Effect of Temperature on the Morphology and Anode Performance of

Oct 13, 2017 - Effect of Temperature on the Morphology and Anode Performance of Binder-Free Carbon Nanofiber/Nickel–Cobalt Hydroxide and Carbon ... ...
1 downloads 14 Views 8MB Size
Article Cite This: J. Phys. Chem. C XXXX, XXX, XXX-XXX

pubs.acs.org/JPCC

Effect of Temperature on the Morphology and Anode Performance of Binder-Free Carbon Nanofiber/Nickel−Cobalt Hydroxide and Carbon Nanofiber/Nickel−Cobalt Oxide for Lithium-Ion Batteries Jia-Min Syu, Ming-Liang Hsiao, and Chieh-Tsung Lo* Department of Chemical Engineering, National Cheng Kung University, number 1, University Road, Tainan City 701, Taiwan S Supporting Information *

ABSTRACT: Hierarchical composite materials composed of electrospun carbon nanofibers and binary nickel−cobalt hydroxide (Ni−Co(OH)2) nanoflakes were synthesized for use as anodes for lithium-ion batteries. To clarify the effect of temperature on the structure and electrochemical performance of the composites, the ternary composite fibers were annealed at various temperatures. When the annealing temperature increased, the interconnected Ni−Co(OH)2 nanoflakes on the carbon surface gradually converted to Ni−Co(OH)2 nanoneedles. Subsequently, the chemical structure of Ni−Co(OH)2 was transformed into binary nickel−cobalt oxide (Ni−CoO) at 300 °C. Further annealing at 400 °C caused the aggregation of nanoneedles, forming flowerlike particles. The incorporation of Ni−Co(OH)2 nanoflakes into carbon nanofibers resulted in a substantial increase in the specific capacity compared with that of the pure carbon nanofibers. The increased specific capacity was because of the synergistic effects of the high theoretical capacity of Ni−Co(OH)2 and easily accessible charge transport in the carbon nanofiber network. The specific capacity of the composite fibers increased with the annealing temperature and reached a maximum of 645 mAh/g at a current density of 150 mA/g at 300 °C. The increased specific capacity was mainly attributed to the change in the specific surface area, mesopore volume, and electrical conductivity of the composite fibers with temperature.



INTRODUCTION To satisfy the increasing demand for multifunctional consumer electronics, electric vehicles, and renewable energy over the past few decades considerable effort has been devoted to developing high-performance energy storage systems. Lithiumion batteries are considered a promising energy storage device because of their high energy density, high working voltage, long cycle stability, and absence of a memory effect. Electrode materials are the major component of lithium-ion batteries that determine their electrochemical performance, and researchers have devoted substantial effort to search for high-performance electrode materials to meet the requirements of the next generation of lithium-ion batteries. Among the materials, transition metal hydroxides and metal oxides are promising anode candidates because they exhibit much higher theoretical capacities than graphite, as well as being low-cost and abundant materials. However, they are associated with the disadvantage of quick capacity fading caused by large volume expansion/ contraction during charging and discharging. Additionally, the aggregation of transition metal hydroxides and metal oxides during the cycling process limits the kinetics of ion diffusion, thereby reducing the performance of batteries. Strategies for improving the electrochemical performance of transition metal hydroxides and metal oxides include the manipulation of morphology and the synthesis of binary mixtures. Nanostructured anodes provide a highly accessible surface area to electrolytes, enabling efficient ion diffusion to electrodes. Additionally, three-dimensional nanostructures offer open space for volume expansion during the cycling process, © XXXX American Chemical Society

thereby promoting electrode stability. Moreover, the combination of two types of transition metal oxides or metal hydroxides, such as binary nickel−cobalt oxide (Ni−CoO)1−5 and binary nickel−cobalt hydroxide (Ni−Co(OH)2),6 may induce the synergistic effects of the two compounds and further surpass the intrinsic properties of the single material. For example, the addition of Co ions to Ni(OH)2 decreased mechanical stress during charging and discharging, preventing electrode failure.7,8 Additionally, mixed materials induced the catalytic decomposition of the solid electrolyte interphase (SEI) during charging, leading to enhanced electrochemical performance.9−11 Extensive research has been conducted for designing nanostructured Ni−CoO or Ni−Co(OH)2 for use as anodes for lithium-ion batteries. Mai et al. synthesized Co-doped NiO nanoflake arrays by using low-temperature chemical bath deposition.1 The presence of Co2+ increased the electrical conductivity of the anode, facilitating charge transport. Consequently, Co-doped NiO nanoflake arrays delivered a specific capacity of 600 mAh/g at a current density of 100 mA/ g for the 50th cycle. By contrast, the pure NiO electrode exhibited a specific capacity of only 463 mAh/g at the same current density. Similarly, Thi et al. prepared anodes composed of Co-doped NiO nanoparticles by using a solvothermal approach.4 The uniform distribution of Co in the crystal lattice Received: July 26, 2017 Revised: September 28, 2017 Published: October 13, 2017 A

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

preparation of carbon fibers through electrospinning can generate fibers on the nanometer scale, providing a large surface area. Additionally, electrospun carbon nanofibers form a web-type structure associated with fiber entanglement. Because of their favorable mechanical and electrical properties, anodes prepared using electrospun carbon nanofibers are binder-free and flexible. However, previous studies have described only a few examples of anodes that synergize the intrinsic properties of nickel hydroxide, cobalt hydroxide, and electrospun carbon nanofibers in fabricating the ternary anodes for lithium-ion batteries.15 In this study, we analyzed the change in the microstructure and electrochemical performance of these ternary anodes with variation of the annealing temperature. As indicated in the literature,2,5,13 thermal treatment on Ni− CoO at different annealing temperatures resulted in morphological changes. These morphologies further altered the microstructure of Ni−CoO, such as the specific surface area and pore size, and consequently affected the anode performance of Ni−CoO. Furthermore, elevated temperatures on the ternary anodes could affect the interfacial integrity between carbon nanofibers and Ni−CoO. This information is crucial for manipulating the structure of the anodes and further developing anode materials for lithium-ion batteries.

of NiO prevented particle aggregation, leading to an increased surface area of NiO. Wang and Zhang synthesized Ni−CoO/C composites by using a coprecipitation and subsequent pyrolytic carbon reduction method.2 The morphology of the composites strongly depended on the thermal treatment. Flowerlike spheres with a diameter of 500 nm were observed at a low annealing temperature, whereas a high annealing temperature resulted in the coexistence of spherical and polyhedral particles with nonuniform particle sizes. This was attributed to the intense crystallization at a high annealing temperature, which caused a destruction of the flowerlike spheres. As an anode for lithium-ion batteries, the composite annealed at a low temperature exhibited the poor cycle stability, which resulted from the poor mechanical integrity of the composite with a flowerlike structure. Zhang et al. synthesized a nanostructured anode consisting of NiO nanoplates and Co3O4 nanoparticles by using a hydrothermal method.12 The NiO nanoplates offered a large surface area, allowing short pathways for ion diffusion, and the Co3O4 nanoparticles acted as a buffer, preventing nanoplate aggregation. The Ni−CoO anode delivered a specific capacity of 633 mAh/g at a current density of 100 mA/g for the 70th cycle. Huang et al. synthesized Ni− CoO mesoporous thorn microspheres by conducting a hydrothermal reaction with urea as a template.5 The unique hierarchical structure of the Ni−CoO microspheres provided both a short diffusion path and a large active surface area for lithium ions. By varying the annealing temperature, the structure of the microspheres changed from the integrated to the stone-pillar-like thorn bundle arrays, which resulted in a substantial decrease in the specific surface area of the microspheres. Furthermore, thermal treatment at 600 °C yielded multiphase crystal structures composed of NiCo2O4, Co3O4, and NiO, which hindered side reactions of the anode with the electrolyte and further reduced the volume expansion of the anode during charging and discharging. Consequently, the Ni−CoO microspheres thermally treated at a low temperature delivered a higher capacity but a lower capacity retention when compared to those annealed at a high temperature. Xu et al. demonstrated the effect of the annealing temperature on the electrode materials composed of Ni−CoO and carbonized kapok fibers.13 Increasing the annealing temperature induced a morphological transition of the composites from the cross-linked fibrous network to nanoflakes, which subsequently enhanced the electrochemical performance of the composites. Ding et al. synthesized Ni− CoO microspheres by using a bubble-template process. The hollow structures provided not only an efficient contact area between the electrodes and electrolyte but also a large open space for accommodating volume changes.14 Consequently, the Ni−CoO hollow microspheres delivered a specific capacity of 730 mAh/g at a current density of 300 mA/g for the 140th cycle. Although Ni−CoO and Ni−Co(OH)2 exhibit excellent electrochemical performance, the addition of binders during electrode preparation is inevitable. Both the amount of binders and the mixing process should be optimized. Additionally, the as-prepared electrodes are not flexible enough to meet the requirements of some developing consumer electronics. In this study, we integrated ternary mixtures of nickel hydroxide, cobalt hydroxide, and electrospun carbon nanofibers as anodes for lithium-ion batteries. Carbon fibers exhibit unique carbon structures that contribute to the high electrical conductivity, high mechanical strength, and lightweight of the fibers, making them useful in many applications. The



EXPERIMENTAL SECTION

Materials and Methods. Electrospun carbon nanofibers were synthesized using polyacrylonitrile (PAN, molecular weight: 150000 g/mol, Sigma-Aldrich Co.) dissolved in N,Ndimethylformamide (DMF, 99.8%) as a precursor solution. The 7 wt % precursor solution was added to a 10 mL glass syringe placed 15 cm from a collector composed of vertically placed aluminum foil. The conventional single-nozzle electrospinning process was conducted at room temperature by using a power supply of 15 kV (You-Shang Technical Corporation). The flow rate of the precursor solution was maintained at 0.7 mL/h by using a syringe pump (KDS 100, KD Scientific). After electrospinning, the as-prepared fiber mats were dried in air to completely remove the residual solvent prior to thermal treatment. The polymer nanofibers were converted to carbon nanofibers through a two-step thermal treatment performed in a tubular quartz furnace. The fiber mats were stabilized by heating in air from room temperature to 250 °C at a heating rate of 2 °C/ min, and the temperature was then maintained at 250 °C for 4 h. Subsequently, carbonization was conducted at 800 °C for 2 h in argon. The carbonized nanofibers were surface-oxidized by using 68 wt % nitric acid at 75 °C for 5 h. Carbon nanofiber/Ni−Co(OH)2 composites were synthesized using a modification of previously described procedures.14,16 In brief, the HNO3-treated carbon nanofibers were mixed with 0.2 g of NiCl2·6H2O, 0.2 g of CoCl2·6H2O, 1.2 g of cetyltrimethylammonium bromide, and 3.0 g of urea in 100 mL of deionized (DI) water. The reaction was performed at 100 °C for 11 h. The resulting carbon nanofiber/Ni−Co(OH)2 composites were subjected to purification through DI water to remove residual chemicals. Subsequently, the composites were annealed at various temperatures for 3 h in nitrogen. The nomenclature of the fibers is as follows: CNF for the carbon nanofibers prepared from neat PAN; CNF-NiCo for the carbon nanofiber/Ni−Co(OH)2 composites without annealing; and CNF−NiCo200, CNF−NiCo300, and CNF−NiCo400 for the composite fibers annealed at 200, 300, and 400 °C, respectively. B

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C Characterization. The surface morphology of the composite fibers was examined using scanning electron microscopy (SEM) on a Hitachi SU8010 scanning electron microscope at an accelerating voltage of 10 kV. Prior to SEM imaging, the composite fibers were sputter-coated with a thin layer of platinum by using a JEOL JFC-1600 coater. The mean fiber diameter was analyzed by observing at least 100 randomly selected fibers per sample on several SEM images. X-ray diffraction (XRD) patterns were measured using a Rigaku RINT-2000 diffractometer with a filtered Cu Kα radiation source (wavelength = 1.5406 Å) at 40 kV and 40 mA. The measurements were performed with a diffraction angle (2θ) between 4° and 70° at a scan rate of 2°/min. X-ray photoelectron spectroscopy (XPS) was conducted on a PHI 5000 VersaProbe modeled X-ray photoelectron spectrometer to record both elemental and chemical information on the composite fibers. The specific surface area and pore size distribution of the composite fibers were determined from the nitrogen sorption isotherms generated using the Brunauer− Emmett−Teller (BET) and Barrett−Joyner−Halenda (BJH) methods, respectively, using a physisorption analyzer (ASAP2020, Micromeritics). Prior to the measurements, the composite fibers were degassed at 200 °C under vacuum to remove other adsorbed species. The four-point probe method was used to measure the electrical conductivity (σ) of the composite fibers, in which σ was calculated using the following equation: σ=

1 Rw

(1) Figure 1. Surface morphology of composite fibers. (a) CNF, (b) CNFNiCo, (c) CNF-NiCo200, (d) CNF-NiCo300, and (e) CNFNiCo400. Insets: high-magnification images of the composite fibers.

where R is the electrical resistance and w is the thickness of the specimen. The electrochemical performance of the composite fibers was analyzed at room temperature by using 2032 coin cells as the lithium-ion batteries (Ubiq Technology Co., Ltd.). The synthesized composite fibers were directly used as a binderfree working electrode, and lithium metal (99.9%, Hongda energy) was used for both the counter and reference electrodes. A PP−PE−PP trilayer (Celgard M824, 12 μm) served as the separator. The electrolyte was prepared by dissolving LiPF6 in a mixture of ethylene carbonate/diethyl carbonate/dimethyl carbonate (vol/vol/vol = 1/1/1) with a concentration of 1 M. The cells were assembled inside an argon-filled glovebox. Cyclic voltammetry (CV) measurements were performed to examine the electrochemical reactivity of the composite fibers by using a CHI-627D potentiostat with a potential window of 0.01−3.50 V at a scan rate of 0.3 mV/s. Galvanostatic charge− discharge tests were performed at potentials ranging from 0.01 to 2.80 V versus Li/Li+ at a constant current density of 150− 900 mA/g by using a BAT750B battery test system (Acutech Systems Co., Ltd.). A total of 100 cycles at a current density of 200 mA/g were performed for cycling stability testing. For the full cell, the cathode was prepared by casting a mixture of LiFePO4, carbon black, and polyvinylidine fluoride in Nmethyl-2-pyrrolidone solution with the weight ratio of 7.5:1.5:1 on aluminum foil. The cycling performance of the full cell was carried out at a constant current density of 50 mA/g.

exhibited defect-free and bead-free fibrous morphology, with a mean fiber diameter of 234 ± 53 nm. The fibers were randomly oriented, forming a netlike structure. When Ni and Co were coprecipitated in the presence of urea, urea dissociated with water to form hydroxyl ions and ammonium ions. The hydroxyl ions further reacted with Ni and Co ions, forming Ni(OH)2 and Co(OH)2. The chemical reactions are as follows:14 CO(NH 2)2 + 3H 2O → 2NH4 + + CO2 + 2OH−

(2)

M2 + + 2OH− → M(OH)2

(3)

(M = Ni/Co)

In Figure 1b, after the coprecipitation of Ni(OH)2 and Co(OH)2 on the fiber surface, Ni−Co(OH)2 exhibited a flakelike, vertically interconnected structure on the surface of the carbon nanofibers. This close connection between Ni− Co(OH)2 and the carbon nanofibers provided excellent electron-diffusion paths. The attachment of Ni−Co(OH)2 to the surface of the carbon nanofibers considerably increased the fiber diameter to 347 ± 64 nm. Furthermore, the as-prepared carbon nanofiber/Ni−Co(OH)2 composites were thermally annealed at various temperatures in nitrogen. After annealing, the composite fibers retained their hierarchical structure; with increasing temperature, the Ni−Co(OH)2 nanoflakes gradually converted to a needlelike structure (Figure 1, panels c and d). When the annealing temperature reached 400 °C, the nanoneedles aggregated to form flowerlike particles randomly dispersed on the fiber surface (Figure 1e). A thermogravimetric analysis (TGA) thermogram (Figure S1) indicated that when Ni(OH)2 and Co(OH)2 were thermally annealed at a



RESULTS AND DISCUSSION Physicochemical Properties of Composite Fibers. Figure 1 shows the representative SEM images of electrospun carbon nanofibers and their composites with Ni−Co(OH)2 annealed at various temperatures. The CNF sample (Figure 1a) C

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C Table 1. Properties of Composite Fibers sample CNF CNF-NiCo CNF-NiCo200 CNF-NiCo300 CNF-NiCo400

D (nm)a

SA (m2/g)b

VT (cm3/g)c

Vmeso (cm3/g)d

Vmicro (cm3/g)e

σ (S/m)f

± ± ± ± ±

15.3 18.4 22.8 50.3 34.1

0.037 0.047 0.054 0.067 0.067

0.035 0.047 0.054 0.060 0.065

0.002 0 0 0.007 0.002

295 76 82 134 187

234 347 366 302 258

53 64 56 50 54

a D: mean fiber diameter. bSA: specific surface area. cVT: total pore volume. dVmeso: pore volume at a pore size of 1.7−300 nm. eVmicro: pore volume at a pore size of 650 mAh/g in the first cycle to 95% at 100 mA/g 77.5% at 300 mA/g ∼75% at 100 mA/g not reported

this study [1] [2] [3] [4] [12] [14] [15] [42]

K

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C



Networks as a Promising Electrode for Supercapacitor Applications. J. Mater. Chem. 2012, 22, 5656−5665. (20) Cheng, Y.; Zhang, H.; Varanasi, C. V.; Liu, J. Improving the Performance of Cobalt-Nickel Hydroxide-Based Self-Supporting Electrodes for Supercapacitors Using Accumulative Approaches. Energy Environ. Sci. 2013, 6, 3314−3321. (21) Lai, F.; Huang, Y.; Miao, Y. − E.; Liu, T. Controllable Preparation of Multi-Dimensional Hybrid Materials of Nickel-Cobalt Layered Double Hydroxide Nanorods/Nanosheets on Electrospun Carbon Nanofibers for High-Performance Supercapacitors. Electrochim. Acta 2015, 174, 456−463. (22) Lai, F.; Miao, Y. − E.; Zuo, L.; Lu, H.; Huang, Y.; Liu, T. Biomass-Derived Nitrogen-Doped Carbon Nanofiber Network: A Facile Template for Decoration of Ultrathin Nickel-Cobalt Layered Double Hydroxide Nanosheets as High-Performance Asymmetric Supercapacitor Electrode. Small 2016, 12, 3235−3244. (23) Li, Y.; Peng, H.; Wei, X.; Xiao, P. Controlled Growth of Hierarchical Nickel and Cobalt Hybrid Inorganic-Organic NanosheetSupported Nanowires for Energy Storage. CrystEngComm 2017, 19, 1555−1563. (24) Goncalves, J. M.; Guimarães, R. R.; Nunes, C. V., Jr.; Duarte, A.; Brandao, B. B. N. S.; Toma, H. E.; Araki, K. Electrode Materials Based on Alpha-NiCo(OH)2 and rGO for High Performance Energy Storage Devices. RSC Adv. 2016, 6, 102504−102512. (25) Hall, D. S.; Lockwood, D. J.; Bock, C.; MacDougall, B. R. Nickel Hydroxides and Related Materials: A Review of Their Structures, Synthesis and Properties. Proc. R. Soc. London, Ser. A 2015, 471, 1−65. (26) Chen, Y.; Pang, W. K.; Bai, H.; Zhou, T.; Liu, Y.; Li, S.; Guo, Z. Enhanced Structural Stability of Nickel-Cobalt Hydroxide via Intrinsic Pillar Effect of Metaborate for High-Power and Long-Life Supercapacitor Electrodes. Nano Lett. 2017, 17, 429−436. (27) Hu, Q.; Gu, Z.; Zheng, X.; Zhang, X. Three-Dimensional Co3O4@NiO Hierarchical Nanowire Arrays for Solid-State Symmetric Supercapacitor with Enhanced Electrochemical Performances. Chem. Eng. J. 2016, 304, 223−231. (28) Zhang, M.; Uchaker, E.; Hu, S.; Zhang, Q.; Wang, T.; Cao, G.; Li, J. CoO-Carbon Nanofiber Networks Prepared by Electrospinning as Binder-Free Anode Materials for Lithium-Ion Batteries with Enhanced Properties. Nanoscale 2013, 5, 12342−12349. (29) Kang, L.; Deng, J.; Liu, T.; Cui, M.; Zhang, X.; Li, P.; Li, Y.; Liu, X.; Liang, W. One-Step Solution Combustion Synthesis of CobaltNickel Oxides/C/Ni/CNTs Nanocomposites as Electrochemical Capacitors Electrode Materials. J. Power Sources 2015, 275, 126−135. (30) Ryu, Z.; Zheng, J.; Wang, M.; Zhang, B. Characterization of Pore Size Distributions on Carbonaceous Adsorbents by DFT. Carbon 1999, 37, 1257−1264. (31) Zhang, J.; Liu, F.; Cheng, J. P.; Zhang, X. B. Binary NickelCobalt Oxides Electrode Materials for High-Performance Supercapacitors: Influence of Its Composition and Porous Nature. ACS Appl. Mater. Interfaces 2015, 7, 17630−17640. (32) Li, C.; Yin, X.; Chen, L.; Li, Q.; Wang, T. Porous Carbon Nanofibers Derived from Conducting Polymer: Synthesis and Application in Lithium-Ion Batteries with High-Rate Capability. J. Phys. Chem. C 2009, 113, 13438−13422. (33) Higuchi, E.; Otsuka, H.; Chiku, M.; Inoue, H. Effect of Pretreatment on the Surface Structure of a Co(OH)2 Electrode. J. Power Sources 2014, 248, 762−768. (34) Wu, M. − S.; Hsieh, H. − H. Nickel Oxide/Hydroxide Nanoplatelets Synthesized by Chemical Precipitation for Electrochemical Capacitors. Electrochim. Acta 2008, 53, 3427−3435. (35) Jiang, Z. − J.; Jiang, Z. Interaction Induced High Catalytic Activities of CoO Nanoparticles Grown on Nitrogen-Doped Hollow Graphene Microspheres for Oxygen Reduction and Evolution Reactions. Sci. Rep. 2016, 6, 27081. (36) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. M. Nano-Sized Transition-Metaloxides as Negative-Electrode Materials for Lithium-Ion Batteries. Nature 2000, 407, 496−499. (37) Grugeon, S.; Laruelle, S.; Herrera-Urbina, R.; Dupont, L.; Poizot, P.; Tarascon, J. M. Particle Size Effects on the Electrochemical

REFERENCES

(1) Mai, Y. J.; Tu, J. P.; Xia, X. H.; Gu, C. D.; Wang, X. L. Co-Doped NiO Nanoflake Arrays toward Superior Anode Materials for Lithium Ion Batteries. J. Power Sources 2011, 196, 6388−6393. (2) Wang, Y. F.; Zhang, L. J. Simple Synthesis of CoO-NiO-C Anode Materials for Lithium-Ion Batteries and Investigation on Its Electrochemical Performance. J. Power Sources 2012, 209, 20−29. (3) Wu, J. B.; Guo, R. Q.; Huang, X. H.; Lin, Y. Ternary Core/Shell Structure of Co3O4/NiO/C Nanowire Arrays as High-Performance Anode Material for Li-Ion Battery. J. Power Sources 2014, 248, 115− 121. (4) Thi, T. V.; Rai, A. K.; Gim, J.; Kim, J. High Performance of CoDoped NiO Nanoparticle Anode Material for Rechargeable Lithium Ion Batteries. J. Power Sources 2015, 292, 23−30. (5) Huang, Z. − D.; Zhang, K.; Zhang, T. − T.; Li, X.; Liu, R. − Q.; Feng, X. − M.; Li, Y.; Lin, X. − J.; He, Y. − B.; Yang, X. − S.; et al. Hierarchical Dispersed Multi-Phase Nickel Cobalt Oxide Mesoporous Thorn Microspheres as Superior Rate Anode Materials for Lithium Ion Batteries. J. Mater. Chem. A 2015, 3, 20886−20891. (6) Quan, W.; Tang, Z.; Hong, Y.; Wang, S.; Zhang, Z. Hydroxyl Compensation Effects on the Cycle Stability of Nickel-Cobalt Layered Double Hydroxides Synthesized via Solvothermal Method. Electrochim. Acta 2015, 182, 445−451. (7) Law, H. H.; Sapjeta, J. Effect of Cobalt on Fibrous NickelHydroxide Electrodes. J. Electrochem. Soc. 1989, 136, 1603−1606. (8) Sjovall, R. The Effect of Co Addition on the Positive Active Material in Ni-Cd Pocket-Plate Batteries. J. Power Sources 2000, 90, 153−155. (9) Huang, X. H.; Tu, J. P.; Zhang, B.; Li, Y.; Yuan, Y. F.; Wu, H. M.; Zhang, C. Q. Electrochemical Properties of NiO-Ni Nanocomposite as Anode Material for Lithium Ion Batteries. J. Power Sources 2006, 161, 541−544. (10) Xiang, J. Y.; Tu, J. P.; Yuan, Y. F.; Wang, X. L.; Huang, X. H.; Zeng, Z. Y. Electrochemical Investigation on Nanoflower-Like CuO/ Ni Composite Film as Anode for Lithium Ion Batteries. Electrochim. Acta 2009, 54, 1160−1165. (11) Pan, Q. M.; Qin, L. M.; Liu, J.; Wang, H. B. Flower-Like ZnONiO-C Films with High Reversible Capacity and Rate Capability for Lithium-Ion Batteries. Electrochim. Acta 2010, 55, 5780−5785. (12) Zhang, Y.; Zhuo, Q.; Lv, X.; Ma, Y.; Zhong, J.; Sun, X. NiOCo3O4 Nanoplate Composite as Efficient Anode in Li-Ion Battery. Electrochim. Acta 2015, 178, 590−596. (13) Xu, W.; Mu, B.; Wang, A. Three-Dimensional Hollow Microtubular Carbonized Kapok Fiber/Cobalt-Nickel Binary Oxide Composites for High-Performance Electrode Materials of Supercapacitors. Electrochim. Acta 2017, 224, 113−124. (14) Ding, C.; Yan, D.; Zhao, Y.; Zhao, Y.; Zhou, H.; Li, J.; Jin, H. A Bubble-Template Approach for Assembling Nickel-Cobalt Oxide Hollow Microspheres with an Enhanced Electrochemical Performance as an Anode for Lithium Ion Batteries. Phys. Chem. Chem. Phys. 2016, 18, 25879−25886. (15) Wei, Y.; Yan, F.; Tang, X.; Luo, Y.; Zhang, M.; Wei, W.; Chen, L. Solvent-Controlled Synthesis of NiO-CoO/Carbon Fiber Nanobrushes with Different Densities and Their Excellent Properties for Lithium Ion Storage. ACS Appl. Mater. Interfaces 2015, 7, 21703− 21711. (16) Lai, C. − C.; Lo, C. − T. Effect of Temperature on Morphology and Electrochemical Capacitive Properties of Electrospun Carbon Nanofibers and Nickel Hydroxide Composites. Electrochim. Acta 2015, 174, 806−814. (17) Oya, A.; Marsh, H. Phenomena of Catalytic Graphitization. J. Mater. Sci. 1982, 17, 309−322. (18) Zhou, H. S.; Zhu, S. M.; Hibino, M.; Honma, I.; Ichihara, M. Lithium Storage in Ordered Mesoporous Carbon (CMK-3) with High Reversible Specific Energy Capacity and Good Cycling Performance. Adv. Mater. 2003, 15, 2107−2111. (19) Zhong, J. H.; Wang, A. L.; Li, G. R.; Wang, J. W.; Ou, Y. N.; Tong, Y. X. Co3O4/Ni(OH)2 Composite Mesoporous Nanosheet L

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C Performance of Copper Oxides toward Lithium. J. Electrochem. Soc. 2001, 148, A285−A292. (38) Hu, Y. − Y.; Liu, Z.; Nam, K. − W.; Borkiewicz, O. J.; Cheng, J.; Hua, X.; Dunstan, M. T.; Yu, X.; Wiaderek, K. M.; Du, L. − S.; et al. Origin of Additional Capacities in Metal Oxide Lithium-Ion Battery Electrodes. Nat. Mater. 2013, 12, 1130−1136. (39) Zhu, Y.; Cao, C. Remarkable Electrochemical Lithium Storage Behaviour of Two-Dimensional Ultrathin α-Ni(OH)2 Nanosheets. RSC Adv. 2015, 5, 83757−83763. (40) Gachot, G.; Grugeon, S.; Armand, M.; Pilard, S.; Guenot, P.; Tarascon, J. − M.; Laruelle, S. Deciphering the Multi-Step Degradation Mechanisms of Carbonate-Based Electrolyte in Li Batteries. J. Power Sources 2008, 178, 409−421. (41) Suresh Kumar, P.; Sahay, R.; Aravindan, V.; Sundaramurthy, J.; Ling, W. C.; Thavasi, V.; Mhaisalkar, S. G.; Madhavi, S.; Ramakrishna, S. Free-Standing Electrospun Carbon Nanofibers − A High Performance Anode Material for Lithium-Ion Batteries. J. Phys. D: Appl. Phys. 2012, 45, 265302. (42) Huang, X.; Wu, J.; Guo, R.; Lin, Y.; Zhang, P. Aligned NickelCobalt Oxide Nanosheet Arrays for Lithium Ion Battery Applications. Int. J. Hydrogen Energy 2014, 39, 21399−21404. (43) Wang, C.; Appleby, A. J.; Little, F. E. Irreversible Capacities of Graphite Anode for Lithium-Ion Batteries. J. Electroanal. Chem. 2002, 519, 9−17. (44) Subramanian, V.; Zhu, H.; Wei, B. High Rate Reversibility Anode Materials of Lithium Batteries from Vapor-Grown Carbon Nanofibers. J. Phys. Chem. B 2006, 110, 7178−7183. (45) Suresh Kumar, P.; Sahay, R.; Aravindan, V.; Sundaramurthy, J.; Ling, W. C.; Thavasi, V.; Mhaisalkar, S. G.; Madhavi, S.; Ramakrishna, S. Free-Standing Electrospun Carbon Nanofibres−A High Performance Anode Material for Lithium-Ion Batteries. J. Phys. D: Appl. Phys. 2012, 45, 265302. (46) Kim, J. S.; Park, Y. T. Characteristics of Surface Films Formed at a Mesocarbon Microbead Electrode in a Li-Ion Battery. J. Power Sources 2000, 91, 172−176. (47) Jin, S. L.; Deng, H. G.; Long, D. H.; Liu, X. J.; Zhan, L.; Liang, X. Y.; Qiao, W. M.; Ling, L. C. Facile Synthesis of Hierarchically Structured Fe3O4/Carbon Micro-flowers and Their Application to Lithium-Ion Battery Anodes. J. Power Sources 2011, 196, 3887−3893. (48) Li, F.; Zou, Q. − Q.; Xia, Y. − Y. CoO-Loaded Graphitable Carbon Hollow Spheres as Anode Materials for Lithium-Ion Battery. J. Power Sources 2008, 177, 546−552. (49) Zhang, J.; Shi, Z.; Wang, C. Effect of Pre-Lithiation Degrees of Mesocarbon Microbeads Anode on the Electrochemical Performance of Lithium-Ion Capacitors. Electrochim. Acta 2014, 125, 22−28. (50) Sun, H.; Del Rio Castillo, A. E.; Monaco, S.; Capasso, A.; Ansaldo, A.; Prato, M.; Dinh, D. A.; Pellegrini, V.; Scrosati, B.; Manna, L.; Bonaccorso, F. Binder-Free Graphene as An Advanced Anode for Lithium Batteries. J. Mater. Chem. A 2016, 4, 6886−6895. (51) Cao, K.; Jiao, L.; Liu, H.; Liu, Y.; Wang, Y.; Guo, Z.; Yuan, H. Lithium-ion Batteries: 3D Hierarchical Porous α-Fe2O3 Nanosheets for High-Performance Lithium-Ion Batteries. Adv. Energy Mater. 2015, 5, 4646−4652. (52) Wang, J.; Wang, G.; Wang, H. Flexible Free-Standing Fe2O3/ Graphene/Carbon Nanotubes Hybrid Films as Anode Materials for High Performance Lithium-Ion Batteries. Electrochim. Electrochim. Acta 2015, 182, 192−201. (53) Hassoun, J.; Bonaccorso, F.; Agostini, M.; Angelucci, M.; Betti, M. G.; Cingolani, R.; Gemmi, M.; Mariani, C.; Panero, S.; Pellegrini, V.; Scrosati, B. An Advanced Lithium-Ion Battery Based on A Graphene Anode and A Lithium Iron Phosphate Cathode. Nano Lett. 2014, 14, 4901−4906. (54) Sivakkumar, S.; Pandolfo, A. Evaluation of Lithium-Ion Capacities Assembled with Pre-Lithiated Graphite Anode and Activated Carbon Cathode. Electrochim. Acta 2012, 65, 280−287. (55) Huang, X. H.; Tu, J. P.; Zhang, C. Q.; Chen, X. T.; Yuan, Y. F.; Wu, H. M. Spherical NiO-C Composite for Anode Material of Lithium Ion Batteries. Electrochim. Acta 2007, 52, 4177−4181.

(56) Martins, P. R.; Parussulo, A. L. A.; Toma, S. H.; Rocha, M. A.; Toma, H. E.; Araki, K. Highly Stabilized Alpha-NiCo(OH) 2 Nanomaterials for High Performance Device Application. J. Power Sources 2012, 218, 1−4. (57) Wu, J.; Lau, W. − M.; Geng, D. − S. Recent Progress in CobaltBased Compounds as High-Performance Anode Materials for Lithium Ion Batteries. Rare Met. 2017, 36, 307−320.

M

DOI: 10.1021/acs.jpcc.7b07412 J. Phys. Chem. C XXXX, XXX, XXX−XXX