Investigations on the rate performance of LiFePO4/CeO2 composite materials via polyol technique for rechargeable lithium batteries

M. Sivakumar*, R. Muruganantham and R. Subadevi
#120, Energy Materials Lab, School of Physics, Alagappa University, Karaikudi-630 004, Tamil Nadu, India. E-mail: susiva73@yahoo.co.in

Received 26th June 2015 , Accepted 28th September 2015

First published on 28th September 2015


Abstract

An attempt has been made to synthesize CeO2 modified LiFePO4 composite cathode materials via a polyol technique with a chemical combination route. The surface modified LiFePO4 samples exhibit superior electrochemical performances compared to bare samples. CeO2 may help to induce the fast lithium-ion diffusivity of LiFePO4 cathode materials to promote high-rate stable cycling and a good coulombic efficiency. The complete coverage of a mild coating of CeO2 under an optimized concentration on LiFePO4 may limit the direct contact of the active material with the electrolyte, which improves the interface stability by preventing the dissolution of Fe ions in the electrolyte.


1. Introduction

Olivine-structured LiFePO4 (LFP) has been highlighted as one of the most promising cathode materials for rechargeable lithium-ion batteries (LIBs) because of its environmental compatibility, low cost, high capacity and thermal stability.1,2 However, the inherent low ionic and electronic conductivities of LiFePO4 significantly diminish the electrochemical performance; especially at a high rate. The surface coating technology plays an important role in the electrochemical performance of cathode materials. Surface coating has been divided into three different configurations such as rough coating,3 core shell structure coating4,5 and ultra thin film coating.6 Surface coating has been proven to be effective for improving the capacity retention, rate capability and even thermal stability of cathode materials for lithium ion batteries.7,8 In recent years, the surface coating/modification of LiFePO4 with different materials such as carbon,9,10 metal oxides,11–13 fluorides14 and metal phosphates,15 etc. has been investigated. However, metal or non metal oxides significantly improved the electrochemical performance and most of the oxides are reported as electrochemically inactive oxides or semiconductor oxides, whereas the electronic conductivity is lower.

Cerium(IV) oxide and CeO2-containing materials are intensively studied as catalysts in numerous three-way catalyst formulations. Yao et al.16 and Liu et al.7 reported that the electrochemical properties of pristine LiFePO4 electrodes were improved when the LiFePO4 electrodes were modified with CeO2. They suggested that modification with CeO2 is an effective way to improve the electrochemical properties of the LiFePO4 cathode material. CeO2 modification could enhance the ionic conductivity of LiFePO4 and produce a good electrical contact between the oxides. Also, the resistance between the electrolyte and electrode has been reduced.17–19 Quan et al.20 introduced the effects of CePO4 tailored LiFePO4 in the form of LiFePO4/C/CePO4 composites. CeO2 is a good oxygen storage material based on the reversible redox reaction. However, the LFP modification using CeO2 via a polyol technique has been rarely carried out in the literature. Therefore, an attempt has been made to prepare the cerium metal oxide (CeO2) tailored LiFePO4 via a suitable economic energy-efficient polyol technique. The structural, morphological, chemical and electrochemical performance of the prepared materials were systematically studied in detail.

2. Experimental

2.1. Synthesis of cerium oxide coated LiFePO4

LiFePO4 cathode material was prepared via a conventional polyol technique.11,21,22 The starting materials, iron(II) sulphate heptahydrate (FeSO4·7H2O, 99.9% of Alfa-Aesar) and lithium dihydrogen phosphate (LiH2PO4, 99.9% of Alfa-Aesar) were taken in stoichiometric molar ratio and dissolved in diethylene glycol (DEG, Aldrich), a polyol solvent. The mixed solution was heated close to the boiling point of the polyol solvent (245 °C) for 18 h under a refluxing process. After that, the reacted solution was washed several times with ethanol and acetone. The resulting particles were separated and dried in a vacuum oven at 150 °C for 48 h to obtain the uncoated-LiFePO4 (bare LFP) sample. 1, 2 and 3 wt% of Ce(NO3)3·6H2O and 0.3 N NH4OH were dissolved in distilled water, in which the prepared LiFePO4 was dispersed at room temperature. The mixture was slightly heated and maintained at a temperature of 50 °C while being stirred in order to evaporate the solvent. The resulting particles were heated at 500 °C for 1 h under an argon atmosphere. Finally, CeO2-coated LiFePO4 composite materials were obtained.

2.2. Structure, morphological and elemental characterizations

The powder X-ray diffraction (XRD) analyses of uncoated and coated LiFePO4 samples were performed using a PANalytical X’-pert diffractometer with CuKα radiation operated at 40 kV, 30 mA and the wavelength of λ = 1.54060 Å in the range 2θ = 10–70°. The functional group vibration was analysed using a Thermo Nicolet 380 FT-IR spectrophotometer using KBr pellets in the range 4000–400 cm−1. The morphology of the prepared materials was studied by using field emission scanning electron microscopy (FESEM, LEO-1530, ZEISS, Germany) and energy dispersive spectrometry (EDS, X-MAX50, 30 kV). The nano particles and coating of CeO2 were observed by high resolution transmission electron microscopy (HR-TEM) (Techni G2 S-TWIN, FEI, Netherlands) techniques. The chemical valence states of the elements were investigated by X-ray photoelectron spectroscopy (XPS, PHI model 5802).

2.3. Fabrication of coin cell and electrochemical studies

A Li-metal |LiPF6(EC + DMC)| CeO2 coated LiFePO4 cell was used to investigate the electrochemical performances of the prepared composite cathodes using CR2032 type coin cells. The cathodes (positive electrodes) were prepared by the mixing of 80 wt% CeO2/LiFePO4 composite powder, 10 wt% super P, and 10 wt% poly-(vinylidene fluoride) (PVdF) in N-methylpyrrolidone (NMP) solvent to form a homogeneous slurry. Then, the mixed slurry was spread uniformly on thin aluminum foil and dried in vacuum at 120 °C for 6 h and then roll pressed; then the samples were punched into circular discs. A polypropylene separator (Celgard 2400, Hoechst Celanese Corp) was drenched in the electrolyte for 24 h prior to use. The coin cell assembling procedures were performed using an Ar-filled glove box by keeping both the oxygen and moisture levels less than 1 ppm. The galvanostatic charge–discharge analysis was performed using a BTS-55 Neware battery testing system between the potential 2.5 and 4.5 V (vs. Li/Li+) at ambient temperature with different C-rates. Electrochemical impedance (EIS) analysis was performed on a CHI 660D electrochemical analyser (CH Instruments) at room temperature in the frequency range 106 to 0.01 Hz with AC signal amplitude of 5 mV.

3. Results and discussion

3.1. Structural studies

Fig. 1 shows the XRD patterns of bare and CeO2-coated LiFePO4 cathode materials with the standard data. The pristine and CeO2-coated LiFePO4 materials have been confirmed to possess a well-defined orthorhombic olivine structure with a space group of Pnma (JCPDS 83-2092). The diffraction peaks of CeO2 not appearing in the 1 wt% based sample is due to the low content of CeO2 in the precursor.23 Fig. 1c and d show the existence of two minor peaks appearing at 2θ = 29 and 33° of (1 1 1) and (2 0 0) planes, corresponding to 2 and 3 wt% of the CeO2 coated LiFePO4 sample, respectively. The presence of impurity peaks suggests that the CeO2 has formed a thin solid solution layer on the surface of the LiFePO4 particles. CeO2 has been confirmed by standard crystal diffraction data and literature reports (JCPDS 89-8436).24,25 Hence, the CeO2 coating does not change the structure of the bare LiFePO4. These results indicate that cerium oxide is just coated on the surface of the LiFePO4 particles. The poor crystallinity of CeO2 is due to the fact that the precursor has been calcined at a low temperature for a short duration (500° for 1 h), which results in the disappearance of the face-centered cubic CeO2 peaks in the CeO2 coated samples. A similar phenomenon can be observed in the earlier reports of rare earth oxide modified samples after calcination at a low temperature (400 °C).16,26,27
image file: c5ra12418b-f1.tif
Fig. 1 XRD patterns of (i) (a) bare LiFePO4, (b–d) 1 to 3 wt% of CeO2 coated LiFePO4 and (ii) enlarged patterns of selected 2θ range.

The crystallite size (D) has been calculated using Scherrer’s equation D = /b[thin space (1/6-em)]cos[thin space (1/6-em)]θ, where D is the crystallite size, K is the shape factor (0.9), λ is the X-ray wavelength, b is the full width at half maximum (FWHM) and θ is the Bragg’s angle. The calculated crystal lattice parameters and average crystallite sizes for all samples are shown in Table 1.

Table 1 The lattice parameters and crystallite sizes of uncoated and coated samples from XRD data
Structural parameters Sample name
JCPDS no. 83-2092 LiFePO4 1 wt% coated LiFePO4 2 wt% coated LiFePO4 3 wt% coated LiFePO4
Lattice constant values and volume a = 10.33 Å a = 10.3142 Å a = 10.3141 Å a = 10.3139 Å a = 10.3137 Å
b = 6.010 Å b = 6.0022 Å b = 6.0020 Å b = 6.0019 Å b = 6.0018 Å
c = 4.693 Å c = 4.6842 Å c = 4.6842 Å c = 4.6843 Å c = 4.6843 Å
V = 291.4 Å3 V = 292.08 Å3 V = 291.81 Å3 V = 292.08 Å3 V = 291.81 Å3
Average crystallite sizes (nm) 51 52 53 59


The slight change in the cell parameter values indicates that the presence of CeO2 on the LiFePO4 surface doesn’t affect the structure of bare LiFePO4. It indicates that CeO2 did not diffuse into the LiFePO4 lattice; it is merely a coating on the surface of LiFePO4 and could form a solid solution. Also, the average crystallite sizes are slightly increased with the increase of the coating content and do not change the phase structure of LiFePO4, similar to other coating materials.26,27 However, due to the fact that the coating procedure was carried out at 500 °C, a few amorphous particles may have been attached on the surface of the core material.

Fig. 2a–d shows the FT-IR spectra of the bare and CeO2 coated LiFePO4 samples. The band between 932 and 1134 cm−1 corresponds to the symmetric and antisymmetric stretching mode of the P–O vibration peak of the PO4 tetrahedron in defect-free LiFePO4. Symmetric and antisymmetric O–P–O bending modes exist in the range of 460–633 cm−1.28 The bending and stretching modes of the bare and coated samples have been observed in the range 460–1134 cm−1. In addition, no additional peaks are observed in the bare LiFePO4 material.


image file: c5ra12418b-f2.tif
Fig. 2 (a–d) FT-IR spectra of bare, 1, 2 and 3 wt% CeO2 coated LiFePO4 samples.

The presence of CeO2 in the coated LiFePO4 is confirmed through the band at 1381 cm−1, which is the characteristic vibration mode of the CeO2 stretching vibration.29,30

The intensity of the vibrational peaks increases with the increase of CeO2 content. The O–H bending and stretching vibrations are located at 1628 and 3436 cm−1. This may be due to the lower calcination temperature or absorbed water molecules from air during the analysis, since ceria is slightly hygroscopic in nature.

3.2. Morphological and elemental analysis

The SEM images of the bare, 1, 2 and 3 wt% of CeO2-coated LiFePO4 materials are shown in Fig. 3a–d. The bare LiFePO4 material has long rod-like particles with the average size 350 nm × 100 nm of length × width. The calcination results in the breaking of the lengthy rods into smaller rods and a few grains on its surface. It exhibits agglomeration between the grains, which are found less in the coated samples than the bare sample. The average particle sizes were measured as length × width (250 × 80), (128 × 56) and (180 × 65) nm respectively for the 1, 2 and 3 wt% of CeO2 coated LiFePO4 samples using the ‘measureIT’ software (Olympus soft imaging solution GMBH product). As the content of CeO2 is increased, the solid solution or nano-sized CeO2 particle growth is increased on the LFP surface. The surface modification on LiFePO4 would affect the electrochemical properties.
image file: c5ra12418b-f3.tif
Fig. 3 (a–d) SEM images of bare, 1, 2 and 3 wt% CeO2 coated LiFePO4 samples.

The energy dispersive X-ray (EDX) analysis was carried out to identify the elements present. Fig. 4a–d illustrates the EDX spectra of the bare, 1, 2 and 3 wt% CeO2 coated LFP samples. According to the EDX results, a new phase containing Ce was detected in the coated samples. These results are in good agreement with the actual CeO2 content used in the coated LFP materials. Also, from this analysis, we can calculate that the ratio of the elemental atoms such as Fe/P is almost equal to 1 and O/Fe or O/P is equal to 4, which gives an additional confirmation of LiFePO4 formation.


image file: c5ra12418b-f4.tif
Fig. 4 (a–d) EDX spectra of bare, 1, 2 and 3 wt% CeO2 coated LiFePO4 samples.

Fig. 5a–c clearly shows the TEM images of 1, 2 and 3 wt% of CeO2 coated LFP particles. It is apparent from the TEM images (Fig. 5a–c) that there are two distinct morphologies; the dark part representing LiFePO4 and the lighter part representing CeO2 particles or the solid solution layer. Fig. 5b demonstrates that 2 wt% of CeO2 has been coated as a thin layer on the surface of LiFePO4 particles. Fig. 4c reveals the existence of a few particles and layers on the surface of the core particles. The related SAED pattern (Fig. 5d–f) exhibits a regular and clear diffraction spot array, which indicates that the particle is single-crystalline and it can be indexed to the orthorhombic phase of LiFePO4. These results are in good agreement with the XRD results.


image file: c5ra12418b-f5.tif
Fig. 5 (a–c) TEM images and (d–f) SAED patterns of 1, 2 and 3 wt% of CeO2 coated LiFePO4 samples and (g and h) higher magnification TEM images of 2 wt% CeO2 coated LFP sample.

The surface composition and the oxidation state of the elements were analysed by X-ray photoelectron spectroscopy (XPS). Fig. 6a shows the wide range core spectra of 2 wt% CeO2 coated LiFePO4 composite and the results confirm the presence of all the elements in the prepared sample. Fig. 6b–f demonstrates the core spectra of Li 1s, Fe 2p, P 2p, O 1s and Ce 3d respectively. Fig. 6c (Fe 2p) has been split into two components, because of spin–orbit coupling, namely Fe 2p3/2 and Fe 2p1/2. LiFePO4 shows Fe 2p3/2 main peaks at 710 eV and 724 eV for Fe 2p1/2, which are in good agreement with Fe2+ in LiFePO4.31


image file: c5ra12418b-f6.tif
Fig. 6 X-ray photoelectron spectra of 2 wt% CeO2 coated LiFePO4 composite: (a) wide range, (b) Li 1s, (c) Fe, (d) P, (e) O, and (f) cerium.

The presence of the carbon (C 1s) peak denotes the minor contribution of oxygenated carbon at the surface of the prepared material. The O 1s binding energy (BE) that has been observed at 530.46 eV corresponds to the lattice oxygen (O2−) of the orthorhombic structure.32,33 The XPS spectrum of Ce is complex and split into Ce 3d3/2 and Ce 3d5/2 with multiple shake-up and shake-down satellites. The peaks between 875 and 895 eV belong to the Ce 3d5/2, while peaks between 895 and 910 eV correspond to the Ce 3d3/2 levels.25,34 The peak at 916 eV is a characteristic satellite peak indicating the presence of CeO2(IV).35 Thus, the XPS analysis confirms the presence of elements in the as-prepared material.

3.3. Electrochemical studies

The electronic conductivity of the cathode material is very important for lithium intercalation/de-intercalation processes in the lithium ion cell. The electrical conductivities of the pristine and CeO2 coated LFP samples are shown in Table 2. The electrical conductivity of the synthesized materials was determined by using a four-probe DC method. It is shown that the electrical conductivities of the coated samples are increased when compared to pristine LFP.36 Electrical conductivity is a physical property reflecting the ability of a material to transfer electrical charge. The ionic mobility depends on many factors, but mostly on the size of ions and ionic bond strength. Therefore, an enhancement of conductivity is used to induce the EC performance of the LFP electrode material. The electronic conductivity of the active material has played a vital role in the transfer of charge from the current collector during the charging and discharging process. The actual lithium ionic conduction takes place between the electrodes through the electrolyte in the Li/electrolyte/CeO2 coated LiFePO4 cell couple. Normally, for bare LiFePO4, the electronic conductivity is very low. Therefore, the charge transfer from the electrode material to the current collector was not up to the mark. But, after coating the CeO2 on LiFePO4 to an optimum level, the transfer of charge is excellent and exhibits appreciable improvement in performance over the bare material.37
Table 2 Electronic conductivities of the pristine and CeO2 coated LiFePO4 samples at room temperature
Sample name Electronic conductivity
Bare LiFePO4 2.88 × 10−2 S cm−1
1 wt% CeO2–LiFePO4 5.01 × 10−2 S cm−1
2 wt% CeO2–LiFePO4 7.03 × 10−2 S cm−1
3 wt% CeO2–LiFePO4 6.25 × 10−2 S cm−1


Fig. 7a shows the initial charge/discharge curves of the pristine, 1, 2 and 3 wt% CeO2-coated LiFePO4 samples at 0.1C rate at room temperature. All the cells exhibited discharge voltage plateaus around 3.45 V, which is the main characteristic of the two-phase reaction of the lithium extraction and insertion between LiFePO4 and FePO4.29 The initial charge/discharge capacity of pristine LiFePO4 is 152/149 mA h g−1, whereas for 1, 2 and 3 wt% CeO2-coated LiFePO4, it is 157/153, 164/163 and 160/159 mA h g−1 at 0.1C rate, respectively. This value of coated LFP is higher when compared to the bare sample, which is due to the improvement of the electronic conductivity after the surface treatment. Obviously, the metal oxide particle coated cathode surface may limit the direct contact of the active material with the electrolyte, i.e., the metal oxide particles cover the core material. This improves the interface stability and prevents the dissolution of Fe-ions in the electrolyte.38


image file: c5ra12418b-f7.tif
Fig. 7 Charge–discharge performance of (a) bare and coated LiFePO4 at 0.1C rate, (b) CeO2 coated LiFePO4 at 1C rate, (c) rate performance of CeO2 coated LiFePO4 at various rates, and (d–f) cyclic behaviour and coulombic efficiency of 1, 2 and 3 wt% of CeO2 coated LiFePO4 at 1C rate at room temperature.

Fig. 7b demonstrates the charge/discharge profile at 1C rate at room temperature for all coated LFP samples up to 50 cycles. Among the samples studied, the 2 wt% of CeO2 coated on LiFePO4 shows a better discharge capacity with cyclic stability and retention when compared with 1 and 3 wt% of CeO2 coated LiFePO4 composites. There are several factors affecting the electrochemical performances of the prepared materials: (i) the capacity increases upon increasing CeO2 content until 2 wt%, which is due to the improvement of the electronic conductivity of the material. The coverage of the solid solution layer on its surface, which provides a better electrode–electrolyte contact and reduces the polarization of bare material (inset Fig. 7a) causes an enhancement in the electronic conductivity. (ii) The 2 wt% of CeO2 coated LFP sample exhibits better capacity, retention and cyclic capability compared to the other samples studied. This is due to the complete coverage of a thin layer of CeO2 solid solution, which may enhance the structural stability and improve the ionic/electronic conductivity of the bare LFP material.24,26 (iii) Further increasing of the CeO2 content causes the aggregation of the cerium oxide particles in a thick layer on the surface of the LiFePO4, which leads to crystallite regions, and the CeO2 particles tend to impede ionic movement by acting as mere insulators. An optimal content of CeO2 addition would result in enhancing the electrochemical properties;24 whereas beyond the optimal content of CeO2, the electrochemical performance decreases since its higher content leads to an inactive or insulating nature. Similar observations have already been made by Ha et al.39 and Liu et al.7

Recently, most of the high rate applications like EV (electric vehicles), HEV (heavy electric vehicles), HED (high energy density) batteries, etc. are occupying the current commercial market due to their appealing rate performance and cycling capability.38 In the present study the high rate studies for the cells have been performed and are given in Fig. 7c. This represents the rate performances of 1, 2 and 3 wt% of CeO2 coated LiFePO4 between 0.1 and 30C rates at room temperature. The electrode discharge capacities are 151, 147, 126, 112, 65, 58 mA h g−1 for 1 wt% of CeO2, 163, 160, 152, 147, 116, 75 mA h g−1 for 2 wt% of CeO2 and 157, 156, 141, 134, 111, 68 mA h g−1 for 3 wt% of CeO2 coated samples respectively, up to 10 cycles at 0.1, 0.5, 1, 10, 20 and 30C rates. The aforementioned result of 2 wt% CeO2 coated LFP electrodes exhibits remarkable improvement in discharge capacity compared to earlier reports viz., carbon, CeO2 and other metal oxide coated LFP (Table 3).

Table 3 The CeO2 coated LiFePO4 samples compared with earlier reports of carbon, CNT, graphene coated and metal oxide modified LiFePO4 samples
Material name Observed discharge capacity (mA h g−1) with current rate of 0.1C Reference
LiFePO4 and LiFePO4–CNT 129 and 155 at 10 mA g−1 Wu et al.40
LiFePO4/C and graphene/LiFePO4/C 146.5 and 157.8 Wang et al.41
C–LiFePO4 and graphene–LiFePO4 ∼150 and ∼153 Kim et al.42
LiFePO4/graphene 160.3 Wang et al.43
LiFePO4/rGO 161 Nagaraju et al.44
Sn-modified LiFePO4 (Sn/Fe ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]99, 3[thin space (1/6-em)]:[thin space (1/6-em)]97 and 5[thin space (1/6-em)]:[thin space (1/6-em)]95) 122, 122 and 114 Ziolkowska et al.13
SiO2-coated LiFePO4 160 Li et al.45
LiFePO4/C/CePO4 156 Quan et al.20
2 wt% of CeO2-coated LiFePO4/C 153.8 Yao et al.16
Pristine LiFePO4 149 In this work
1 wt% of CeO2–LFP 153
2 wt% of CeO2–LFP 163
3 wt% of CeO2–LFP 159


Fig. 7d–f illustrate the 1, 2 and 3 wt% of CeO2 coated LFP samples for 100 cycles of discharge capacity and coulombic efficiency at 1C rate at room temperature. The coulombic efficiency of CeO2 coated LFP composites are almost the same (about 97–99%) at 1C rate for the coated samples. The 2 wt% of CeO2 coated LFP sample has shown an excellent rate with an appealing cyclic performance and more stable coulombic efficiency than the rest. It is observed that the coulombic efficiency has been improved upon increasing the CeO2 content until 2 wt%; further addition of CeO2 affected this trend. This could be due to the fact that Ce4+ is not electrochemically active and, therefore, the presence of excess CeO2 could lower the discharge capacity.35 Consequently, the optimized amount of CeO2 coating could improve the discharge capability and coulombic efficiency of the electrode materials.

The electrochemical kinetics of the pristine and CeO2 coated LiFePO4 composite has been studied by Electrochemical Impedance Spectroscopy (EIS). Fig. 8 shows the Nyquist impedance plots of the materials and the equivalent circuit (inset of Fig. 8). The Nyquist plots are composed of a semi-circle in the high frequency region and an inclined line in the low frequency region. The intercept of the high frequency region corresponds to the ohmic resistance (Re), which represents the resistance of the electrolyte and electrode. The semicircle in the middle frequency range indicated the charge transfer resistance (Rct). The inclined line corresponds to the diffusion of Li+ in the bulk electrode, namely the Warburg impedance (Zw). The charge-transfer resistance of CeO2 coated LiFePO4 electrodes is much less than that of pristine LiFePO4. This indicates that the coating is more favourable for the insertion and de-insertion of lithium ions during the charge and discharge process. Among the CeO2 coated LiFePO4 samples, the 2 wt% of CeO2 coated LiFePO4 shows the smallest Rct of 230 Ω and Zw, which is clearly seen in the plots (Fig. 8).


image file: c5ra12418b-f8.tif
Fig. 8 Nyquist plots of the bare and CeO2 coated LiFePO4 composite electrodes.

The small resistance indicates that Li+ ion and electron transfer are more feasible on the electrode, which may be attributed to the decreasing electronic resistance of the composite material. The impedance plot of the bare and CeO2 coated LFP electrodes are represented by a large depressed semicircle with an inductive line and a very small semicircle at room temperature, indicating a lower electronic/ionic conductivity of bare LFP than the values of CeO2 coated LFP sample.

The impedance of the 2 wt% of the CeO2 coated LiFePO4 composite is significantly smaller than that of the other samples, which indicates an enhancement in both the ionic and electronic conductivity as well as lower interface impedance, relating to a superior rate performance.37 CeO2 can serve as protective layer and prevents the direct contact between LiFePO4 and the electrolyte solution like a preformed SEI, which can reduce side reactions and improve the structure, cycle stability and decrease charge-transfer resistance.24 The CeO2 coating enhances the reversibility of the electrode reaction. However, when the amount of CeO2 is increased to 3 wt%, the initial discharge capacity and cycling stability are decreased, which may be due to an excessive amount of CeO2. The 2 wt% of CeO2 coated on LiFePO4 is found to be an optimum level of coating content which improves EC performance.46

4. Conclusions

The olivine type orthorhombic structure of bare LiFePO4 has been successfully prepared via the polyol technique with a low boiling point solvent (DEG), a simple binary precursor and also without post-heat treatment, any special environment or carbon materials. In addition, the CeO2 coated LFP has been subjected to a one step short-time heat treatment without any change in the bare LFP structure. Due to the heat treatment, the particle size of the coated samples has been reduced. The 2 wt% of CeO2 coated LFP surface was formed in a thin layer. It exhibits a superior electrochemical performance than other coated and bare LFP samples. It denotes that the appropriate content of CeO2 coating has better ionic and electronic conduction, which enhances the ionic and electronic transport in the LiFePO4 electrode. However, the higher content of the inactive materials may be impeding the mass and charge transfer and may reduce the reversible capacity. Therefore, an optimized content of CeO2 protects the LiFePO4 electrode from electrolyte corrosion and maintains the structural stability of the LiFePO4. It is beneficial to use in high-rate battery applications. This is due to the excellent rate capability and cycle stability of CeO2 tailored LiFePO4.

Acknowledgements

The authors M. Sivakumar and R. Muruganantham gratefully acknowledge the financial support to carry out this work from the Department of Science and Technology (DST), New Delhi, Govt. of India under the DST-SERC major research project (SR/S2/CMP-0049/2008).

Notes and references

  1. A. K. Padhi, K. S. Nanjundaswamy and J. B. Goodenough, J. Electrochem. Soc., 1997, 144, 1188–1194 CrossRef CAS PubMed.
  2. J. Wang and X. Sun, Energy Environ. Sci., 2015, 8, 1110–1138 CAS.
  3. J.-T. Son, J. Korean Electrochem. Soc., 2010, 13, 246–250 CrossRef CAS.
  4. F. Pan and W.-l. Wang, J. Solid State Electrochem., 2012, 16, 1423–1427 CrossRef CAS.
  5. C.-H. Hsu, H.-Y. Liao and P.-L. Kuo, Energy Technol., 2014, 2, 409–413 CrossRef CAS PubMed.
  6. X. Rui, X. Zhao, Z. Lu, H. Tan, D. Sim, H. H. Hng, R. Yazami, T. M. Lim and Q. Yan, ACS Nano, 2013, 7, 5637–5646 CrossRef CAS PubMed.
  7. Y. Liu, C. Mi, C. Yuan and X. Zhang, J. Electroanal. Chem., 2009, 628, 73–80 CrossRef CAS PubMed.
  8. J. Wang, X. He, R. Kloepsch, S. Wang, B. Hoffmann, S. Jeong, Y. Yang and J. Li, Energy Technol., 2014, 2, 188–193 CrossRef CAS PubMed.
  9. X. Zhi, G. Liang, L. Wang, X. Ou, L. Gao and X. Jie, J. Alloys Compd., 2010, 503, 370–374 CrossRef CAS PubMed.
  10. B. Ding, G. Ji, Z. Sha, J. Wu, L. Lu and J. Y. Lee, Energy Technol., 2015, 3, 63–69 CrossRef CAS PubMed.
  11. M. Sivakumar, R. Muruganantham and R. Subadevi, Appl. Surf. Sci., 2015, 337, 234–240 CrossRef PubMed.
  12. S. Liu and H. Wang, Mater. Lett., 2014, 122, 151–154 CrossRef CAS PubMed.
  13. D. Ziolkowska, K. P. Korona, B. Hamankiewicz, S.-H. Wu, M.-S. Chen, J. B. Jasinski, M. Kaminska and A. Czerwinski, Electrochim. Acta, 2013, 108, 532–539 CrossRef CAS PubMed.
  14. M. Leblanc, V. Maisonneuve and A. Tressaud, Chem. Rev., 2015, 115, 1191–1254 CrossRef CAS PubMed.
  15. G.-M. Song, Y. Wu, Q. Xu and G. Liu, J. Power Sources, 2010, 195, 3913–3917 CrossRef CAS PubMed.
  16. J. Yao, F. Wu, X. Qiu, N. Li and Y. Su, Electrochim. Acta, 2011, 56, 5587–5592 CrossRef CAS PubMed.
  17. T.-F. Yi, J.-Z. Wu, M. Li, Y.-R. Zhu, Y. Xie and R.-S. Zhu, RSC Adv., 2015, 5, 37367–37376 RSC.
  18. Q. Li and V. Thangadurai, Fuel Cells, 2009, 09, 684–698 CrossRef CAS PubMed.
  19. Y. Wang, C. X. Guo, J. Liu, T. Chen, H. Yang and C. M. Li, Dalton Trans., 2011, 40, 6388–6391 RSC.
  20. W. Quan, Z. Tang, J. Zhang and Z. Zhang, Mater. Chem. Phys., 2014, 147, 333–338 CrossRef CAS PubMed.
  21. D.-H. Kim and J. K. Kim, Electrochem. Solid-State Lett., 2006, 9, A439 CrossRef CAS PubMed.
  22. S. W. Oh, Z.-D. Huang, B. Zhang, Y. Yu, Y.-B. He and J.-K. Kim, J. Mater. Chem., 2012, 22, 17215 RSC.
  23. F. Wu, Electrochim. Acta, 2009, 54, 6803–6807 CrossRef CAS PubMed.
  24. Y. Yang, R. Guo, G. Cai, C. Zhang, L. Liu, S. Wang, C. Wu and Y. Wan, J. Electrochem. Soc., 2014, 161, A2153–A2159 CrossRef CAS PubMed.
  25. H.-W. Ha, N. J. Yun, M. H. Kim, M. H. Woo and K. Kim, Electrochim. Acta, 2006, 51, 3297–3302 CrossRef CAS PubMed.
  26. W. Yuan, H. Z. Zhang, Q. Liu, G. R. Li and X. P. Gao, Electrochim. Acta, 2014, 135, 199–207 CrossRef CAS PubMed.
  27. S. J. Shi, J. P. Tu, Y. J. Zhang, Y. D. Zhang, X. Y. Zhao, X. L. Wang and C. D. Gu, Electrochim. Acta, 2013, 108, 441–448 CrossRef CAS PubMed.
  28. C. M. Burba and R. Frech, J. Electrochem. Soc., 2004, 151, A1032–A1038 CrossRef CAS PubMed.
  29. Y. Hou, X. Wang, Y. Zhu, C. Hu, Z. Chang, Y. Wu and R. Holze, J. Mater. Chem. A, 2013, 1, 14713–14718 CAS.
  30. B. Yan and H. Zhu, J. Nanopart. Res., 2008, 10, 1279–1285 CrossRef CAS.
  31. W. Xiong, Q. Hu and S. Liu, Anal. Methods, 2014, 6, 5708–5711 RSC.
  32. A. Fedorková, R. Orináková, A. Orinák, M. Kupková, H.-D. Wiemhöfer, J. N. Audinot and J. Guillot, Solid State Sci., 2012, 14, 1238–1243 CrossRef PubMed.
  33. J. Ni and Y. Wang, RSC Adv., 2015, 5, 30537–30541 RSC.
  34. S. A. Miller, V. Y. Young and C. R. Martin, J. Am. Chem. Soc., 2001, 123, 12335 CrossRef CAS PubMed.
  35. D. Arumugam and G. P. Kalaignan, Electrochim. Acta, 2010, 55, 8709–8716 CrossRef CAS PubMed.
  36. S. Gopukumar, C. Nithya, P. H. Maheshwari, R. Ravikumar, R. Thirunakaran, A. Sivashanmugam, S. K. Dhawanb and R. B. Mathur, RSC Adv., 2012, 2, 11574–11577 RSC.
  37. C. Wang and J. Hong, Electrochem. Solid-State Lett., 2007, 10, A65–A69 CrossRef CAS PubMed.
  38. W. Wei, L. Guo, X. Qiu, P. Qu, M. Xu and L. Guo, RSC Adv., 2015, 5, 37830–37836 RSC.
  39. H.-W. Ha, N. J. Yun and K. Kim, Electrochim. Acta, 2007, 52, 3236 CrossRef CAS PubMed.
  40. G. Wu, Y. Zhou, X. Gao and Z. Shao, Solid State Sci., 2013, 24, 15–20 CrossRef CAS PubMed.
  41. Z. Wang, H. Guo and P. Yan, Ceram. Int., 2014, 40, 15801–15806 CrossRef CAS PubMed.
  42. W. K. Kim, W. H. Ryu, D. W. Han, S. J. Lim, J. Y. Eom and H. S. Kwon, ACS Appl. Mater. Interfaces, 2014, 6, 4731–4736 CAS.
  43. L. Wang, H. Wang, Z. Liu, C. Xiao, S. Dong, P. Han, Z. Zhang, X. Zhang, C. Bi and G. Cui, Solid State Ionics, 2010, 181, 1685–1689 CrossRef CAS PubMed.
  44. D. H. Nagaraju, M. Kuezma and G. S. Suresh, J. Mater. Sci., 2015, 50, 4244–4249 CrossRef CAS.
  45. Y.-D. Li, S.-X. Zhao, C.-W. Nan and B.-H. Li, J. Alloys Compd., 2011, 509, 957–960 CrossRef CAS PubMed.
  46. J. Zhai, M. Zhao, D. Wang and Y. Qiao, J. Alloys Compd., 2010, 502, 401–406 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.