Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Urea-assisted hydrothermal synthesis of a hollow hierarchical LiNi0.5Mn1.5O4 cathode material with tunable morphology characteristics

Xing Qinabc, Mushang Zhouabc, Bo Zongabc, Jianling Guoabc, Jiajia Gongabc, Li Wang*abc and Guangchuan Liangabc
aInstitute of Power Source and Ecomaterials Science, Hebei University of Technology, Tianjin 300130, China. E-mail: wangli_hebut@163.com
bKey Laboratory of Special Functional Materials for Ecological Environment and Information (Hebei University of Technology), Ministry of Education, Tianjin 300130, China
cKey Laboratory for New Type of Functional Materials in Hebei Province, Hebei University of Technology, Tianjin 300130, China

Received 8th July 2018 , Accepted 19th August 2018

First published on 24th August 2018


Abstract

A hollow hierarchical LiNi0.5Mn1.5O4 cathode material has been synthesized via a urea-assisted hydrothermal method followed by a high-temperature calcination process. The effect of reactant concentration on the structure, morphology and electrochemical properties of the carbonate precursor and corresponding LiNi0.5Mn1.5O4 product has been intensively investigated. The as-prepared samples were characterized by XRD, FT-IR, SEM, CV, EIS, GITT and constant-current charge/discharge tests. The results show that all samples belong to a cubic spinel structure with mainly Fd3m space group, and the Mn3+ content and impurity content initially decrease and then increase slightly with the reactant concentration increasing. SEM observation shows that the particle morphology and size of carbonate precursor can be tailored by changing reactant concentration. The LiNi0.5Mn1.5O4 sample obtained from the carbonate precursor hydrothermally synthesized at a reactant concentration of 0.3 mol L−1 exhibits the optimal overall electrochemical properties, with capacity retention rate of 96.8% after 100 cycles at 1C rate and 10C discharge capacity of 124.9 mA h g−1, accounting for 99.9% of that at 0.2C rate. The excellent electrochemical performance can be mainly attributed to morphological characteristics, that is, smaller particle size with homogeneous distribution, in spite of lower Mn3+ content.


Introduction

Lithium ion batteries (LIBs) are being intensively pursued for high-power and high-energy applications such as electric vehicles (EVs) and hybrid electric vehicles (HEVs).1,2 The energy density of LIB is generally limited by the cathode material. One practical and cost-effective way to improve the energy density of cathode material is elevating the working voltage. Among all cathode materials, LiNi0.5Mn1.5O4 spinel has been regarded as one of the most promising cathodes for LIB, owing to high working voltage (4.7 V vs. Li/Li+) and high energy density (650 W h kg−1).3,4 In addition, the inherently fast Li+ ion diffusion within its 3D spinel structure leads to good rate capability and cycling stability.5 However, simultaneously achieving satisfactory cyclability and rate performance for LiNi0.5Mn1.5O4 remains challenging due to complex performance-influencing factors and electrolyte/structure instability under high potential.6,7

To date, there are three main approaches to solve the above problem, such as surface modification, cation doping and particle size reduction to nanoscale level. The nanosized materials generally show improved rate performance by shorter Li+ ion diffusion channel and larger surface area. Unfortunately, low thermodynamic stability of nanoparticles results in electrochemical agglomeration and raises the risk of side reaction with electrolyte, thus leading to inferior cycling stability.8,9 Recently, the cathodes with hierarchical hollow structure have attracted considerable interests for the co-improvement of rate capability and cycle performance.10,11 On the one hand, this kind of structure can provide a shorter path for Li+ ion diffusion and meanwhile have larger electrode/electrolyte contact area for Li+ flux across the interface, resulting in better rate capability. On the other hand, the hollow interior can improve the structural integrity by alleviating the mechanical strain induced by volume change during the repeated Li+ ion insertion/extraction processes, which can promote the cycling stability.5,10,12–16

At present, the LiNi0.5Mn1.5O4 cathode material with hollow hierarchical structure has been generally synthesized by high-temperature calcination using MnCO3,11,15,17 manganese oxides converted from MnCO3,5,10,18 or Ni0.25Mn0.75CO3 (ref. 19–21) as self-template and Mn/Ni source. In the above literatures, most of MnCO3 or Ni0.25Mn0.75CO3 has been synthesized by direct precipitation method between NH4HCO3/Na2CO3 precipitant and Mn/Ni salt solution, which may lead to the inhomogeneous particle size distribution of as-prepared carbonate precursor and corresponding LiNi0.5Mn1.5O4 product. To address above-mentioned problem, in this work, the Ni–Mn carbonate precursor was initially synthesized via a homogenous precipitation-hydrothermal method using urea as a precipitant reservoir, and LiNi0.5Mn1.5O4 cathode material with hollow hierarchical structure was achieved by high-temperature calcination process with lithium source. As is well known, the key point to fabricate hierarchical material is control of the precursor. During the hydrothermal process, the urea is slowly decomposed to generate precipitating agent (CO32−) in the reaction mixture, allowing particle ripening during precipitation and often resulting in better crystallinity, regular shape and size of particles.8 There are many influencing factors during the hydrothermal process, such as hydrothermal temperature, hydrothermal duration, molar ratio of urea to transition metal ions, solvent, reactant concentration, etc. The hydrothermal temperature, hydrothermal duration, molar ratio of urea to transition metal ions and solvent have been investigated in our previous work, which have been determined as 180 °C, 10 h, 2[thin space (1/6-em)]:[thin space (1/6-em)]1, mixed solvent of H2O and ethylene glycol (v/v = 6[thin space (1/6-em)]:[thin space (1/6-em)]1), respectively. Although there are many reports on the urea-assisted hydrothermal method,22–28 the effect of reactant concentration on the structure, morphology and electrochemical properties of carbonate precursor and corresponding product was rarely involved, which was just the focus of this work. It has been found that the reactant concentration exerts great influence on the particle morphology and size of carbonate precursors, thus resulting in different electrochemical properties of corresponding LiNi0.5Mn1.5O4 products.

Experimental

Sample synthesis

Firstly, the carbonate precursor was synthesized by a urea-assisted hydrothermal method. The specific preparation process was just as follows. Ni(CH3COO)2·4H2O, Mn(CH3COO)2·4H2O and urea were dissolved in 75 mL of a mixture of deionized water and EG (6[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) under stirring, then transferred to a 100 mL Teflon-lined stainless-steel autoclave and heated at 180 °C for 10 h. The total molar concentration of transition metal ions (reactant concentration) was selected as 0.1, 0.2, 0.3 and 0.4 mol L−1, respectively. The molar ratio of urea to transition metal ions was set as 2[thin space (1/6-em)]:[thin space (1/6-em)]1. After cooling to room temperature naturally, the resulting carbonate precursor was centrifuged, and washed with water and ethanol several times. The dried carbonate precursor and stoichiometric amount of Li2CO3 (5% excess to make up for the lithium volatilization during calcination) were dispersed in ethanol, and the ethanol was evaporated slowly at room temperature under continuous stirring. Finally, the mixture was calcined at 750 °C for 12 h in the air to obtain the LiNi0.5Mn1.5O4 product. The carbonate precursors synthesized at reactant concentrations of 0.1, 0.2, 0.3, 0.4 mol L−1 were labelled as NMC-0.1, NMC-0.2, NMC-0.3, NMC-0.4, and the resulting LiNi0.5Mn1.5O4 products were named as LNMO-0.1, LNMO-0.2, LNMO-0.3, LNMO-0.4, respectively.

Sample characterization

XRD and FT-IR were performed to analyse the structural properties of as-synthesized carbonate precursor and LiNi0.5Mn1.5O4 powders by Bruker D8-FOCUS X-ray diffractometer (CuKα radiation) and Bruker V80 Fourier Transformation Infrared Spectrometer, respectively. SEM was performed to observe the particle morphology and size on a FEI Nova Nano SEM450 instrument. Brunauer–Emmett–Teller (BET) analysis of the sample was measured with an ASAP 2460 surface area and pore size analyser. The pore size distribution was calculated by the Barrett–Joyner–Halenda (BJH) method using nitrogen adsorption data.

The Ni/Mn content in LiNi0.5Mn1.5O4 product was measured by a chemical titration method, and the residual Ni2+ content in the filtrate was measured by a spectrophotometer (T6, PERSEE, Beijing). The detailed procedures are illustrated in ESI.

Electrochemical measurements

The electrochemical measurements were carried out using CR2032 coin cells. Test electrodes were prepared as follows: 80 wt% as-prepared LiNi0.5Mn1.5O4 powder, 15 wt% ethylene black and 5 wt% PTFE were mixed in alcohol and then ultrasonically dispersed uniformly to form a slurry. After the evaporation of alcohol, the mixture was rolled to a thin film with a thickness of 0.14 mm. The round electrode plate with a diameter of 12 mm and active material loading of ∼10 mg was cut down for coin assembly. The Li metal, Celgard 2400 and a solution of 1 M LiPF6 in EC[thin space (1/6-em)]:[thin space (1/6-em)]DMC[thin space (1/6-em)]:[thin space (1/6-em)]EMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 in volume) were used as the counter electrode, separator and electrolyte, respectively. The coin cells were assembled in an argon-filled glove box. For rate capability tests, the cells were charged and discharged between 3.5 and 4.95 V (vs. Li/Li+) at 0.2C, 1C, 5C and 10C rates. For cycling experiments, the cells were charged and discharged at 1C rate. The electrochemical measurement was carried out using Land battery measurement system (Wuhan, China) at room temperature. In order to calculate the lithium ion diffusion coefficient during charge/discharge process, galvanostatic intermittent titration technique (GITT) was also performed on this apparatus at 0.1C. The cell was charged/discharged for 10 min and then followed by a relaxation time of 30 min between 3.5 and 4.95 V during the 3rd change/discharge process after two activation cycles at 0.2C rate. The electrochemical impedance spectroscopy (EIS) was measured using an electrochemical workstation (CHI660D) in the frequency range of 100 kHz to 10 mHz with an AC amplitude of 5 mV at full-discharged state after cycling 3 times at 1C rate. The cyclic voltammetry (CV) was measured from 3.4 to 5.0 V at scan rate of 0.1 mV s−1.

Results and discussion

Structure of samples

Fig. 1(a) shows the XRD patterns of carbonate precursors synthesized at different reactant concentrations. It can be seen that the diffraction patterns of all four precursors can be indexed to a typical carbonate structure based on MnCO3 (JCPDS card no.: 44-1472) and NiCO3 (JCPDS card no.: 12-0771). It is inferred that the precursors synthesized at different reactant concentrations consist of both MnCO3 and NiCO3 phases. The diffraction peak becomes gradually broadened with the increase of reactant concentration, implying a gradual decrease in crystal size, which can be verified by the following SEM observation in Fig. 3. In addition, the peak intensity of NiCO3 is gradually decreased with the reactant concentration, implying the gradually incomplete precipitation of NiCO3 (Ksp = 1.42 × 10−7, 25 °C) due to its higher solubility product than MnCO3 (Ksp = 2.24 × 10−11, 25 °C) and higher complex equilibrium constant of [Ni(NH3)n]2+ (lg[thin space (1/6-em)]Kf = 8.12 for n = 4) than [Mn(NH3)n]2+ (lg[thin space (1/6-em)]Kf = 1.3 for n = 4).29 This can be verified by the remaining Ni2+ contents in the filtrate measured by spectrophotometry method, which are 7.14, 35.45, 63.32 and 105.36 mg L−1, respectively, for NMC-0.1, NMC-0.2, NMC-0.3 and NMC-0.4 samples. The contents of Ni and Mn in LiNi0.5Mn1.5O4 product were measured via the chemical titration method as illustrated in ESI. The obtained molar ratios of Ni and Mn are 0.495[thin space (1/6-em)]:[thin space (1/6-em)]1.5, 0.490[thin space (1/6-em)]:[thin space (1/6-em)]1.5, 0.485[thin space (1/6-em)]:[thin space (1/6-em)]1.5 and 0.479[thin space (1/6-em)]:[thin space (1/6-em)]1.5, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples, in consistence with the residual Ni2+ concentrations in the filtrate. It can be seen that all four samples show different levels of nickel deficiency compared to the stoichiometric LiNi0.5Mn1.5O4, and the nickel deficiency becomes gradually more severe with the reactant concentration.
image file: c8ra05817b-f1.tif
Fig. 1 XRD patterns of carbonate precursors (a) and LiNi0.5Mn1.5O4 samples (b) synthesized at different reactant concentrations.

image file: c8ra05817b-f2.tif
Fig. 2 FT-IR spectra of LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations.

image file: c8ra05817b-f3.tif
Fig. 3 SEM images of carbonate precursor (a–d) and LiNi0.5Mn1.5O4 samples (e–k) (a and e) 0.1 mol L−1, (b, f and i) 0.2 mol L−1, (c, g and j) 0.3 mol L−1, (d, h and k) 0.4 mol L−1.

Fig. 1(b) shows the XRD patterns of LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations. It can be seen that all four patterns can be assigned to cubic spinel LiNi0.5Mn1.5O4 (JCPDS card no.: 80-2162). Some weak peaks at 2θ = 37.5, 43.6 and 63.4° (marked with arrow) can be ascribed to a rock-salt impurity phase (LixNi1−xO, and/or Ni6MnO8),30,31 which can be clearly seen in local enlarged image between 35° and 45° in Fig. S1. The content of impurity phase can be roughly estimated by calculating the ratio of peak intensity at 2θ = 43.6° of impurity phase to that at 2θ = 44.3° of spinel phase.32 The intensity ratios are 3.62[thin space (1/6-em)]:[thin space (1/6-em)]100, 3.20[thin space (1/6-em)]:[thin space (1/6-em)]100, 2.96[thin space (1/6-em)]:[thin space (1/6-em)]100 and 3.36[thin space (1/6-em)]:[thin space (1/6-em)]100, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples, that is, the impurity content decreases in the order of LNMO-0.1 > LNMO-0.4 > LNMO-0.2 > LNMO-0.3. Among them, the LiNi0.5Mn1.5O4 sample synthesized at reactant concentration of 0.3 mol L−1 has the lowest impurity content, which is believed to be advantageous to electrochemical property, because it is recognized that the presence of rock-salt impurity phase may lower discharge capacity and block Li+ mobility in the material.33 What's more, the variation rule of impurity phase content is the same as that of Mn3+ content, so the reason will be given in the following discussion of Mn3+ content.

Table 1 lists the lattice parameters of four samples obtained from the Rietveld refinement results, as shown in Fig. S2, based on Fd3m space group. From the refinement factors listed in Table S1, it can be clearly seen that the refinement yields good agreement factors, with Rwp values less than 10% and GOF values between 1 and 2, suggesting an acceptable refinement. As shown in Table 1, the lattice parameter a is 8.1708, 8.1688, 8.1666 and 8.1669 Å for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples, respectively. It can be seen that the lattice parameter initially decreases and then increases slightly with the reactant concentration increasing, which may be caused by the variation of Mn3+ content, as Mn3+ (0.645 Å) has a larger ionic radius than Mn4+ (0.53 Å). The variation of Mn3+ content is reported to be closely related to the particle size. Xue et al.32 found that under the same calcination condition, the different Mn3+ content in spinel sample could be ascribed to an inhomogeneous migration of atoms. During calcination process, Ni and Mn atoms have to migrate a distance to form a homogeneous LiNi0.5Mn1.5O4 phase. The shorter the distance, the more uniform the diffusion. The smaller particle size means shorter migration distance of atoms, thus leading to more homogeneous Ni/Mn distribution and lower Mn3+ content. Due to the smaller particle size and more homogeneous distribution, LNMO-0.3 sample exhibits the lowest Mn3+ content. As for LNMO-0.4 sample, the slightly higher Mn3+ content may be ascribed to the existence of some larger agglomerate particles.

Table 1 Lattice parameters and I311/I400 ratios of LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations
Sample a (Å) V3) I311/I400
LNMO-0.1 8.1708 545.51 1.038
LNMO-0.2 8.1688 545.09 0.880
LNMO-0.3 8.1666 544.67 0.991
LNMO-0.4 8.1669 544.72 0.942


For spinel LiMn2O4, the I311/I400 peak intensity ratio can reflect the degree of tetragonal distortion from cubic spinel structure.34 Therefore, the I311/I400 ratio can also reflect the structural integrity of LiNi0.5Mn1.5O4 spinel structure.35,36 The smaller the I311/I400 ratio, the better the structural stability. As shown in Table 1, the I311/I400 ratios are 1.038, 0.880, 0.991 and 0.942 for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples, respectively. It is reasonable to infer that LNMO-0.2 sample should have the best structural stability, which can be verified by the following cycling performance curves in Fig. 6.


image file: c8ra05817b-f4.tif
Fig. 4 Charge/discharge curves at various rates (a) and rate capability curves (b) for LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations.

image file: c8ra05817b-f5.tif
Fig. 5 Cyclic voltammograms of LiNi0.5Mn1.5O4 samples at scan rate of 0.1 mV s−1 between 3.4 and 5.0 V.

image file: c8ra05817b-f6.tif
Fig. 6 Cycling performance and coulombic efficiency of LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations at 1C rate.

It is well known that the LiNi0.5Mn1.5O4 spinel takes on two distinct space groups according to the ordering of transition metal cations: the simple cubic P4332 space group with Ni2+ and Mn4+ ions occupying the 12d and 4b sites, respectively, and the face centered cubic Fd3m space group with Ni2+ and Mn4+ ions randomly distributing on 16d sites.37,38 XRD characterization cannot distinguish between the two space groups due to the similar scattering factors of Ni and Mn, which can be accomplished by FT-IR spectra. As mentioned by Kunduraci et al.,39 P4332 space group has eight absorption peaks in FT-IR spectrum, while Fd3m space group only has five ones, and the intensity ratio of I(576 cm−1)/I(615 cm−1) increases with the cation ordering degree. From Fig. 2 it can be seen that all samples can be classified as Fd3m space group. The intensity ratios of I(576 cm−1)/I(615 cm−1) for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNNO-0.4 samples are 0.930, 0.971, 1.002 and 0.994, respectively, that is, the disordering degree (Mn3+ content) initially decreases and then increases slightly with the increase of reactant concentration, in accordance with XRD result.

Morphology of samples

As is well known, the structure and morphology have great influences on the electrochemical performance of electrode materials for LIB. The SEM images of the carbonate precursors prepared with different reactant concentrations are shown in Fig. 3(a–d). It can be seen that the precursor samples exhibit a gradually decreasing particle size with the increase of reactant concentration. Generally, the growth process of crystals can be separated into two steps, an initial nucleation stage and a subsequent crystal growth process. It is believed that chemical precipitation is a process of competition between crystal nucleation and crystal growth, which has a great relationship with the particle size. The increase of reactant concentration will lead to an increasing nucleation rate, thus leading to the gradual decrease in particle size. However, as for NMC-0.4 sample, there exists some large particles (3–4 μm in diameter) caused by the agglomeration of small particles, leading to its nonuniform particle size distribution. Among the four precursors, NMC-0.3 sample shows the most homogeneous particle size distribution with smaller average particle size of 3 μm. It is generally accepted that both particle size and distribution homogeneity exert an influence on the electrochemical performance of final product, therefore better electrochemical performance of LNMO-0.3 sample can be expected.

In addition, the particle morphology also varies with the reactant concentration. From Fig. 3 we can see that NMC-0.1 sample exhibits shuttle-like morphology, NMC-0.2 and NMC-0.3 samples show peanut-like morphology, or rather, the combination of two spherical particles with each other, and NMC-0.4 sample shows spherical morphology. It can be seen that the sphericity of the particle gradually increases with reactant concentration, in other words, the higher the reactant concentration, the easier it is to form spherical particles, in consistence with the finding of ref. 40. Bai et al.40 synthesized a three-dimensional fusiform hierarchical micro/nano Li-rich Li1.2Ni0.2Mn0.6O2 cathode material through a hydrothermal method followed by heat-treatment process. They suggested that in order to make sure the morphology of material to be non-spherical, the amount of transition metal ion should be low, because higher the concentration of reagent is, easier to make the crystal nucleation grow to be sphere-shape, which is in consistence with this work.

SEM images of the as-prepared LiNi0.5Mn1.5O4 samples are shown in Fig. 3(e–k). It can be seen that the LiNi0.5Mn1.5O4 samples exhibit a good inherited morphology from carbonate precursors, that is, shuttle-like morphology for LNMO-0.1, peanut-like morphology for LNMO-0.2 and LNMO-0.3, as well as spherical morphology for LNMO-0.4. Compared with corresponding carbonate precursor, the LiNi0.5Mn1.5O4 product exhibits a relatively smaller secondary particle size caused by volume shrinkage during calcination and a relatively larger primary particle size resulting from the fusion of subunits, whose surface is much rougher than before calcination with a significant fraction of the particles having cracks. But, by and large, with the increase of reactant concentration, the particle morphology and size of LiNi0.5Mn1.5O4 product vary in a similar trend of carbonate precursor. Therefore, it is a feasible way to tailor the particle morphology and size of LiNi0.5Mn1.5O4 by tailoring their precursors' morphology and size, which can be controlled by reactant concentration.

The typical broken hollow particles of LiNi0.5Mn1.5O4 samples synthesized at 0.2–0.4 mol L−1 reactant concentrations are shown in Fig. 3(i–k), the magnified image of the region marked with a circle in Fig. 3(f–h), and the hollow interior can be clearly observed from the broken part. From the broken particle in Fig. 3(i–k), the wall of LiNi0.5Mn1.5O4 hollow particles is composed of interconnected submicron polyhedral subunits. It is worth noting that the wall thickness of LiNi0.5Mn1.5O4 samples synthesized at 0.2–0.4 mol L−1 reactant concentrations is about 0.7, 0.6 and 0.3 μm, respectively, that is, the wall thickness gradually decreases with reactant concentration. It is speculated that the wall of LNMO-0.1 sample is so thick that the hollow structure cannot be observed from Fig. 3(e), and that of partial LNMO-0.4 breaks into small particles due to thinner wall, as shown in Fig. 3(k). The variation of wall thickness can be explained as follows. It has been reported that the particles in the core of carbonate precursor is formed by nucleation process, which have a relatively smaller size and a lower density due to the initial higher supersaturation. However, the wall is formed by gradual crystal growth process, whose compositional particles have a relatively larger size and a higher density.41 During high-temperature calcination, the carbonate precursor will go through volume shrinkage to obtain LiNi0.5Mn0.5O4 particle. So the core with low density becomes sparse or even disappeared, and the wall with high density is still compact in LiNi0.5Mn0.5O4. The increase of reactant concentration results in the increased nucleation rate, thus leading to the gradual decrease in wall thickness of LiNi0.5Mn1.5O4 product. Wang et al.41 thought that the formation of hollow structure may be caused by Ostwald ripening. Due to the high surface energy of the small crystallites in core compared to the larger ones in wall, the smaller crystallites in the cores transferred to the wall during calcination process. This hollow hierarchical structure can provide a larger surface area and a reduced diffusion distance for both Li+ ions and electrons, and the internal void space can effectively accommodate the volume change resulting from the repeated Li+ ion intercalation/deintercalation, which is favourable to electrochemical performance.

To further confirm the hollow hierarchical structure of as-prepared samples, nitrogen adsorption/desorption measurements were performed. Fig. S3 shows the adsorption/desorption isotherms and corresponding BJH pore size distribution plots of the materials. According to IUPAC classification, all isotherms can be attributed to type IV, indicating the presence of mesopores. From the pore size distribution plots it can be seen that there are abundant mesopores ranged from 2 to 20 nm existing in the materials. The loops observed can be ascribed to type H1, corresponding to the particle assembly, in consistence with the particle structure of as-prepared samples. The surface areas obtained according to BET method are 2.386, 3.616, 4.262 and 3.935 m2 g−1, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples. Using the BJH method and the desorption branch of the nitrogen isotherm, the cumulative pore volumes are 8.09 × 10−3, 12.80 × 10−3, 14.95 × 10−3 and 13.96 × 10−3 cm3 g−1 for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples, respectively. The relatively larger specific surface area and higher porosity of LNMO-0.3 sample could offer larger electrode/electrolyte contact area and promote Li+ ion diffusion,42,43 therefore, better electrochemical performance can be expected.

Electrochemical properties

To evaluate the rate capability, the cells were tested from 0.2C to 10C in the voltage range of 3.5 to 4.95 V and the charge rates were equal to the discharge rates. At each rate, a constant-current charge step was followed by a constant-voltage step until the current decreased to 1/10 of the applied current. Fig. 4(a) displays the charge/discharge curves of as-prepared LiNi0.5Mn1.5O4 samples at rates of 0.2C, 1C, 5C and 10C. All discharge curves show a long plateau at ∼4.7 V, arising from Ni2+/Ni4+ redox couple, as well as a short plateau at ∼4.0 V, associated with Mn3+/Mn4+ redox couple. The presence of 4.0 V plateau implies that the as-prepared LiNi0.5Mn1.5O4 samples mainly have disordered structure with Fd3m space group, in consistence with FT-IR result. Based on the discharge curves at 0.2C rate, the Mn3+ contents can be calculated from the discharge capacity between 3.8 and 4.25 V being divided by the total discharge capacity,44 which are 12.06%, 11.33%, 9.34% and 10.62%, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples. It can be seen that the Mn3+ content first decreases and then increases with the reactant concentration, and LNMO-0.3 sample has the lowest Mn3+ content, in consistence with the above XRD and FT-IR results. In addition, with increasing current density, the voltage difference between charge and discharge plateau increases due to the limited Li+ ion diffusion kinetics, meaning increasing polarization. A smaller voltage difference between charge and discharge plateau indicates better rate capability. Besides, the rate capability can also be evaluated by proportion of constant-current charge step in total charge capacity at large currents.11 The constant-current charge step contributes 87.3%, 92.0%, 92.1% and 88.8% to the total charge capacity at 5C in LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4, respectively. The above results indicate that LNMO-0.3 sample exhibits the optimal rate capability, whose discharge capacities at 0.2C, 1C, 5C and 10C are 125.0, 128.5, 127.4 and 124.9 mA h g−1, respectively. The discharge capacity at 10C rate accounts for 99.9% of that at 0.2C rate, higher than 99.2% for LNMO-0.1, 98.5% for LNMO-0.2 and 96.3% for LNMO-0.4.

To further evaluate the rate performance of samples, the cells were cycled successively at current densities from 0.2C to 10C, five times at each current density, as shown in Fig. 4(b). Interestingly, the capacities for all electrodes at 1C is higher than 0.2C, which is consistent with the finding of ref. 45. They reported that training at stepwise increasing discharge rates (0.2–1C, up to ca. tenth cycle) significantly improved the discharge capacities of the electrodes, probably due to better accessibility of deeper parts of LiNi0.5Mn1.5O4 crystallites and formation of a solid electrolyte interphase layer on the inner parts. From the figure, we can see that LNMO-0.3 sample exhibits higher discharge capacities at all test rates, manifesting its better rate capability, which can be mainly attributed to its smaller particle size and more homogenous distribution, in spite of lower Mn3+ content. In addition, the lower impurity content also contributes to its better rate capability. The worst rate performance of LNMO-0.4 sample may be partly ascribed to the existence of many small particles, because it has been reported that when particles are too small, only a small number of particles can directly contact the conductive additive and many interfaces exist between the particles.41 It can be concluded that excessive decrease in particle size is unfavorable to rate capability, which is consistent with the finding of ref. 41. In addition, the hollow hierarchical structure with connected primary subunits can facilitate rapid Li+ ion diffusion and efficient mass-electrolyte contact because of the significantly shortened diffusion route and a large surface-to-volume ratio, which may also be conductive to rate capability.

Fig. 5 shows the cyclic voltammograms of as-prepared LiNi0.5Mn1.5O4 samples between 3.4 and 5.0 V at scan rate of 0.1 mV s−1. For all curves, there appear redox peaks at about 4.0 V corresponding to Mn3+/Mn4+ reaction and at about 4.7 V corresponding to Ni2+/Ni4+ reaction, in accordance with the charge/discharge curve in Fig. 4(a). This is characteristic of LiNi0.5Mn1.5O4 with Fd3m space group.46,47 It is noteworthy that a minor redox peak appears at around 4.9 V, which is reported to be caused by the electrolyte decomposition.48 Moreover, it has been reported that LiNi0.5Mn1.5O4 with a disordered structure shows two split peaks at around 4.7 V while that an ordered structure only gives one peak at around 4.7 V,49 which also confirms the disordered Fd3m space group of as-prepared samples, consistent with the FT-IR analysis. From the inset magnified image of 4.0 V peak, it can be seen that the Mn3+ content decreases in the order of LNMO-0.1 > LNMO-0.2 > LNMO-0.4 > LNMO-0.3, in consistence with the Mn3+ content calculated from initial discharge curves at 0.2C rate in Fig. 4(a).

The voltage differences between anodic (φa) and cathodic (φc) peaks reflect the polarization degree of the electrode, which is listed in Table S2. The bigger the potential difference between Li+ ions intercalation and deintercalation is, the stronger the electrode polarization is. It can be observed that LNMO-0.3 sample exhibits the smallest potential difference (for Ni2+/Ni3+ and Ni3+/Ni4+) among all samples, indicating its best reversibility of Li+ ions during intercalation/deintercalation process, which in turn ensures its optimal rate capability and cycling performance.50 In addition, the higher peak intensity of LNMO-0.3 sample also confirms its higher specific capacity, in consistence with Fig. 4(a).

Cycling performance is also of significant importance for powering LIB. Fig. 6 shows cycling performance and coulombic efficiency of LiNi0.5Mn1.5O4 materials in the voltage range of 3.5–4.95 V at 1C rate and room temperature. From the initial charge/discharge curves in Fig. S4, we can get that the initial coulombic efficiencies are 89.7%, 89.5%, 90.7% and 89.9%, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples. The coulombic efficiencies of all four samples are gradually improved in initial several cycles and then stabilized at ∼99%. The formation of solid electrolyte interphase (SEI) film results in the relatively low coulombic efficiency during initial cycles, which is of great importance in preventing electrolyte decomposition, thus leading to the stabilization of coulombic efficiency in the following cycles. In consideration of the capacity increase in initial several cycles, the 5th cycle was chosen for relative capacity retention. Based on Fig. 6 we can get that the discharge capacities at 5th cycle and 100th cycle are 122.6/118.1, 125.7/123.7, 130.8/126.6 and 126.5/122.4 mA h g−1, and the corresponding capacity retention rates are 96.3%, 98.4%, 96.8% and 96.8%, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples. It can be seen that LNMO-0.2 sample shows the best cycling performance, which is consistent with its highest I311/I400 intensity ratio obtained from XRD pattern.

It has been reported that the factors influencing the cycling performance of LiNi0.5Mn1.5O4 cathode material mainly include the side reactions with electrolyte51 and Mn dissolution at high voltage,52,53 as well as the presence of rock-salt impurity phase,30,33,47 etc. The side reaction is closely relevant to the contact area between electrode and electrolyte, and the higher the contact area, the more severe the side reaction with electrolyte. The Mn dissolution mainly results from the disproportionation reaction of Mn3+ (2Mn3+ = Mn2+ + Mn4+),54 that is, the existence of Mn3+ in the cathode is the main reason for the Mn dissolution,55 therefore higher Mn3+ content may lead to severe Mn dissolution and severe strain from Jahn–Teller distortion.3,56,57 In addition, it has been reported that the Mn dissolution of LiMn2O4 spinel is often accelerated with larger specific surface area,58,59 which should be also applicable to LiNi0.5Mn1.5O4, a derivative of LiMn2O4. The impurity phase tends to react with the electrolyte and thus causes capacity fading.47 Based on above analysis, it can be inferred that the better cycling performance of LNMO-0.2 sample can be attributed to its appropriate particle size, Mn3+ content and impurity content. The inferior cycling performance of LNMO-0.1 sample may be attributed to the higher Mn3+ and impurity contents, while that of LNMO-0.3 and LNMO-0.4 samples may be mainly ascribed to the smaller particle size.

But then again, the as-prepared LiNi0.5Mn1.5O4 samples with hollow hierarchical structure show superior cycling performance to that of bulk electrode,31,60–62 which can be mainly attributed to that the small primary particles and interior void space allow for facile penetration by the electrolyte and effectively release the strain caused by the volume change during the repeated Li+ insertion/extraction cycles.5,63 The superior structural integrity of hollow structure can also be confirmed by the SEM image of four electrodes after cycling 100 times at 1C rate, as shown in Fig. S5. The very small particles in the image are conductive additive (acetylene black). Compared with the original SEM images in Fig. 3, all samples can maintain their original morphology after 100 cycles, and the hollow structure of LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples is still visible. It means that the hollow structure of all samples is very stable, which is not broken in the charge/discharge cycling process. Therefore, their excellent cycling performance can be partly ascribed to buffering effect of the hollow interior during the repeated Li+ ion insertion/extraction processes. Taking the discharge capacity, rate capability and capacity fading factor into account, LNMO-0.3 sample shows a balance of electrochemical properties, which agrees with the deduction obtained from Fig. 3.

Electrochemical impedance spectroscopy (EIS) is an important technique for evaluating the interfacial electrochemistry as well as the reaction kinetics in LIB materials. Impedance spectra were conducted on the LiNi0.5Mn1.5O4/Li cells to further understand the different electrochemical performances of LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations. Nyquist plots were obtained in the frequency range from 100 kHz to 10 mHz at full-discharged state after 3 cycles at 1C rate, as shown in Fig. 7. It can be seen that each plot contains a semicircle and an oblique line. The intercept at Z′ axis is relative to the ohmic resistance (Re) of electrolyte. The semicircle at the high-to-medium frequency region is attributed to the impedance (Rf) and constant phase element (CPE-1) of the SEI film, and the charge-transfer impedance (Rct) and constant phase element (CPE-2) of the electrode/electrolyte interface. While the oblique line at the low frequency region can be ascribed to the Warburg impedance (W0) related to Li+ ion solid-state diffusion in the electrode.64 The Nyquist plots were fitted by ZView software based on the equivalent circuit in Fig. S6, and the fitting lines match with the experimental data very well, as shown in Fig. 7. From the fitting parameters listed in Table S3, it can be clearly observed that the difference of Re and Rf between the four electrodes is negligible. Meanwhile, the charge-transfer resistance (Rct) exhibits considerable difference between these four electrodes, which are greatly higher than the resistances of SEI film, indicating that the charge-transfer resistance is a dominant factor of the electrochemical kinetics of electrode. As shown in Table S3, the charge-transfer resistance is 193.40, 63.47, 38.50 and 143.90 Ω for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples after 3 cycles at 1C rate, respectively, that is, the charge-transfer resistance first decreases and then increases with the reactant concentration increasing. Among them, LNMO-0.3 sample has the lowest charge-transfer resistance, indicative of the fastest Li-intercalation kinetics. In general, for LIB cathode materials, electronic and ionic conductivity are two key aspects to the charge-transfer resistance. In detail, the materials able to exhibit higher electronic and ionic conductivity possess better rate performances.65 Therefore, the lower charge-transfer resistance and higher Li+ ion diffusion coefficient are accountable for the high capacity and excellent rate performance of LNMO-0.3 sample, which can be mainly attributed to the smaller particle size with homogeneous distribution and lowest impurity content, despite a less disordered structure.


image file: c8ra05817b-f7.tif
Fig. 7 Nyquist plots for LiNi0.5Mn1.5O4 samples synthesized at different reactant concentrations.

The better cycling performance of samples is further supported by the EIS spectra after 100 cycles at 1C rate, which are shown in Fig. S7. The fitted results obtained by ZView software are shown in Table S3. It can be seen that both SEI film resistance and charge-transfer resistance are increased after cycling 100 times in comparison with cycling 3 times, and the charge-transfer resistances are 203.50, 77.41, 48.36 and 158.00 Ω, respectively, for LNMO-0.1, LNMO-0.2, LNMO-0.3 and LNMO-0.4 samples. It can be seen that in comparison with the EIS spectra after cycling 3 times (Fig. 7), the Rct values increase slightly, which suggests that the interparticle resistance of the electrode is suppressed by the hollow hierarchical structure. Therefore, the electrons and Li+ ions can transfer efficiently on the interface of active material and electrolyte, thus resulting in enhanced electrode reaction kinetics and excellent cycling performance of the samples.

In order to determine the Li+ ion diffusion coefficients (DLi+), galvanostatic intermittent titration technique (GITT) was carried out for all electrodes. In this experiment, a small current pulse (0.2C) was applied for 10 min and the cell was then allowed to rest at open circuit for 30 min. Fig. S8(a) shows the GITT curves of as-prepared LiNi0.5Mn1.5O4 samples in the 3rd charge/discharge process. Obviously, LNMO-0.3 samples has smaller polarization than other samples. Fig. S8(b) shows a typical E vs. τ profile at a selected titration for LNMO-0.3 sample. According to the GITT measurement, the chemical diffusion coefficient of Li+ ion can be calculated by the following eqn (1):66,67

 
image file: c8ra05817b-t1.tif(1)
where VM is the molar volume of samples which is deduced from the crystallographic data, M and m are the molecular weight and the mass of active material. A is the electrode contact area with the electrolyte obtained from the BET surface area, and L is the thickness of the electrode. τ is the constant current pulse time and ΔEs is the difference between the steady potentials. Since E vs. τ1/2 is linear over the entire time period of current flux (Fig. S8(c)), then eqn (1) can be further simplified as eqn (2):68,69
 
image file: c8ra05817b-t2.tif(2)
in which, ΔEτ is the total transient voltage change of the cell for an applied galvanostatic current for the time τ. Based on eqn (2), the obtained DLi+ values at different potentials during the 3rd discharge process are presented in Fig. 8. It can be seen that the diffusion coefficients strongly depend on the potential. There is absence of the constant diffusion coefficients in the two-phase coexisting region around 4.7 V for the material with Fd3m space group, where the Li+ ion diffusivity is slower, in consistence with ref. 70. However, the four samples exhibit similar DLi+ values in the range of 10−8 to 10−11 cm2 s−1 during the lithium insertion process irrespective of the particle size, higher than those reported in ref. 69–71, which can be attributed to the hollow hierarchical structure of as-prepared samples. Whereas, the relatively small particle size and more homogeneous distribution of LNMO-0.3 sample could shorten the diffusion distance of Li+ ions, thus leading to its superior rate capability and cycling performance, in consistence with Fig. 4 and 6.


image file: c8ra05817b-f8.tif
Fig. 8 Li+ ion diffusion coefficient for all electrodes during 3rd discharge process.

Conclusions

Hollow hierarchical LiNi0.5Mn1.5O4 cathode material has been successfully synthesized via a urea-assisted hydrothermal method followed by high-temperature calcination process. The effects of reactant concentration on the structure, morphology and electrochemical properties of carbonate precursor and corresponding LiNi0.5Mn1.5O4 product have been intensively investigated. XRD analysis shows that all four samples belong to cubic spinel structure, and the LiNi0.5Mn1.5O4 sample synthesized at 0.3 mol L−1 has the lowest impurity content. FT-IR spectra illustrate that all four samples have mainly the disordered structure with Fd3m space group, and the disordering degree (Mn3+ content) initially decreases and then increases slightly with the reactant concentration increasing. Among them, the LNMO-0.3 sample has the lowest Mn3+ content. SEM observation shows that the average particle size of carbonate precursor exhibits a decreasing trend with the reactant concentration, and the one synthesized at 0.3 mol L−1 has relatively smaller particle size with more homogenous distribution. In addition, the particle morphology changes from shuttle to peanut to sphere with the increase of reactant concentration, in other words, the sphericity of the particle gradually increases with the reactant concentration. It can be concluded that the particle morphology and size of LiNi0.5Mn1.5O4 material can be controlled by changing their precursors' morphology and size, which can be controlled by reactant concentration. Taking the discharge capacity, rate capability and capacity fading factor into account, the LiNi0.5Mn1.5O4 sample synthesized at 0.3 mol L−1 reactant concentration shows optimal overall electrochemical properties, which can be mainly attributed to smaller particle size with homogeneous distribution and lowest impurity content, despite a less disordered structure. In addition, the excellent electrochemical performance of as-prepared LiNi0.5Mn1.5O4 material can be partly attributed to the peculiar hollow hierarchical structure, which can not only provide extra active sites and short Li+ ion diffusion path but also buffer the volume changes and relieve the mechanical strain caused by the repeated insertion/extraction of Li+ ions. On the basis of this work, this facile urea-assisted homogeneous precipitation-hydrothermal preparation of LiNi0.5Mn1.5O4 material with uniform hollow hierarchical architecture will provide an effective and promising strategy for the synthesis of high energy density cathode materials for lithium ion batteries.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors are grateful to the Key R&D Plan Self-raised Project of Hebei Province (Grant number 16214406) and Technology Innovation Foundation Project for Outstanding Youth of Hebei University of Technology (Grant number 2013009) for the financial support of this work.

References

  1. J. M. Tarascon and M. Armand, Nature, 2001, 414, 359 CrossRef PubMed.
  2. M. Armand and J. M. Tarascon, Nature, 2008, 451, 652 CrossRef PubMed.
  3. D. Liu, W. Zhu, J. Trottier, C. Gagnon, F. Barray, A. Guerfi, A. Mauger, H. Groult, C. M. Julien, J. B. Goodenough and K. Zaghib, RSC Adv., 2014, 4, 154 RSC.
  4. K. Zhang, X. Han, Z. Hu, X. Zhang, Z. Tao and J. Chen, Chem. Soc. Rev., 2015, 44, 699 RSC.
  5. J.-H. Kim, N. P. W. Pieczonka and L. Yang, ChemPhysChem, 2014, 15, 1940 CrossRef PubMed.
  6. X. Zhang, F. Cheng, J. Yang and J. Chen, Nano Lett., 2013, 13, 2822 CrossRef PubMed.
  7. M. R. Jo, Y.-I. Kim, Y. Kim, J. S. Chae, K. C. Roh, W.-S. Yoon and Y.-M. Kang, ChemSusChem, 2014, 7, 2248 CrossRef PubMed.
  8. F. Wu, Z. Wang, Y. Su, Y. Guan, Y. Jin, N. Yan, J. Tian, L. Bao and S. Chen, J. Power Sources, 2014, 267, 337 CrossRef.
  9. L. Zhou, K. Zhang, Z. Hu, Z. Tao, L. Mai, Y.-M. Kang, S.-L. Chou and J. Chen, Adv. Energy Mater., 2017, 8, 1701415 CrossRef.
  10. W. W. Wu, H. F. Xiang, G. B. Zhong, W. Su, W. Tang, Y. Zhang, Y. Yu and C. H. Chen, Electrochim. Acta, 2014, 119, 206 CrossRef.
  11. Y. Xue, Z.-B. Wang, L.-L. Zheng, F.-D. Yu, B.-S. Liu, Y. Zhang and Y.-X. Zhou, RSC Adv., 2015, 5, 100730 RSC.
  12. L. Zhou, D. Zhao and X. W. Lou, Adv. Mater., 2012, 24, 745 CrossRef PubMed.
  13. G. Zhang, L. Yu, H. B. Wu, H. E. Hoster and X. W. Lou, Adv. Mater., 2012, 24, 4609 CrossRef PubMed.
  14. J. Li, S. Xiong, Y. Liu, Z. Ju and Y. Qian, Nano Energy, 2013, 2, 1249 CrossRef.
  15. H. Luo, P. Nie, L. Shen, H. Li, H. Deng, Y. Zhu and X. Zhang, ChemElectroChem, 2015, 2, 127 CrossRef.
  16. Z. Wang, L. Zhou and X. W. Lou, Adv. Mater., 2012, 24, 1903 CrossRef PubMed.
  17. X. Liu, D. Li, Q. Mo, X. Guo, X. Yang, G. Chen and S. Zhong, J. Alloys Compd., 2014, 609, 54 CrossRef.
  18. C. Zhu and T. Akiyama, RSC Adv., 2014, 4, 10151 RSC.
  19. Y. Xiao, Y. Zhu, T. Gao, B. Zhong and X. Guo, Ionics, 2017, 23, 27 CrossRef.
  20. W. Luo, J. Alloys Compd., 2015, 636, 24 CrossRef.
  21. D. Lu, L. Yuan, Z. Chen, R. Zeng and Y. Cai, J. Alloys Compd., 2018, 730, 509 CrossRef.
  22. J. Zhang, X. Guo, S. Yao, W. Zhu and X. Qiu, J. Power Sources, 2013, 238, 245 CrossRef.
  23. Y. Liu, M. Zhang, Y. Xia, B. Qiu, Z. Liu and X. Li, J. Power Sources, 2014, 256, 66 CrossRef.
  24. A. Cao and A. Manthiram, Phys. Chem. Chem. Phys., 2012, 14, 6724 RSC.
  25. L. Wang, G. Liu, W. Wu, D. Chen and G. Liang, J. Mater. Chem. A, 2015, 3, 19497 RSC.
  26. J. Q. Zhao and Y. Wang, ECS Trans., 2014, 61, 83 CrossRef.
  27. D. Luo, S. Fang, Q. Tian, L. Qu, S. Shen, L. Yang and S. Hdirano, J. Mater. Chem. A, 2015, 3, 22026 RSC.
  28. W.-H. Ryu, S.-J. Lim, W.-K. Kim and H. S. Kwon, J. Power Sources, 2014, 257, 186 CrossRef.
  29. A. van Bommel and J. R. Dahn, Chem. Mater., 2009, 21, 1500 CrossRef.
  30. J.-H. Kim, A. Huq, M. Chi, N. P. W. Pieczonka, E. Lee, C. A. Bridges, M. M. Tessema, A. Manthiram, K. A. Persson and B. R. Powell, Chem. Mater., 2014, 26, 4377 CrossRef.
  31. J. Li, D. Lu, Z. Chen, H. Zeng, A. Li and Y. Cai, Ionics, 2017, 23, 2275 CrossRef.
  32. Y. Xue, Z. Wang, L. Zheng, F. Yu, B. Lin, Y. Zhang and K. Ke, Sci. Rep., 2015, 5, 13299 CrossRef PubMed.
  33. J. Song, D. W. Shin, Y. Lu, C. D. Amos, A. Manthiram and J. B. Goodenough, Chem. Mater., 2012, 24, 3101 CrossRef.
  34. Z. Bai, N. Fan, Z. Ju, C. Sun and Y. Qian, Mater. Lett., 2012, 76, 124 CrossRef.
  35. M. M. Thackeray, Prog. Solid State Chem., 1997, 25, 1 CrossRef.
  36. S. J. Bao, C. M. Li, H. L. Li and J. H. T. Luong, J. Power Sources, 2007, 164, 885 CrossRef.
  37. J.-H. Kim, S.-T. Myung, C. S. Yoon, S. G. Kang and Y.-K. Sun, Chem. Mater., 2004, 16, 906 CrossRef.
  38. Y. Idemoto, H. Narai and N. Koura, J. Power Sources, 2003, 119–121, 125 CrossRef.
  39. M. Kunduraci, J. F. Al-Sharab and G. G. Amatucci, Chem. Mater., 2006, 18, 3585 CrossRef.
  40. Y. Li, Y. Bai, C. Wu, J. Qian, G. Chen, L. Liu, H. Wang, X. Zhou and F. Wu, J. Mater. Chem. A, 2016, 4, 5942 RSC.
  41. Y. Xue, Z.-B. Wang, L.-L. Zheng, F.-D. Yu, B.-S. Liu and Y.-X. Zhou, ChemistrySelect, 2017, 2, 4325 CrossRef.
  42. P. G. Bruce, B. Scrosati and J.-M. Tarascon, Angew. Chem., Int. Ed., 2008, 47, 2930 CrossRef PubMed.
  43. P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J.-M. Tarascon, Nature, 2000, 407, 496 CrossRef PubMed.
  44. T. Risthaus, J. Wang, A. Friesen, A. Wilken, D. Berghus, M. Winter and J. Li, J. Power Sources, 2015, 293, 137 CrossRef.
  45. A. V. Potapenko, S. I. Chernukhin, I. V. Romanova, K. S. Rabadanov, M. M. Gafurov and S. A. Kirillov, Electrochim. Acta, 2014, 134, 442 CrossRef.
  46. Y.-C. Jin, C.-Y. Lin and J.-G. Duh, Electrochim. Acta, 2012, 69, 45 CrossRef.
  47. X. Zhang, F. Cheng, K. Zhang, Y. Liang, S. Yang, J. Liang and J. Chen, RSC Adv., 2012, 2, 5669 RSC.
  48. Z. Yang, Y. Jiang, J.-H. Kim, Y. Wu, G.-L. Li and Y.-H. Huang, Electrochim. Acta, 2014, 117, 76 CrossRef.
  49. G. B. Zhong, Y. Y. Wang, Y. Q. Yu and C. H. Chen, J. Power Sources, 2012, 205, 385 CrossRef.
  50. O. A. Shlyakhtin, S.-H. Choi, Y. S. Yoon and Y.-J. Oh, J. Power Sources, 2005, 141, 122 CrossRef.
  51. A. Manthiram, K. Chemelewski and E. S. Lee, Energy Environ. Sci., 2014, 7, 1339 RSC.
  52. N. M. Asl, J.-H. Kim, N. P. W. Pieczonka, Z. Liu and Y. Kim, Electrochem. Commun., 2013, 32, 1 CrossRef.
  53. J.-H. Kim, N. P. W. Pieczonka, Z. Li, Y. Wu, S. Harris and B. R. Powell, Electrochim. Acta, 2013, 90, 556 CrossRef.
  54. W. Liu, J. Liu, K. Chen, S. Ji, Y. Wan, Y. Zhou, D. Xue, P. Hodgson and Y. Li, Chem.–Eur. J., 2014, 20, 824 CrossRef PubMed.
  55. C. Zhan, T. Wu, J. Lu and K. Amine, Energy Environ. Sci., 2018, 11, 243 RSC.
  56. O. K. Park, Y. Cho, S. Lee, H.-C. Yoo, H.-K. Song and J. Cho, Energy Environ. Sci., 2011, 4, 1621 RSC.
  57. N. P. W. Pieczonka, Z. Liu, P. Lu, K. L. Olsen, J. Moote, B. R. Powell and J.-H. Kim, J. Phys. Chem. C, 2013, 117, 15947 CrossRef.
  58. D. H. Jang, Y. J. Shin and S. M. Oh, J. Electrochem. Soc., 1996, 143, 2204 CrossRef.
  59. L.-F. Wang, C.-C. Ou, K. A. Striebel and J.-S. Chen, J. Electrochem. Soc., 2003, 150, A905 CrossRef.
  60. L. Wang, D. Chen, J. Wang, G. Liu, W. Wu and G. Liang, RSC Adv., 2015, 5, 99856 RSC.
  61. L. Wang, D. Chen, J. Wang, G. Liu, W. Wu and G. Liang, Powder Technol., 2016, 292, 203 CrossRef.
  62. M.-H. Liu, H.-T. Huang, C.-M. Lin, J.-M. Chen and S.-C. Liao, Electrochim. Acta, 2014, 120, 133 CrossRef.
  63. Y.-L. Ding, X.-B. Zhao, J. Xie, G.-S. Cao, T.-J. Zhu, H.-M. Yu and C.-Y. Sun, J. Mater. Chem., 2013, 1, 7077 RSC.
  64. R. Jin, H. Jiang, Y. Sun, Y. Ma, H. Li and G. Chen, Chem. Eng. J., 2016, 303, 501 CrossRef.
  65. J. Zeng, Y. Cui, D. Qu, Q. Zhang, J. Wu, X. Zhu, Z. Li and X. Zhang, ACS Appl. Mater. Interfaces, 2016, 8, 26082 CrossRef PubMed.
  66. J. Zheng, W. Shi, M. Gu, J. Xiao, P. Zuo, C. Wang and J.-G. Zhang, J. Electrochem. Soc., 2013, 160, A2212 CrossRef.
  67. H. Li, X. Wei, P. Yang, Y. Ren, S. Wang, Y. Xing and S. Zhang, Electrochim. Acta, 2018, 261, 86 CrossRef.
  68. R. Zhao, Z. Yang, J. Liang, D. Lu, C. Liang, X. Guan, A. Gao and H. Chen, J Alloys Compd, 2016, 689, 318 CrossRef.
  69. X. Hao and B. M. Bartlett, J. Electrochem. Soc., 2013, 160, A3162 CrossRef.
  70. A. Ito, D. Li, Y. Lee, K. Kobayakawa and Y. Sato, J. Power Sources, 2008, 185, 1429 CrossRef.
  71. J.-H. Kim, S.-T. Myung, C. S. Yoon, I.-H. Oh and Y.-K. Sun, J. Electrochem. Soc., 2004, 151, A1911 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ra05817b

This journal is © The Royal Society of Chemistry 2018