dispersive Xray detector (EDX). X-ray photoelectron spectroscopy (XPS; VGESCA-LABMK II) was used to determine the ion valence states in the metal oxide. Differential scanning calorimetry (DSC; MettlerToledo DSC 1) was used to evaluate the thermodynamic stability of the samples. 2.3. Electrochemical Tests. Electrochemical tests were performed on two-electrode coin cells (type: 2025) at ambient temperature. In this case, lithium metal served as both the reference electrode and the counter electrode. The working electrode was composed of 80 wt % active material, 10 wt % conductive additive (acetylene black), and 10 wt % binder (polyvinylidene difluoride, PVDF) with an aluminum foil current collector. A lithium salt solution, 1 M LiPF6 in a mixture of equal volumes of dimethyl carbonate (DMC), ethylene carbonate (EC), and diethyl carbonate (DEC) was used as the electrolyte. Cell assembly was carried out in a glovebox filled with argon, the oxygen and moisture contents of which were maintained below 1 ppm. Galvanostatic charge/discharge tests were implemented on a battery tester (LAND-CT2001A) with a voltage window of 2.0−4.8 V at selected current rates. Cells were cycled under 0.1 C for three cycles before continuous high rate (e.g., 1 and 5 C) tests. Cyclic voltammetry (CV) tests were carried out on a CHI660d electrochemical workstation with the same voltage window. Electrochemical impedance spectroscopy (EIS) tests were also carried out on the CHI660d in the frequency range of 0.1 MHz to 0.01 Hz with an AC perturbation signal of 5 mV.

electrode. Also, a high specific surface area tends to promote side reactions between the electrode and electrolyte, leading to inferior cycle performance and poor safety. To make the nano strategy work, the better option is to form a functional structure by artificially arranging and stabilizing nanosized constructions in the structural hierarchy of the material in a controlled and expected way. This is more than only seemingly reasonable. Indeed, Ozin used a whole book to introduce the building of good structure−function relationships using chemistry a number of years ago.25 Meanwhile, a category of successful designs for cathode materials in which primary and secondary particles are clearly defined in a nano/microstructure hierarchy has acted as a demonstration, including those in our 2015 study.11,13,17,18,26 Herein, we adopt a hard-templating method to synthesize 3D reticular Li1.2Ni0.2Mn0.6O2. In the hierarchy of 3D reticular Li1.2Ni0.2Mn0.6O2, microscopic particles of the material are constructed from gyroid-like nano-3D networks inherited from a 3D reticular Ni0.75Mn2.25O4 precursor synthesized using mesoporous silica (KIT-6) as a template. In addition to controlling the morphology of synthesized samples, the template also filled a certain amount of space in the particles. When the template was etched away, a mesopore channel system was created in each individual particle. As a cathode material, 3D reticular Li1.2Ni0.2Mn0.6O2 showed a significant improvement in electrochemical performance compared with that of the bulk counterpart.

2. EXPERIMENTAL SECTION 2.1. Preparation of Materials. Mn(NO3)2·4H2O (99%), Ni(NO3)2·6H2O (99%), NaOH (99.3%), LiOH·H2O (99%), 1-butanol (99.4%), ethanol (99.0%), aqueous ammonium hydroxide (28−30%), and tetraethyl orthosilicate (98%) were purchased from Aladdin (Shanghai, China). Pluronic P123 (Mn = 5800) was purchased from Sigma-Aldrich (St. Gallen, United States). To synthesize the 3D reticular Ni0.75Mn2.25O4 precursor, 1.5 g of the two metal nitrates in a molar ratio of 1:3 was dissolved in 20 mL of ethanol followed by the addition of 1 g of mesoporous silica (KIT-6) synthesized according to the literature.27,28 After the mixture was stirred at room temperature until almost dry, the sample was heated slowly to 700 °C and then held at that temperature for 5 h. The resulting material was treated three times with a hot solution of 2.0 M NaOH to remove the silica template. After centrifugation and washing 1517

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

3. RESULT AND DISCUSSION 3.1. Composition and Morphology Analysis of rLNMO. A 3D r-LNMO particle consists of two interlaced 3D networks and a mesopore channel system. Ideally, the interface of the particle should be a double-gyroid surface with the Ia3̅d space group. In practice, the obtained structure is slightly rough but still provides excellent performance for electrochemical applications. Scheme 1 briefly shows an r-LNMO electrode and the hard-templating process used to synthesize r-LNMO. XRD patterns were collected to investigate the crystallographic characteristics of the as-prepared r-LNMO, as shown in Figure 1. Most peaks in the XRD pattern of r-LNMO could be

sponding to impurity phases were observed. In contrast, the assynthesized Ni0.75Mn2.25O4 precursor inherited the cubic spinel structure (space group Fd3̅ m) with interconnecting O 6 octahedra as alternate layers along the [111] direction. According to the literature, the similarity between the radius of Li+ and Ni2+ plays a crucial role in making cubic-to-layered topotaxy on the basis of [0001]h/[111]c possible.31 During the transformation, the O6 octahedra stacking layers function as the fundamental framework. The structural diagrams presented in Figures 1b and c provide a more vivid demonstration of this topotaxy. XPS measurements were performed to learn more about the valence state of the transition metals in r-LNMO. The results were calibrated with reference to the C 1s binding energy at 284.5 eV and are presented in Figure 2. The dominant peaks at 854.5 and 642.0 eV are attributed to Ni2+ and Mn4+, respectively.32 As shown in Figure 2b, the intense Ni 2p3/2 peak at about 854.5 eV was accompanied by a shakeup peak at about 861.0 eV. This particular pattern is characteristic of Ni2+. Meanwhile, a less prominent peak at about 855.8 eV was also observed, indicating the existence of Ni3+. In the Mn 2p spectrum (Figure 2c), two main spin−orbit lines at about 642.0 eV for 2p3/2 and 653.7 eV for 2p1/2 with a separation of 11.7 eV indicated a majority of Mn4+. However, a pair of less prominent peaks at about 643.4 and 655.1 eV revealed the presence of Mn3+. According to the literature, minor contributions of Ni3+ and Mn3+ are inevitable owing to valence degeneracy.33 The O 1s spectrum (Figure 2d) of r-LNMO contained two peaks. The peak at 529.2 eV is typical of metal oxygen bonds, while that at 531.5 eV is associated with low coordinative oxygen ions at the surface of the material.34 Brunauer−Emmett−Teller (BET) analysis (Figure 3a) showed that r-LNMO possessed a specific surface area as large as 61.695 m2 g−1. A large specific area directly enlarges the solid−liquid interface in the final electrodes and simultaneously promotes rate performance. The average pore width of rLNMO was 5.872 nm, which may promise a performance of rLNMO similar to that of other mesoporous cathode materials.13,28,35 Higher capacity is one of the effects of mesopores.13 The 3D reticular morphology of r-LNMO was first confirmed by the obtained SAXS data. As shown in Figure 3b, the relatively sharp diffraction peaks in the SAXS pattern could be indexed to the corresponding reflections in the cubic Ia3̅d space group, which indicated the presence of a doublegyroid morphology as intended.36 The SEM image (Figure 3c) confirmed that the r-LNMO particles were microscale. Indications of networks and the mesopore channel system were found on the surface of the material in the high magnification SEM image (Figure 3c). The 3D reticular structure could be easily identified in TEM images (Figure 3e). A honeycomb-like structure along the [111] direction was clearly observed. Ideally, electrolyte cloud fluxes through the whole particle along this direction. The r-LNMO was further analyzed by EDX analysis (Figure 3f). According to the results, the ratio of the transition metals in the synthesized sample was very close to stoichiometric, and both the elements were uniformly distributed in the particle. 3.2. Electrochemical Performance of r-LNMO. Galvanostatic charge/discharge curves measured for r-LNMO at 0.1 C rate (defined as 20 mA g−1) between 2.0 and 4.8 V are displayed in Figure 4a. The r-LNMO material delivered initial charge/discharge capacities of 364.55 and 259.14 mA h g−1, respectively. As mentioned before, the typical high charge

Figure 1. XRD pattern of the 3D reticular Li1.2Ni0.2Mn0.6O2 sample (a), crystallographic structure of spinel Ni0.75Mn2.25O4 (b), and layered Li1.2Ni0.2Mn0.6O2 (c).

indexed to the layered rock-salt form with hexagonal αNaFeO2-type structure (R3̅m space group). The low-intensity peaks within the range of 20−25° (dashed circle) corresponded to Li2MnO3 with C/2m space group, which is characteristic of Li-rich Mn-based materials.30 The clear (006/102) and (018/ 110) split peaks were evidence of a high degree of crystallization. Meanwhile, no distinct reflection peaks corre1518

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

Figure 2. XPS spectra of (a) C 1s, (b) O 1s, (c) Ni 2p, and (d) Mn 2p for 3D r-LNMO.

Figure 3. Nitrogen adsorption/desorption isotherms and pore size distribution of 3D reticular Li1.2Ni0.2Mn0.6O2 (r-LNMO) (a), SAXS curve of rLNMO (b), SEM images of r-LNMO (c and d), TEM image of r-LNMO (e), and EDX results of r-LNMO (f).

capacity of r-LNMO in the first cycle is caused by the electrochemical activation of the Li2MnO3 component when

charged over 4.5 V. Much charge/discharge capacities higher than those of conventional cathode materials (e.g., LiCoO2 and 1519

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

Figure 4. Charge/discharge curves of a 3D reticular Li1.2Ni0.2Mn0.6O2 (r-LNMO) electrode at 0.1 C in the voltage range between 2.0−4.8 V (a) and CV curves of an r-LNMO electrode in the voltage range of 2.0−4.8 V at the scan rate of 0.1 mV s−1 (b).

Figure 5. Cycle performances for 3D reticular and bulk Li1.2Ni0.2Mn0.6O2 (r-LNMO and b-LNMO, respectively) electrodes at 1 and 0.2 C rates (a and b), rate performance of r-LNMO and b-LNMO electrodes (c), high temperature performance of r-LNMO (d), Nyquist plots of r-LNMO and bLNMO electrodes in the frequency range from 100 kHz to 0.1 Hz (e), and performance comparison table of prepared samples (f).

1520

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

electrolyte and conductive additives, allowing easy access to the center of the material, and thus may ensure that Li+, electrons, and the active material are in the same place at the same time. Therefore, both a good working voltage and rate capability are guaranteed. A well-designed hierarchical structure in cathode materials can slow structural breakage by buffering the mechanical strain introduced by volume change during cycling to slow aging and can also relax side reactions, which often trouble thermodynamically unstable isolated nanoparticles with the same Li+ diffusion distance. Hence, the cycle stability of r-LNMO is also improved. EIS of r-LNMO further consolidates the above-mentioned deduction because rapid lithium-ion intercalation/deintercalation and fast electron transfer also means small impedance. As shown in Figure 5e, the impedance spectra of both r-LNMO and b-LNMO electrodes contained a partial semicircle in the high to medium frequency region and an inclined line in the low-frequency region. The semicircle in the high to medium frequency region is usually assigned to the charge transfer impedance (Rct), which is related to charge transfer through the electrode/electrolyte interface. Meanwhile, the quasi-straight line in the low frequency region represents the Warburg impedance, which reflects the solid-state diffusion of Li in electrode materials. According to Figure 5e, r-LNMO (Rct = 59.57 Ω) possessed a charge transfer resistance obviously lower than that of b-LNMO (Rct = 102.40 Ω). To be concise, a comparison table (Figure 5f) was appended to the end of Figure 5 to summarize the above four paragraphs and the thermal stability test described below. 3.3. Thermal Stability of r-LNMO. Thermal stability is an important safety related property for cathode materials considered for applications in commercial energy storage systems. Figure 6 shows the DSC curves of a r-LNMO

LiFePO4) were still maintained during the successive cycles, which is an important factor of a higher energy density for rLNMO. The discharge capacity retention after 25 cycles was 93%, which is among the best in previous reports.2,4,6,9,37 Cyclic voltammetry curves of a typical r-LNMO electrode measured between 2.0 and 4.8 V with a scanning rate of 0.1 mV s−1 are presented in Figure 4b and give a more detailed illustration of the electrochemical behavior of r-LNMO. More than one oxidation or reduction reaction occurred during cycling. In the first cycle, the first anodic peak around 3.7 V is associated with Ni oxidation from Ni2+ to Ni4+, while the second anodic peak at a higher potential (∼4.7 V) is attributed to irreversible Li2O extraction from the Li2MnO3 component to form MnO2. The reduction reactions remained almost the same. There were three distinguishable reduction peaks at ∼4.3, ∼3.7, and ∼3.3 V. The rather inconspicuous reduction peak at ∼4.3 and the distinct reduction peak ∼3.7 V are associated with the reduction of Ni4+, while the peak at ∼3.3 V can be assigned to the reduction of Mn4+. Hence, the reversible oxidation/ reduction reactions of r-LNMO involved both Ni and Mn. The changes in the cycle profile from the second cycle onward were quite smooth, indicating a good reversibility of r-LNMO. The contrasting electrochemical performances of r-LNMO and b-LNMO are shown in Figure 5. As shown in Figure 5a, the discharge capacity of r-LNMO at 1 C rate was 187.0 mA h g−1 after 50 cycles, while that of b-LNMO was 117.6 mA h g−1, representing 95.6 and 64.5% capacity retention, respectively. Meanwhile, the initial discharge capacity of r-LNMO was a little higher than that of b-LNMO. Thus, r-LNMO possessed both better cycle stability and higher capacity. Similar superiority of r-LNMO was also observed in the test under 0.2 C (Figure 5b). Meanwhile, Coulombic efficiency of r-LNMO was higher in both conditions. In addition to high specific capacity and good cycling stability, good rate capability is also crucial for cathode materials. Figure 5c presents the rate performance of both rLNMO and b-LNMO at a charge rate of 0.1 C and various discharge rates (0.1−5 C). The discharge capacities of the rLNMO electrode were 261.2, 225.6, 214.4, 188.5, 166.9, and 135.4 mA h g−1 at a current rate of 0.1, 0.2, 0.5, 1, 2, and 5 C, respectively. Notably, the electrode could still deliver a capacity of more than 135.4 mA h g−1 even at 5 C. Additionally, the discharge capacity of r-LNMO recovered to 250.6 mA h g−1 when the discharge rate was dropped back down to 0.1 C, close to its initial discharge capacity under 0.1 C rate. In contrast, the capacity drop observed for b-LNMO was much more drastic when the current rate was increased. The unique 3D reticular structure of r-LNMO is considered to be responsible for its superior electrochemical performance, namely, high capacity and excellent cycle stability and rate capability. A quick high temperature test under 60 °C was also performed (Figure 5d). It is feasible to deduce that r-LNMO possesses beneficial features similar to those of other kinds of nano/microhierarchical cathode materials because r-LNMO provides no less fine characteristics on the basis of its particle and subparticle structure. Specifically, the 3D reticular structure can increase the utilization of the cathode material and introduces new lithium storage space in its mesopores, leading to high capacity. The presence of mesopores and nanosized material dimensions can shorten the Li+ ion diffusion path and provide a larger electrode−electrolyte interface for Li+ to flux across, which results in rapid lithium-ion intercalation/ deintercalation and fast electron transfer. Ideally, the mesopore channel system in this structure provides room for the

Figure 6. DSC profile of 3D r-LNMO electrodes and electrodes constructed with ur-LNMO.

electrode and an electrode constructed with ultrasound-treated r-LNMO (ur-LNMO). The major exothermal peak of the rLNMO electrode emerged at 241.5 °C, about 10 °C higher than that of the ur-LNMO electrode. The high exothermal temperature (good thermal stability) indicates that the rLMNO had good structural stability and safety characteristics. The designed 3D reticular structure wins isolated thermodynamically unstable nanoparticles by combining them. Likely, this novel structure reduced the amount of side reactions between the active material and electrolyte, as seen for other kinds of effective hierarchical structures, improving its safety.6 1521

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces

Li1.2Ni0.2Mn0.6O2 by Analytical Electron Microscopy. J. Power Sources 2008, 178 (1), 422−433. (8) Hy, S.; Su, W. N.; Chen, J. M.; Hwang, B. J. Soft X-ray Absorption Spectroscopic and Raman Studies on Li1.2Ni0.2Mn0.6O2 for Lithium-Ion Batteries. J. Phys. Chem. C 2012, 116 (48), 25242−25247. (9) Li, Y.; Bai, Y.; Wu, C.; Qian, J.; Chen, G. H.; Liu, L.; Wang, H.; Zhou, X. Z.; Wu, F. Three-Dimensional Fusiform Hierarchical Micro/ Nano Li1.2Ni0.2Mn0.6O2 with a Preferred Orientation (110) Plane as a High Energy Cathode Material for Lithium-Ion Batteries. J. Phys. Chem. A 2016, 4 (16), 5942−5951. (10) Hong, J.; Seo, D. H.; Kim, S. W.; Gwon, H.; Oh, S. T.; Kang, K. Structural Evolution of Layered Li1.2Ni0.2Mn0.6O2 upon Electrochemical Cycling in a Li Rechargeable Battery. J. Mater. Chem. 2010, 20 (45), 10179−10186. (11) Chen, L.; Su, Y. F.; Chen, S.; Li, N.; Bao, L. Y.; Li, W. K.; Wang, Z.; Wang, M.; Wu, F. Hierarchical Li1.2Ni0.2Mn0.6O2 Nanoplates with Exposed {010} Planes as High-Performance Cathode Material for Lithium-Ion Batteries. Adv. Mater. 2014, 26 (39), 6756−6760. (12) Yan, P. F.; Nie, A. M.; Zheng, J. M.; Zhou, Y. G.; Lu, D. P.; Zhang, X. F.; Xu, R.; Belharouak, I.; Zu, X. T.; Xiao, J.; Amine, K.; Liu, J.; Gao, F.; Shahbazian-Yassar, R.; Zhang, J. G.; Wang, C. M. Evolution of Lattice Structure and Chemical Composition of the Surface Reconstruction Layer in Li1.2Ni0.2Mn0.6O2 Cathode Material for Lithium Ion Batteries. Nano Lett. 2015, 15 (1), 514−522. (13) Zhang, K.; Han, X.; Hu, Z.; Zhang, X.; Tao, Z.; Chen, J. Nanostructured Mn-Based Oxides for Electrochemical Energy Storage and Conversion. Chem. Soc. Rev. 2015, 44 (3), 699−728. (14) Feng, X.; Gao, Y. R.; Ben, L. B.; Yang, Z. Z.; Wang, Z. X.; Chen, L. Q. Enhanced Electrochemical Performance of Ti-Doped Li1.2Mn0.54Co0.13Ni0.13O2 for Lithium-Ion Batteries. J. Power Sources 2016, 317, 74−80. (15) Zhao, Y. J.; Sun, Y. C.; Yue, Y. Y.; Hu, X. S.; Xia, M. H. Carbon Modified Li-rich Cathode Materials Li1.26Fe0.22Mn0.52O2 Synthesized via Molten Salt Method with Excellent Rate Ability for Li-Ion Batteries. Electrochim. Acta 2014, 130, 66−75. (16) Qing, R.-P.; Shi, J.-L.; Xiao, D.-D.; Zhang, X.-D.; Yin, Y.-X.; Zhai, Y.-B.; Gu, L.; Guo, Y.-G. Enhancing the Kinetics of Li-rich Cathode Materials through the Pinning Effects of Gradient Surface Na+ Doping. Adv. Energy Mater. 2016, 6 (6), 1. (17) Li, L.; Wang, L.; Zhang, X.; Xie, M.; Wu, F.; Chen, R. Structural and Electrochemical Study of Hierarchical Lini1/3Co1/3Mn1/3O2 Cathode Material for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7 (39), 21939−21947. (18) Sun, M.-H.; Huang, S.-Z.; Chen, L.-H.; Li, Y.; Yang, X.-Y.; Yuan, Z.-Y.; Su, B.-L. Applications of Hierarchically Structured Porous Materials from Energy Storage and Conversion, Catalysis, Photocatalysis, Adsorption, Separation, and Sensing to Biomedicine. Chem. Soc. Rev. 2016, 45 (12), 3479−3563. (19) Alivisatos, A. P. Semiconductor Clusters, Nanocrystals, and Quantum Dots. Science 1996, 271 (5251), 933−937. (20) Namgung, R.; Mi Lee, Y.; Kim, J.; Jang, Y.; Lee, B. H.; Kim, I. S.; Sokkar, P.; Rhee, Y. M.; Hoffman, A. S.; Kim, W. J. Poly-Cyclodextrin and Poly-Paclitaxel Nano-Assembly for Anticancer Therapy. Nat. Commun. 2014, 5, 3702. (21) Link, S.; El-Sayed, M. A. Shape and Size Dependence of Radiative, Non-Radiative and Photothermal Properties of Gold Nanocrystals. Int. Rev. Phys. Chem. 2000, 19 (3), 409−453. (22) Roduner, E. Size Matters: Why Nanomaterials are Different. Chem. Soc. Rev. 2006, 35 (7), 583−592. (23) Burda, C.; Chen, X. B.; Narayanan, R.; El-Sayed, M. A. Chemistry and Properties of Nanocrystals of Different Shapes. Chem. Rev. 2005, 105 (4), 1025−1102. (24) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Synthesis and Characterization of Monodisperse Nanocrystals and Close-Packed Nanocrystal Assemblies. Annu. Rev. Mater. Sci. 2000, 30, 545−610. (25) Cademartiri, L.; Arsenault, A.; Ozin, G. A. Nanochemistry: A Chemical Approach to Nanomaterials, 2nd ed.; Royal Society of Chemistry: London, 2008.

4. CONCLUSION In summary, we successfully adopted a hard-templating route to synthesize 3D reticular Li1.2Ni0.2Mn0.6O2. The as-synthesized 3D reticular Li1.2Ni0.2Mn0.6O2 delivered high capacity, excellent cycle performance, and good rate capability as a lithium-ion battery cathode material. The excellent performance of the material is mainly ascribed to its function-promoting structure based on the nano-3D network and mesopore channel system combination. Simply put, this particular structure provides nanoscale lithium pathways, large specific surface area, good thermal and mechanical stability, and easy access to the center of the material. On the basis of this work, 3D reticular Li1.2Ni0.2Mn0.6O2 can be considered to be a potential cathode candidate for the development of future lithium-ion batteries. More importantly, the hard-templating strategy verified in this work may also provide an effective general approach to improve the cycle stability and rate capability of cathode materials for lithium-ion batteries.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; Tel: +86 10 68912508; Fax: +86 10 68451429. ORCID

Renjie Chen: 0000-0002-7001-2926 Author Contributions §

L. Li, L. Wang, and R. Chen contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The experimental work of this study was supported by the Chinese National 973 Program (Grant 2015CB251106), the Joint Funds of the National Natural Science Foundation of China (Grant U1564206), and the Major achievements Transformation Project for Central University in Beijing and Beijing Science and Technology Project (Grant D151100003015001).



REFERENCES

(1) Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334 (6058), 928− 935. (2) Zhang, X. F.; Meng, X. B.; Elam, J. W.; Belharouak, I. Electrochemical Characterization of Voltage Fade of Li1.2Ni0.2Mn0.6O2 Cathode. Solid State Ionics 2014, 268, 231−235. (3) Liu, Y. J.; Lv, J.; Liu, S. B.; Chen, L.; Chen, X. H. Improved Electrochemical Performance of Li1.2Ni0.2Mn0.6O2 Cathode Materials by Ball Milling and Carbon Coating. Powder Technol. 2013, 239, 461− 466. (4) Liu, Y. J.; Gao, Y. Y.; Lv, J.; Chen, L. A Facile Method to Synthesize Carbon Coated Li1.2Ni0.2Mn0.6O2 with Improved Performance. Mater. Res. Bull. 2013, 48 (11), 4930−4934. (5) Guan, X. T.; Ding, B.; Liu, X. F.; Zhu, J. J.; Mi, C. H.; Zhang, X. G. Enhancing the Electrochemical Performance of Li1.2Ni0.2Mn0.6O2 by Surface Modification with Nickel-Manganese Composite Oxide. J. Solid State Electrochem. 2013, 17 (7), 2087−2093. (6) Zhang, L. J.; Wu, B. R.; Li, N.; Mu, D. B.; Zhang, C. Z.; Wu, F. Rod-like Hierarchical Nano/Micro Li1.2Ni0.2Mn0.6O2 as High Performance Cathode Materials for Lithium-Ion Batteries. J. Power Sources 2013, 240, 644−652. (7) Lei, C. H.; Bareno, J.; Wen, J. G.; Petrov, I.; Kang, S. H.; Abraham, D. P. Local Structure and Composition Studies of 1522

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523

Research Article

ACS Applied Materials & Interfaces (26) Shi, X. X.; Liao, S. X.; Yuan, B.; Zhong, Y. J.; Zhong, B. H.; Liu, H.; Guo, X. D. Facile Synthesis of 0.6Li2MnO3-0.4LiNi0.5Mn0.5O2 with Hierarchical Micro/Nanostructure and High Rate Capability as Cathode Material for Li-Ion Battery. Acta Phys. -Chim. Sin. 2015, 31 (8), 1527−1534. (27) Kim, T.-W.; Kleitz, F.; Paul, B.; Ryoo, R. MCM-48-Like Large Mesoporous Silicas with Tailored Pore Structure: Facile Synthesis Domain in a Ternary Triblock Copolymer-Butanol-Water System. J. Am. Chem. Soc. 2005, 127 (20), 7601−7610. (28) Ren, Y.; Ma, Z.; Bruce, P. G. Ordered Mesoporous NiCoMnO4: Synthesis and Application in Energy Storage and Catalytic Decomposition of N2O. J. Mater. Chem. 2012, 22 (30), 15121−15127. (29) Taolin, Z.; Li, L.; Renjie, C.; Huiming, W.; Xiaoxiao, Z.; Shi, C.; Man, X.; Feng, W.; Jun, L.; Amine, K. Design of Surface Protective Layer of LiF/FeF3 Nanoparticles in Li-Rich Cathode for HighCapacity Li-Ion Batteries. Nano Energy 2015, 15, 164−176. (30) Shi, J. L.; Zhang, J. N.; He, M.; Zhang, X. D.; Yin, Y. X.; Li, H.; Guo, Y. G.; Gu, L.; Wan, L. J. Mitigating Voltage Decay of Li-Rich Cathode Material via Increasing Ni Content for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8 (31), 20138−20146. (31) Li, J. F.; Xiong, S. L.; Liu, Y. R.; Ju, Z. C.; Qian, Y. T. Uniform LiNi1/3Co1/3Mn1/3O2 Hollow Microspheres: Designed Synthesis, Topotactical Structural Transformation and Their Enhanced Electrochemical Performance. Nano Energy 2013, 2 (6), 1249−1260. (32) Bomben, K. D.; Moulder, J. F.; Sobol, P. E.; Stickle, W. F. Handbook of X-Ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of XPS Data; Physical Electronics: 1995. (33) Huang, Z. D.; Liu, X. M.; Oh, S. W.; Zhang, B. A.; Ma, P. C.; Kim, J. K. Microscopically Porous, Interconnected Single Crystal LiNi1/3Co1/3Mn1/3O2 Cathode Material for Lithium Ion Batteries. J. Mater. Chem. 2011, 21 (29), 10777−10784. (34) Marco, J. F.; Gancedo, J. R.; Gracia, M.; Gautier, J. L.; Rios, E.; Berry, F. J. Characterization of the Nickel Cobaltite, NiCo2O4 Prepared by Several Methods: An XRD, XANES, EXAFS, and XPS Study. J. Solid State Chem. 2000, 153 (1), 74−81. (35) Tuysuz, H.; Comotti, M.; Schuth, F. Ordered Mesoporous Co3O4 as Highly Active Catalyst for Low Temperature Co-oxidation. Chem. Commun. 2008, 34, 4022−4024. (36) Liu, H.; Du, X.; Xing, X.; Wang, G.; Qiao, S. Z. Highly Ordered Mesoporous Cr2O3 Materials with Enhanced Performance for Gas Sensors and Lithium Ion Batteries. Chem. Commun. 2012, 48 (6), 865−867. (37) Tang, D. C.; Liu, D. T.; Liu, Y. Y.; Yang, Z. Z.; Chen, L. Q. Investigation on the Electrochemical Activation Process of Li1.20Ni0.32Co0.004Mn0.476O2. Prog. Nat. Sci. 2014, 24 (4), 388−396.

1523

DOI: 10.1021/acsami.6b13229 ACS Appl. Mater. Interfaces 2017, 9, 1516−1523