Next Article in Journal
Lithium Ion Cell/Batteries Electromagnetic Field Reduction in Phones for Hearing Aid Compliance
Next Article in Special Issue
Durability and Reliability of Electric Vehicle Batteries under Electric Utility Grid Operations. Part 1: Cell-to-Cell Variations and Preliminary Testing
Previous Article in Journal / Special Issue
Calculation of Constant Power Lithium Battery Discharge Curves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Carbon Additive Effect on Electrochemical Performance of LiFe0.5Mn0.5PO4/C Composites by a Simple Solid-State Method for Lithium Ion Batteries

by
Chun-Chen Yang
1,2,*,
Yen-Wei Hung
1 and
Shingjiang Jessie Lue
3
1
Department of Chemical Engineering, Ming Chi University of Technology, New Taipei City 243, Taiwan
2
Battery Research Center of Green Energy, Ming Chi University of Technology, New Taipei City 243, Taiwan
3
Department of Chemical and Materials Engineering and Green Technology Research Center, Chang Gung University, Kwei-shan, Taoyuan 333, Taiwan
*
Author to whom correspondence should be addressed.
Batteries 2016, 2(2), 18; https://doi.org/10.3390/batteries2020018
Submission received: 28 April 2016 / Revised: 25 May 2016 / Accepted: 7 June 2016 / Published: 15 June 2016
(This article belongs to the Special Issue Lithium Ion Batteries)

Abstract

:
This work reported a solid-state method to prepare LiFe0.5Mn0.5PO4/C (LFMP/C) composite cathode materials by using LiH2PO4, MnO2, Fe2O3, citric acid (C6H8O7), and sucrose (C12H22O11). The citric acid was used as a complex agent and C12H22O11 was used as a carbon source. Two novel hollow carbon sphere (HCS) and nanoporous graphene (NP-GNS) additives were added into the LFMP/C composite to enhance electrochemical performance. The HCS and NP-GNS were prepared via a simple hydrothermal process. The characteristic properties of the composite cathode materials were examined by micro-Raman spectroscopy, X-ray diffraction (XRD), scanning electron microscopy (SEM), elemental analysis (EA), and alternating current (AC) impedance methods. The coin cell was used to investigate the electrochemical performance at various rates. It was found that the specific discharge capacities of LFMP/C + 2% NP-GNS + 2% HCS composite cathode materials were 161.18, 154.71, 148.82, and 120.00 mAh·g−1 at 0.1C, 0.2C, 1C, and 10C rates, respectively. Moreover, they all showed the coulombic efficiency ca. 97%–98%. The advantage of the one-pot solid-state method can be easily scaled up for mass-production, as compared with the sol-gel method or hydrothermal method. Apparently, the LFMP/C composite with HCS and NP-GNS conductors can be a good candidate for high-power Li-ion battery applications.

Graphical Abstract

1. Introduction

Li-ion batteries are appropriate for use in cell phones, laptop computers, digital cameras, renewable energy storage, and smart grid applications because of their relatively high energy density, low cost, and high rate capabilities. Li-ion batteries used in hybrid-electric vehicle or electric vehicle applications must be charged and discharged rapidly; therefore, the electrodes must perform at a high rate and maintain high discharge capacity and reliable cycle-life stability. Compared with the LiFePO4 cathode material [1], the LiMnPO4 cathode material exhibits a higher working potential at 4.1 V (versus Li/Li+) and is compatible with conventional liquid carbonate-based electrolytes. The energy density of the LiMnPO4 (697 Wh·kg−1) [2] is higher than that of the LiFePO4 (586 Wh·kg−1); however, the electrical conductivity of the LiMnPO4 material (less than 10−10 S·cm−1) is much lower than that of the LiFePO4 material (at 1.8 × 10−8 S·cm−1). The electrochemical performance can be improved by coating the samples’ surface with carbon, or by doping with Fe atoms or nano-sized cathode materials. Olive phosphate materials combined with mixed transition-metal ions, such as LiFe1−xMnxPO4/C, have recently attracted considerable attention [3,4,5,6,7,8,9,10,11,12,13,14]. Zou et al. [15] prepared LiFe0.2Mn0.8PO4/C cathode materials by a solid-state method and sucrose used as the carbon source. The as-prepared cathode materials achieved capacities of 150 mAh g−1 and 110 mAh g−1 at 1/20C and 1C, respectively. Liu et al. [16] recently synthesized a LiMn1−xFexPO4/C (x = 0.1, 0.2, 0.3) cathode material by a ball-milling process with MnPO4·H2O precursors, LiOH, NH4H2PO4, and 30 wt% glucose. The as-prepared LiMn0.7Fe0.3PO4/C cathode material had a capacity of 140 mAh·g−1 and high rate performance. Mi et al. [17] prepared mesoporous LiFe0.6Mn0.4PO4/C materials—comprising of different amounts of muti-wall carbon nanotubes (MWCNTs) by using a two-step carbon coating and spray drying process. The LiFe0.6Mn0.4PO4/C materials with 2 wt% MWCNTs showed excellent performance by delivering a discharge capacity of 163.3 mAh·g−1 at 0.1C and yielding an ultra-high rate capacity of 64.23 mAh·g−1 at 50C. Hu et al. [18] prepared Fe-doped LiMn1−xFexPO4 (x = 0.5) nanomaterials by using a solvothermal method. LiMn0.5Fe0.5PO4 notably delivered a 100% capacity retention with a discharge capacity of 147 mAh·g−1 at 1C rate. Zhong et al. [19] most recently synthesized a carbon-coated LiFe0.5Mn0.5PO4/C (LFMP/C) material by a rheological phase reaction method with stearic acid as the carbon source (i.e., a solid-state method). The LFMP/C material exhibited the highest electrochemical performance compared with the LiMn0.2Fe0.8PO4/C and LiMn0.8Fe0.2PO4/C materials. The LFMP/C material delivered discharge capacities of 138, 99, 80, 72, 67, and 55 mAh·g−1 at 0.1C, 1C, 5C, 10C, 15C, and 20C rates, respectively. The LFMP/C material achieved long and stable cycling performance with a capacity of 103 mAh·g−1 at 1C rate during a 100-cycle test.
In this work, nanostructured carbon materials show a variety of fascinating and admirable properties for Li-ion batteries, such as high surface area, low diffusion distance, and high electron and ionic conductivity [20,21]. Therefore, nano-sized carbon materials show extremely promising to improve the discharge capacity, the power capability, and the long cycling stability. We studied the carbon conductor effect on performances of LFMP/C composite. Both hollow carbon sphere (HCS) and nanoporous graphene (NP-GNS) were synthesized by a hydrothermal process. The characteristic properties of the LFMP/C composite cathode materials were examined by micro-Raman spectroscopy, X-ray diffraction (XRD), scanning electron microscopy (SEM), elemental analysis (EA), and alternating current (AC) impedance method.

2. Experimental

2.1. Preparation of Hollow Carbon Sphere and Nanoporous Graphene Carbon Conductors

The HCS was also synthesized by a simple, one-pot, green, hydrothermal process. In a typical procedure [22], 40 g glucose (Aldrich, St. Louis, MO, USA) and 4 g sodium dodecyl sulfate (SDS, Aldrich) were first dissolved into in 460 mL distilled water. Then, the resultant solution was stirred for three days at 50 °C. The obtained mixture solution was transferred into a Teflon-lined autoclave of 600 mL capacity with a stainless steel shell, followed by a hydrothermal treatment at 190 °C for 10 h. After that, the autoclave was allowed to reach room temperature. The as-prepared carbon sphere materials was filtered and washed with distilled water and ethanol several times, and then dried at 80 °C for overnight. The HCS materials were obtained by sintering the as-prepared powders at 900 °C for 4 h in N2 atmosphere.
The nanoporous graphene oxide (GO) was fabricated by a solvothermal process [23]. Both the as-prepared GO and ferrocene were used as the precursors. The concentration of GO solution was around 4 mg·mL−1. The ratio of GO to ferrocene was kept at 1:20. The solvothermal process was carried out at 180 °C for 10 h. The mixture solid product was separated through filtration, washed with ethanol for several times to remove the residual ferrocene, and then dried. The ferrocene/GO mixed composite was annealed at 800 °C for 4 h in a 95% Ar/5% H2 atmosphere; at the same time, the ferrocene was decomposed into iron/Austenite nanosphere. The iron/Austenite nanosphere was further removed with 12 M HCl solution, then nano-porous graphene sheet was obtained through filtration and annealed in Ar at 800 °C for 1 h. The NP-GNS was generated through a simple thermal reduction process.

2.2. Preparation of LiFe0.5Mn0.5PO4/C Materials

The LFMP/C cathode materials were prepared using a wet-balled method. Appropriate amounts of LiH2PO4, Fe2O3, MnO2, carbon black (BP2000), sucrose, and citric acid (Aldrich, 99%) as starting materials were dissolved in acetone. The molar ratio of Li:Fe:Mn:P was 1:0.5:0.5:1. Approximately 10 wt% sucrose and 5 wt% citric acid were used as the carbon source and complex agent, respectively. The sucrose and citric acid were mixed with the acetone and starting raw materials and then wet ball-milled (Planetary Ball-Milling Apparatus, ball-to-powder weight ratio = 10:1, Primium Line, Fritsch, Idar-Oberstein, Germany) at 400 rpm for approximately 10–15 h. The ball-milling process was performed in an inert atmosphere that was controlled to avoid the Fe2+ and Mn2+ species from oxidation. The ball-milled mixture was dried and subsequently heated at 650, 675, and 700 °C for 15 h in an argon atmosphere. The carbon content was approximately 5%–9%. The LFMP/C + 2% graphene (denoted as NP-GNS) +2% carbon sphere (denoted as CS) composites were prepared by adding nano-sized graphene (16 nm thickness, 50 m2·g−1) and carbon sphere (prepared by a hydrothermal process). The obtained LFMP/C composite precursor was further calcined at 650, 675, and 700 °C for 15 h in an argon atmosphere. The residual carbon content was approximately 5%–9%.

2.3. Material Characterization

The crystal structures of the as-prepared LFMP/C samples were examined using an XRD spectrometer (D2 Phaser, Bruker, Karlsruhe, Germany). The surface morphology and the residual carbon morphology were examined using SEM (Hitachi, Toyo, Chiba, Japan) and high-resolution transmission electron microscopy (HR-TEM, JEOL 2010F, Toyo), respectively. The micro-Raman spectra were recorded using a confocal micro-Renishaw fitted with a 632 nm He–Ne excitation laser (Renishaw inVia, London, UK). The residual carbon content in the sample was examined using an elemental analyzer (FLASH 2000, Thermo Fisher Scientific, Waltham, MA, USA).

2.4 Electrochemical Performance of LiFe0.5Mn0.5PO4/C

The electrochemical performances of the LFMP/Li and LFMP/C + 2% NP-GNS + 2% HCS half-cells were measured using a two-electrode system (i.e., CR-2032 coin cells assembled in an Ar-filled glove box). The LFMP/C composite electrodes were prepared by mixing active LFMP/C materials with Super P and poly(vinyl fluoride) (PVDF) binder at a weight ratio of 80:10:10 [24]. The resulting sample was coated on an Al foil (Aldrich) and dried in a vacuum oven at 80 °C for 12 h. Li foils (Aldrich) were used as the counter and reference electrodes, and a microporous polyethylene (PE) film (Celgard 2400, Charlotte, NC, USA) was used as the separator. The active material loading on Al current collector is around 2–3 mg·cm−2, and the thickness of the dried LFMP/C electrode is about 30 μm. The total thickness of the electrode is 50 μm; the Al current collector is 20 μm. The electrolyte was 1 M of LiPF6 in a mixture of EC and DEC (1:1 in v/v, Merck, Kenilworth, NJ, USA). The LFMP/Li half-cells were charged at a constant current (CC) and a potential range of 2.0–4.5 V (versus Li/Li+) at 0.1C-rate, using a battery tester (Arbin BT2000, College Station, TX, USA). However, the LFMP/Li cells were charged using a CC–CV protocol, and discharged using a CC profile, at a potential range of 2.0–4.5 V (versus Li/Li+) at various C-rates (i.e., 0.2–10C-rates) by using a battery tester (Arbin BT2000). The second CV charge step of 4.5 V was terminated when the charge current was less than 0.1C. The AC spectra of LFMP/Li half-cells were measured at room temperature an Autolab PGSTAT-30 electrochemical system with GPES 4.8 Package Software (Eco Chemie, B.V., Utrecht, The Netherlands).

3. Results and Discussion

Figure 1a shows the typical SEM image of HCS via a hydrothermal carbonation of glucose. The size of HCS is around 1–2 μm; they have hollow and semi-hollow structure. Figure 1b shows the XRD pattern of the HCS sample. Two broad peaks are located at 24° and 43° that correspond to the (002) and (101) planes, indicating the graphitic state of HCS sample. Figure 1c shows the SEM image of NP-GNS sample. They show a highly curved morphology with a nanoporous structure.
Figure 2 shows the XRD patterns of the LFMP/C and LFMP/C + 2% NP-GNS + 2% HCS samples prepared using the wet ball-milling method at 650–700 °C. All of the Bragg diffraction peaks of the compounds were well-indexed based on a single-phase olivine-type structure indexed with the orthorhombic Pnma space group. No impurity peaks, such as Li3PO4 and Mn2P2O7 phases, were found. By comparison, the lattice parameters of the LFP/C, LFMP/C, and SP-LFMP/C composite samples are listed in Table 1. The carbon peaks cannot be observed because of the small amount of the residual carbon content (C% ca. 5%–9%) with amorphous configuration. Fe2+ and Mn2+ ions were located at the tetrahedral 4c sites in LiFe1−yMnyPO4 (y = 0.5). The Mn2+ ion (with an ionic radius of 0.97 Å) has a larger ionic size than the Fe2+ ion (with an ionic radius of 0.92 Å). It was found that the LFMP/C composite materials increased as the Mn2+-doping concentration increased. In addition, the unit cell volume of the LFMP/C composite materials increased linearly with the Mn2+-substitution. As shown in Figure 1, the peaks of the LFMP/C sample (i.e., 301, 121, and 410) shifted to the lower angle direction compared with those of the LFP/C sample [20]. The primary crystal sizes of the LFMP/C and the LFMP/C + 2% NP-GNS + 2% HCS composite materials estimated using Scherrer’s equation were ca. 50–65 nm. The content of residual carbon is studied by elemental analysis (EA) in the composites lists in Table 2. The residual carbon content of the LFMP/C materials with 5 wt% sucrose, 20 wt% citric acid, and 5 wt% BP2000 at different sintering temperatures are around 7.4%–4.8%; in contrast, the residual carbon of the LFMP/C + 2% NP-GNS + 2% HCS composite materials with adding extra 2% NP-GNS + 2 wt% HCS carbon conductor is around 9.31%. Clearly, the residual carbon content decreases when the sintering temperature was increased. It was also found that the electrical conductivity of the LFMP/C + 2% NP-GNS + 2% HCS composite materials can be improved to ca. 10−3 S·cm−1.
Figure 3 shows the SEM images of the LFMP/C-650, LFMP/C-670, LFMP/C-700, and LFMP/C + 2% NP-GNS + 2% HCS-650 composites prepared via the wet ball-milled method. The secondary particle sizes of the LFMP/C composite ranged from 2 μm to 10 μm with a spherical morphology. The surface is highly porous with a macro/nano three-dimensional (3D) hierarchical structure. Figure 3d also shows the SEM images of the LFMP/C + 2% NP-GNS + 2% HCS-650 composites. The LFMP/C + 2% NP-GNS + 2% HCS-650 composite with two conductive carbons added shows a good conductive network because of the uniform coverage of the carbon on the surface of LFMP/C material. In other words, the double carbons greatly improved the coating uniformity on the LFMP/C composite material (see later for transmission electron microscopy (TEM) and micro-Raman results). More importantly, the LFMP/C composite showed slightly smaller particles size around 50 nm; it is due to the high content of carbon being added (C% = 9%). It is well accepted that the carbon precursors can inhibit the LFMP crystalline growth [24]. Therefore, it was well accepted that the smaller particle size of LFPM/C will have much shorter Li+ ion diffusion length. It will greatly improve the electrochemical performance, in particular at high rate. Figure 4 shows the HR-TEM images of the LFMP/C and LFMP/C + 2% NP-GNS + 2% HCS samples sintered at 650–700 °C. It was found that a uniform carbon coating on LFMP/C samples sintered at temperatures of 650, 675, and 700 °C with the thickness of approximately 10, 3, and 2 nm, respectively, was observed. In contrast, Figure 4h presents the TEM image for the LFMP/C + 2% NP-GNS + 2% HCS composite sintered at 650 °C. It was shown the much thick carbon layer is around 4.7 nm.
Figure 5 shows the micro-Raman spectra of the LFMP/C samples and the LFMP/C + 2% NP-GNS + 2% HCS composite, indicating three major Raman vibration peaks at 951 cm−1, 1328 cm−1, and 1585 cm−1. Several Raman peaks at approximately between 947–950 cm−1 and 596–446 cm−1 were identified as the vibration of the P–O bond, and the peak located at 638 cm−1 was identified as the vibration of the FeOx groups. Two Raman peaks of the LFMP/C samples and the LFMP/C + 2% NP-GNS + 2% HCS composite were located at approximately 1328 cm−1 and 1585 cm−1; these peaks were attributed to the residual carbon source. Raman peaks at 1326 cm−1 (D-band) and 1588 cm−1 (G-band) was observed in the LFMP/C + 2% NP-GNS + 2% HCS composite. Broadening the D (A1g symmetry) and G (E2g symmetry) bands with a strong D-band indicated a localized in-plane sp2 graphitic crystal domain and a disordered sp3 amorphous carbon, respectively. The intensity ratio of the D-band to the G-band (i.e., R = ID/IG) was used to estimate the carbon quality of the LFMP/C samples. The R-value of the LFMP/C samples at various sintered temperatures of 650–700 °C (R = 1.11–1.12) was slightly lower than that of the LFMP/C + 2% NP-GNS + 2% HCS composite (R = 1.15). The discharge capacities and rate capabilities of the LFMP/C samples are strongly related to the intensity ratio of the D- and G-bands [22,23,24]. Table 3 shows the R values of all LFMP/C samples by micro-Raman analysis in detail.
Figure 6 shows the initial charge/discharge profiles of three LFMP/C samples, i.e., LFMP-650, LFMP/C-675, and LFMP/C-700, and the LFMP/C + 2% NP-GNS + 2% HCS composite at 0.1C rate and 25 °C, indicating that the LFMP/Li half-cell displayed two flat discharge plateaus at ca. 4.00–4.03 V and ca. 3.50–3.55 V (versus Li/Li+), respectively, associated with two Mn3+/Mn2+ and Fe3+/Fe2+ redox pairs. The LFMP/C, based on a wet ball-milled method, half-cells based on the as-synthesized materials yielded an initial discharge capacity of 148, 139, and 144 mAh·g−1 at 0.1C, respectively. By contrast, as shown in Figure 6, the LFMP/C + 2% NP-GNS + 2% HCS composite showed an initial discharge capacity of around 161.2 mAh·g−1 at 0.1C, indicating that it is superior in performance to the LFMP/C composite with a suitable amount of carbon additives, ca. 4 wt% here. Figure 6 reveals the typical charge/discharge curves of the LFMP/C + 2% NP-GNS + 2% HCS composite at various discharge rates, i.e., at 0.2–10C.
The LFMP/C + 2% NP-GNS + 2% HCS composite exhibited discharge capacities of 154.7, 150.3, 148.8, 136.7, 130.3, and 120.0 mAh·g−1 at charge/discharge rates of 0.2C/0.2C, 0.2C/0.5C, 0.2C/1C, 0.2C/3C, 0.2C/5C, and 0.2C/10C, respectively. These results indicate that the LFMP/C + 2% NP-GNS + 2% HCS composite exhibited excellent high-rate capability and reliable cycle-life stability. In a previous study [17], the discharge capacities of a LiFe0.6Mn0.4PO4/C sample containing 2 wt% MWCNTs and prepared using a spray drying method—were ca. 163.3 mAh·g−1 and 148.7 mAh·g−1 at 0.1C-rate and 1C-rate, respectively. These results are comparable with our results, which are ca. 161 mAh·g−1 and 149 mAh·g−1 at 0.1C and 1C rate, respectively. However, in Zhong et al. [19], the discharge capacities of a LFMP/C material prepared by using a rheological phase method—were 138 mAh·g−1 and 100 mAh·g−1 at 0.1 and 1C-rate, respectively; these findings show that our sample results are clearly superior to their results.
Figure 7 presents a comparison of the rate capability performance of the LFMP/Li half-cell based on the LFMP/C samples (without carbon additives) and the LFMP/C + 2% NP-GNS + 2% HCS composite comprised of extra 2 wt% NP-GNS + 2 wt% HCS carbon conductors, at 0.2–10C rates. We observed that the LFMP/C + 2% NP-GNS + 2% HCS composite exhibited higher rate performance than that of the LFMP/C sample. As shown in Figure 8, the discharge specific capacities of the LFMP/C + 2% NP-GNS + 2% HCS composite sintered at 650 °C decreased from 154 mAh·g−1 to 120 mAh·g−1 as the discharge rate was increased from 0.2C to 10C. By contrast, the discharge-specific capacities of the LFMP/C samples sintered at the same temperature of 650 °C decreased from 141 mAh·g−1 to 113 mAh·g−1. We observed that the rate capability performance of the LFMP/C + 2% NP-GNS + 2% HCS composite with a two-type of carbon conductors, namely, two-dimensional (2D)-graphene and 3D carbon sphere, is greatly improved, as compared to the LFMP/C material without any carbon conductor.
Figure 9A shows the cycle-life performance of the LFMP/C-650, LFMP/C-675, and LFMP/C-700 samples, and the LFMP/C + 2% NP-GNS + 2% HCS composite at charge/discharge rate of 0.1C/0.1C for 30 cycles for comparison. It was found that the discharge capacity of the LFMP/C-650 sample was maintained at 141–139 mAh·g−1 at 0.1C-rate during the 30 cycling test, and the average discharge capacity was approximately 141 mAh·g−1, demonstrating excellent cycle stability with no apparent capacity after the 30-cycle test. We observed that the average current efficiency of the LFMP/C-650 sample at 0.1C was approximately 99.32% during the 30-cycle test. Moreover, the discharge capacity of the LFMP/C-675 sample was maintained at 145–137 mAh·g−1 at 0.1C-rate during the 30 cycles. The average discharge capacity was approximately 137 mAh·g−1 with a fading rate of −0.57%/cycle at 0.1C rate during the cycling test. In addition, it was found that the average coulombic efficiency of the LFMP/C-675 sample at 0.1C was approximately 99.3% during the 30-cycle test. Furthermore, the discharge capacity of the LFMP/C-700 sample was maintained at 148–139 mAh·g−1 at 0.1C-rate during the 30 cycles. The average discharge capacity was approximately 140 mAh·g−1 with a fading rate of −0.63%/cycle at 0.1C rate during the cycling test. In addition, it was revealed that the average current efficiency of the LFMP/C-700 sample at 0.1C was approximately 99.3% during the 30-cycle test.
As expected, the discharge capacity of the LFMP/C + 2% NP-GNS + 2% HCS composite was achieved at 161–155 mAh·g−1 at 0.1C-rate during the 30 cycles. The average discharge capacity was approximately 158 mAh·g−1 with a fading rate of −0.23%/cycle at 0.1C rate during the cycling test. In addition, it was found that the average current efficiency of the LFMP/C + 2% NP-GNS + 2% HCS composite at 0.1C was approximately 98.55% during the 30-cycle test. These results indicated that all LFMP/C sample and the LFMP/C + 2% NP-GNS + 2% HCS composite exhibit excellent and stable electrochemical cycling properties. Figure 9B shows the cycle-life performance of the LFMP/C-650, LFMP/C-675, and LFMP/C-700 samples, and the LFMP/C + 2% NP-GNS + 2% HCS composite at charge/discharge rate of 1C/1C for 100 cycles for comparison. The discharge capacity of the LFMP/C-650 sample was maintained at 122–114 mAh·g−1 at 1C/1C rate during the 100-cycle test, and the average discharge capacity was approximately 118.2 mAh·g−1, demonstrating high cycle stability with no apparent capacity decay after the 100-cycle test. It also observed that the average current efficiency of the LFMP/C-650 sample was approximately over 99%. Moreover, the discharge capacity of the LFMP/C-675 sample was kept at 130–107 mAh·g−1 at a 1C discharge rate during the 100-cycle test. Additionally, the average discharge capacity of the LFMP/C-700 was around 115.7 mAh·g−1 during the 100-cycle test and the average current efficiency of this sample was approximately 97%–98%. Particularly, the discharge capacity of the LFMP/C + 2% NP-GNS + 2% HCS composite was achieved at 141–130 mAh·g−1 at a 1C discharge rate and the average discharge capacity was around 138 mAh·g−1 during the 100-cycle test. It was found that the average current efficiency of this LFMP/C + 2% NP-GNS + 2% HCS composite was approximately 99%. These results clearly demonstrate that the LFMP/C + 2% NP-GNS + 2% HCS composite exhibits excellent and stable electrochemical cycling properties. This may be due to the composite with good electrical conducting channels being built on the electrode material by using 2D graphene and 3D carbon sphere additives. The polarization (the charge transfer resistance) of the LFMP/C + 2% NP-GNS + 2% HCS composite was markedly reduced.
The AC impedance spectroscopy was applied to investigate the interface properties of the LFMP/C samples. The AC spectrum of the LFMP/C samples and the LFMP/C + 2% NP-GNS + 2% HCS composite is illustrated in Figure 10. The equivalent circuit for LFMP/C is displayed in inset of Figure 10. Each AC plot comprised of one semicircle at a higher frequency followed by a linear portion at a lower frequency. The lower frequency region of the straight line was considered as the Warburg impedance, which was used for long-range lithium-ion diffusion in the bulk phase. Rb indicates the bulk resistance at the electrolyte, Rct refers to the charge transfer resistance at the active material interface, and CPE represents the double-layer capacitance and surface film capacitances. Table 4 summarizes the values of the Rb and Rct for all LFMP/C samples. The Rb and Rct values of the LFMP/C-650 sample were approximately 5.46 Ω and 156 Ω, respectively. The Rb and Rct values of the LFMP/C-675 sample were approximately 4.36 Ω and 234 Ω, respectively. Moreover, the Rb and Rct values of the LFMP/C-700 sample were approximately 6.06 Ω and 193.6 Ω, respectively. By contrast, the LFMP/C + 2% NP-GNS + 2% HCS composite show 5.3 Ω and 63.95 Ω for the Rb and Rct values. In fact, the LFMP/C + 2% NP-GNS + 2% HCS composite showed the lowest charge transfer resistance, as compared with other LFMP/C samples. We prepared the LFMP/C composite material with 2D and 3D carbon additives, and they can build up the electrical conducting channels within LFMP electrode, which facilitate the electron transport and also greatly reduce the electrode polarization. In conclusion, the LFMP/C + 2% NP-GNS + 2% HCS composite can be prepared by a simple one-pot solid-state ball-milled process. The LFMP/C + 2% NP-GNS + 2% HCS composite showed excellent electrochemical performance; this is due to the reduction of the charge transfer resistance.

4. Conclusions

In this work, a LFMP/C material was first prepared using a wet ball-milling solid-state method. A LFMP/C + 2% NP-GNS + 2% HCS composite with mixed carbon conductors was then prepared and examined. The properties of the resulting samples were examined using XRD, SEM, TEM, micro-Raman spectroscopy, and the AC impedance spectroscopy, and galvanostatic charge-discharge methods. The LFMP/C + 2% NP-GNS + 2% HCS composite exhibited discharge capacities of 154.7, 150.3, 148.8, 136.7, 130.3, and 120.0 mAh·g−1 at charge/discharge rates of 0.2C/0.2C, 0.2C/0.5C, 0.2C/1C, 0.2C/3C, 0.2C/5C, and 0.2C/10C, respectively. These results indicate that the LFMP/C + 2% NP-GNS + 2% HCS composite exhibited excellent high-rate capability and reliable cycle-life stability. This work demonstrated that the excellent performance of the LFMP/C can be synthesized by a one-pot solid-state method when the suitable amount of carbon conducting additives (optimum around 4 wt%) was added. Those carbon conductors play a key role in improving electrochemical performance. This study shows that the LFMP/C composite material is a good candidate for application in high-power Li-ion batteries.

Acknowledgments

Financial support from the Ministry of Science and Technology, Taiwan (Project No: MOST 103-2632-E-131 -001) is gratefully acknowledged.

Author Contributions

Chun-Chen Yang and Shinhjiang Jessie Lue wrote the paper, Yen-Wei Hung designed the experiment and did the electrochemical performance analysis. All authors examined and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Padhi, A.K.; Najundaswamy, K.S.; Goodenough, J.B. The root cause of the rate performance improvement after metal doping: A case study of LiFePO4. J. Electrochem. Soc. 1997, 144, 1188–1194. [Google Scholar] [CrossRef]
  2. Oh, S.; Myung, S.; Park, J.; Scrosati, B.; Amine, K.; Sun, Y. Double-structured LiMn0.85Fe0.15PO4 coordinated with LiFePO4 for rechargeable lithium batteries. Angew. Chem. 2012, 51, 1853–1856. [Google Scholar] [CrossRef] [PubMed]
  3. Oh, S.; Jung, H.; Yoon, C.; Myung, S.; Chen, Z.; Amine, K.; Sun, Y. Enhanced electrochemical performance of carbon-LiMn1−xFexPO4 nanocomposite cathode for lithium-ion batteries. J. Power Sources 2011, 196, 6924–6928. [Google Scholar] [CrossRef]
  4. Osorio-Guillén, J.; Holm, B.; Ahuja, R.; Johansson, B. A theoretical study of olivine LiMPO4 cathodes. Solid State Ion. 2004, 167, 221–227. [Google Scholar] [CrossRef]
  5. Yao, J.; Bewlay, S.; Konstantionv, K.; Drozd, V.; Liu, R.; Wang, X.; Liu, H.; Wang, G. Characterisation of olivine-type LiMn1−xFexPO4 cathode materials. J. Alloys Compd. 2006, 425, 362–366. [Google Scholar] [CrossRef]
  6. Mi, C.H.; Zhang, X.G.; Zhao, X.B.; Li, H.L. Synthesis and performance of LiMn0.6Fe0.4PO4/nano-carbon webs composite cathode. Mater. Sci. Eng. B 2006, 129, 8–13. [Google Scholar] [CrossRef]
  7. Burba, C.; Frech, R. Local structure in the Li-ion battery cathode material Lix(MnyFe1−y)PO4 for 0 < x ≤ 1 and y = 0.0, 0.5 and 1.0. J. Power Sources 2007, 172, 870–876. [Google Scholar]
  8. Shin, Y.; Kim, J.; Cheruvally, G.; Ahn, J.; Kim, K. Li(Mn0.4Fe0.6)PO4 cathode active material: Synthesis and electrochemical performance evaluation. J. Phys. Chem. Solids 2008, 69, 1253–1256. [Google Scholar] [CrossRef]
  9. Zaghib, K.; Mauger, A.; Gendron, F.; Massot, M.; Julien, C. Insertion properties of LiFe0.5Mn0.5PO4 electrode materials for Li-ion batteries. Ionics 2008, 14, 371–376. [Google Scholar] [CrossRef]
  10. Chen, Y.C.; Chen, J.M.; Hsu, C.H.; Yeh, J.W.; Shih, H.C.; Chang, Y.S.; Sheu, H.S. Structure studies on LiMn0.25Fe0.75PO4 by in-situ synchrotron X-ray diffraction analysis. J. Power Sources 2009, 189, 790–793. [Google Scholar] [CrossRef]
  11. Baek, D.H.; Kim, J.K.; Shin, Y.J.; Chauhan, G.S.; Ahn, J.H.; Kim, K.W. Effect of firing temperature on the electrochemical performance of LiMn0.4Fe0.6PO4/C materials prepared by mechanical activation. J. Power Sources 2009, 189, 59–65. [Google Scholar] [CrossRef]
  12. Kim, J.; Chauhan, G.; Ahn, J.; Ahn, H. Effect of synthetic conditions on the electrochemical properties of LiMn0.4Fe0.6PO4/C synthesized by sol-gel technique. J. Power Sources 2009, 189, 391–396. [Google Scholar] [CrossRef]
  13. Chen, Y.C.; Chen, J.M.; Hsu, C.H.; Lee, J.F.; Yeh, J.W.; Shih, H.C. In-situ synchrotron X-ray absorption studies of LiMn0.25Fe0.75PO4 as a cathode material for lithium ion batteries. Solid State Ion. 2009, 180, 1215–1219. [Google Scholar] [CrossRef]
  14. Martha, S.K.; Grinblat, J.; Haik, O.; Zinigrad, E.; Drezen, T.; Miners, J.H.; Exnar, I.; Kay, A.; Markovsky, B.; Aurbach, D. LiMn0.8Fe0.2PO4: An advanced cathode material for rechargeable lithium batteries. Angew. Chem. Int. Ed. 2009, 48, 8559–8563. [Google Scholar] [CrossRef] [PubMed]
  15. Zou, Q.Q.; Zhu, G.N.; Xia, Y.Y. Preparation of carbon-coated LiFe0.2Mn0.8PO4 cathode material and its application in a novel battery with Li4Ti5O12 anode. J. Power Sources 2012, 206, 222–229. [Google Scholar] [CrossRef]
  16. Liu, T.; Wu, B.; Wu, X. Realizing Fe substitution through diffusion in preparing LiMn1−xFexPO4-C cathode materials from MnPO4·H2O. Solid State Ion. 2014, 254, 72–77. [Google Scholar] [CrossRef]
  17. Mi, Y.Y.; Gao, P.; Liu, W.; Zhang, W.; Zhou, H. Carbon nanotube-loaded mesoporous LiFe0.6Mn0.4PO4/C microspheres as high performance cathodes for lithium-ion batteries. J. Power Sources 2014, 267, 459–468. [Google Scholar] [CrossRef]
  18. Hu, L.; Qiu, B.; Xia, Y.; Qin, Z.; Qin, L.; Zhou, X.; Liu, Z. Solvothermal synthesis of Fe-doping LiMnPO4 nanomaterials for Li-ion batteries. J. Power Sources 2014, 248, 246–252. [Google Scholar] [CrossRef]
  19. Zhong, Y.J.; Li, J.T.; Wu, Z.G.; Guo, X.D.; Zhong, B.H.; Sun, S.G. LiMn0.5Fe0.5PO4 solid solution materials synthesized by rheological phase reaction and their excellent electrochemical performances as cathode of lithium ion battery. J. Power Sources 2013, 234, 217–222. [Google Scholar] [CrossRef]
  20. Scosati, B.; Garche, J. Lithium batteries: Status, prospects and future. J. Power Sources 2010, 195, 2419–2430. [Google Scholar] [CrossRef]
  21. Goriparti, S.; Miele, E.; Angelis, F.D.; Fabrizio, E.D.; Zaccaria, R.P.; Capiglia, C. Review on recent progress of nanostructured anode materials for Li-ion batteries. J. Power Sources 2014, 257, 421–443. [Google Scholar] [CrossRef]
  22. Zhao, W.; Liu, Y.; Zhang, X. Preparation and characterization of hollow Co3O4 spheres. Mater. Lett. 2008, 62, 772–774. [Google Scholar] [CrossRef]
  23. Zhang, J.; Guo, B.; Yang, Y.; Shen, W.; Wang, Y.; Zhou, X.; Wu, H.; Guo, S. Gas molecule adsorption in carbon nanotubes and nanotube bundles. Carbon 2015, 84, 467–478. [Google Scholar]
  24. Yang, C.C.; Chen, Y.C.; Liao, Y.C. Comparison of electrochemical performances of LiFePO4/C composite materials by two preparation routes. Mater. Res. Bull. 2012, 47, 2622–2626. [Google Scholar] [CrossRef]
Figure 1. (a) Scanning electron microscopy (SEM) image for hollow carbon sphere (HCS); (b) X-ray diffraction (XRD) pattern for HCS; and (c) SEM image for nanoporous graphene (NP-GNS) at 3k×.
Figure 1. (a) Scanning electron microscopy (SEM) image for hollow carbon sphere (HCS); (b) X-ray diffraction (XRD) pattern for HCS; and (c) SEM image for nanoporous graphene (NP-GNS) at 3k×.
Batteries 02 00018 g001
Figure 2. XRD patterns of all LiFe0.5Mn0.5PO4/C (LFMP/C) sample.
Figure 2. XRD patterns of all LiFe0.5Mn0.5PO4/C (LFMP/C) sample.
Batteries 02 00018 g002
Figure 3. SEM images of LFMP/C sample in 5 k×: (a) LFMP/C-650; (b) LFMP/C-675; (c) LFMP/C-700; and (d) LFMP/C + 2% NP-GNS + 2% HCS-650.
Figure 3. SEM images of LFMP/C sample in 5 k×: (a) LFMP/C-650; (b) LFMP/C-675; (c) LFMP/C-700; and (d) LFMP/C + 2% NP-GNS + 2% HCS-650.
Batteries 02 00018 g003
Figure 4. High-resolution transmission electron microscopy (HR-TEM) images of LFMP/C sample: (a,b) LFMP/C-650; (c,d) LFMP/C-675; (e,f) LFMP/C-700; and (g,h) LFMP/C + 2% NP-GNS + 2% HCS-650.
Figure 4. High-resolution transmission electron microscopy (HR-TEM) images of LFMP/C sample: (a,b) LFMP/C-650; (c,d) LFMP/C-675; (e,f) LFMP/C-700; and (g,h) LFMP/C + 2% NP-GNS + 2% HCS-650.
Batteries 02 00018 g004
Figure 5. Micro-Raman curve of all LFMP/C samples.
Figure 5. Micro-Raman curve of all LFMP/C samples.
Batteries 02 00018 g005
Figure 6. The initial sample charge/discharge curves of LFMP/C samples at 0.1C/0.1C at 25 °C.
Figure 6. The initial sample charge/discharge curves of LFMP/C samples at 0.1C/0.1C at 25 °C.
Batteries 02 00018 g006
Figure 7. The rate capability curves of all LFMP/C samples at 0.2–10C.
Figure 7. The rate capability curves of all LFMP/C samples at 0.2–10C.
Batteries 02 00018 g007
Figure 8. The charge/discharge curves of the LFMP/C + 2% NP-GNS + 2% HCS composite sintered at 650 °C at 0.2–10C rates.
Figure 8. The charge/discharge curves of the LFMP/C + 2% NP-GNS + 2% HCS composite sintered at 650 °C at 0.2–10C rates.
Batteries 02 00018 g008
Figure 9. The cycle-life performance of all LFMP/C samples at: (A) 0.1C/0.1C rate; and (B) 1C/1C rate.
Figure 9. The cycle-life performance of all LFMP/C samples at: (A) 0.1C/0.1C rate; and (B) 1C/1C rate.
Batteries 02 00018 g009
Figure 10. The Nyquist plots of all LFMP/C samples at open circuit potential (OCP); the inset for equivalent circuit.
Figure 10. The Nyquist plots of all LFMP/C samples at open circuit potential (OCP); the inset for equivalent circuit.
Batteries 02 00018 g010
Table 1. The lattice parameters of all LFMP/C samples from XRD analysis.
Table 1. The lattice parameters of all LFMP/C samples from XRD analysis.
SampleLFMP/C-650LFMP/C + 2% NP-GNS + 2% HCS-650LFMP/C-675LFMP/C-700
a (nm)0.472160.471380.471840.47177
b (nm)1.039571.038261.039261.03904
c (nm)0.605750.604990.605450.60542
V (nm3)0.297330.296090.296900.29686
Table 2. The residual carbon contents of all LFMP/C samples by elemental analysis (EA).
Table 2. The residual carbon contents of all LFMP/C samples by elemental analysis (EA).
SampleSample#1/mgSample#2/mgC#1/wt%C#2/wt%Cavg/wt%
LFMP/C-6501.882.337.347.407.37
LFMP/C-6751.972.45.195.315.25
LFMP/C-7001.711.944.764.854.81
LFMP/C + 2% NP-GNS + 2% HCS-6502.111.889.319.319.31
Table 3. Micro-Raman spectroscopic analysis of LFMP/C samples.
Table 3. Micro-Raman spectroscopic analysis of LFMP/C samples.
SamplesIDIGPO43−R = ID/IG(ID + IG)/PO43−
LFMP/C-65019,62717,65637611.1129.91
LFMP/C-67524,02120,94077731.1475.78
LFMP/C-70015,96214,16642861.1247.03
LFMP/C + 2% NP-GNS + 2% HCS-65022,39919,50882701.1485.07
Table 4. The parameters of alternating current (AC) impedance analysis.
Table 4. The parameters of alternating current (AC) impedance analysis.
SampleRb (Ω)Rct (Ω)
LFMP/C-6505.46156.24
LFMP/C-6754.36234.36
LFMP/C-7006.06193.66
LFMP/C + 2% NP-GNS + 2% HCS-6505.3063.95

Share and Cite

MDPI and ACS Style

Yang, C.-C.; Hung, Y.-W.; Lue, S.J. The Carbon Additive Effect on Electrochemical Performance of LiFe0.5Mn0.5PO4/C Composites by a Simple Solid-State Method for Lithium Ion Batteries. Batteries 2016, 2, 18. https://doi.org/10.3390/batteries2020018

AMA Style

Yang C-C, Hung Y-W, Lue SJ. The Carbon Additive Effect on Electrochemical Performance of LiFe0.5Mn0.5PO4/C Composites by a Simple Solid-State Method for Lithium Ion Batteries. Batteries. 2016; 2(2):18. https://doi.org/10.3390/batteries2020018

Chicago/Turabian Style

Yang, Chun-Chen, Yen-Wei Hung, and Shingjiang Jessie Lue. 2016. "The Carbon Additive Effect on Electrochemical Performance of LiFe0.5Mn0.5PO4/C Composites by a Simple Solid-State Method for Lithium Ion Batteries" Batteries 2, no. 2: 18. https://doi.org/10.3390/batteries2020018

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop