Graphite Composite as

Jun 29, 2017 - Lithium-ion batteries (LIBs) are one of the most important/recognized rechargeable energy storage and conversion technologies in the mo...
5 downloads 28 Views 4MB Size
Research Article www.acsami.org

Nanostructured Phosphorus Doped Silicon/Graphite Composite as Anode for High-Performance Lithium-Ion Batteries Shiqiang Huang,†,‡ Ling-Zhi Cheong,§ Deyu Wang,*,† and Cai Shen*,† †

Ningbo Institute of Materials Technology & Engineering Chinese Academy of Sciences, 1219 Zhongguan Road, Zhenhai District, Ningbo, Zhejiang 315201, China ‡ University of Chinese Academy of Sciences, 19 A Yuquan Rd, Shijingshan District, Beijing 100049, P. R. China § School of Marine Science, Ningbo University, Ningbo 315211, China S Supporting Information *

ABSTRACT: Silicon as the potential anode material for lithium-ion batteries suffers from huge volume change (up to 400%) during charging/discharging processes. Poor electrical conductivity of silicon also hinders its long-term cycling performance. Herein, we report a two-step ball milling method to prepare nanostructured P-doped Si/graphite composite. Both P-doped Si and coated graphite improved the conductivity by providing significant transport channels for lithium ions and electrons. The graphite skin is able to depress the volume expansion of Si by forming a stable SEI film. The as-prepared composite anode having 50% P-doped Si and 50% graphite exhibits outstanding cyclability with a specific capacity of 883.4 mAh/g after 200 cycles at the current density of 200 mA/g. The cost-effective materials and scalable preparation method make it feasible for large-scale application of the P-doped Si/ graphite composite as anode for Li-ion batteries. KEYWORDS: lithium-ion battery, silicon anode, in situ, atomic force microscopy, ball milling



INTRODUCTION Lithium-ion batteries (LIBs) are one of the most important/ recognized rechargeable energy storage and conversion technologies in the mobile world. It is one of the best battery technologies to supply power to portable electronic devices and even electric vehicles in the future.1,2 Development of new kinds of electrode materials and structures with high energy density, long life cycle, high safety performance, and ultrafast charging and discharging rates is important for LIBs.3−9 Among various anode materials,10−17 extremely high theoretical specific capacity (4200 mAh/g, Li4.4Si) makes silicon (Si) one of the potential anode materials. In addition, Si has the advantages of low working potential, natural abundance and well-established manufacturing technologies.18−21 However, low intrinsic electrical conductivity and enormous volume expansion (up to 400%) during lithium alloy/dealloy processes limit the commercial use of Si in LIBs. Such huge volume change leads to capacity loss with cycling and is mainly due to particle pulverization, breakdown of electric conductive network, loss of contact with the conductive network or the current collector, and continual formation of stable solid electrolyte interface (SEI) on the fresh Si surfaces.22−25 In order to eliminate the aforementioned problems and acquire highly reversible Si-based anodes, a lot of designing concepts have been proposed over the past few years. Nanostructured Si materials including Si nanoparticles,26−28 Si nanospheres,29,30 Si nanowires,31−33 Si nanotubes,34−36 and Si nanosheets37,38 which have been shown to improve the © 2017 American Chemical Society

electrochemical performance of LIBs are becoming increasingly popular. Some of the materials worth mentioning include 3D porous Si anodes,39−41 metal or polymer scaffold-supported Si anodes,42,43 surface coating or core/shell structured Si anodes,44−47 Si/C composite anodes,48−50 and novel adhesion binders.51,52 Among these new structures and methods, C coated Si composite anodes seem to be one of the most promising approaches. The advantages of such composite structures are clear: (1) the carbon matrix can restrain the volume expansion of Si, (2) the carbon skins can make up low electronic conductivity of intrinsic Si and provide significant electrons and lithium ions channels, and (3) the carbon skins can minimize the direct contact of Si to the electrolyte and prevent any new SEI layer formed on fresh Si.53,54 In this paper, we prepared a nanostructured P-doped Si/ graphite (PSG) composite by a two-step ball milling process using commercially available phosphorus, silicon, and graphite as primary materials (Scheme 1). Instead of the gas deposition method,55 herein, we applied ball milling to synthesize the Pdoped Si, which is more cost-effective. Obtained P-doped Si was further ball milled with graphite to form the PSG composite. The as-prepapred PSG composite anode exhibits high reversible capacity, Coulombic efficiency and good cycle Received: March 31, 2017 Accepted: June 29, 2017 Published: June 29, 2017 23672

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

ACS Applied Materials & Interfaces



Research Article

RESULTS AND DISCUSSION Electrochemical impedances of different kinds of P-doped Si with 0.1%, 0.5%, 1%, and 2% of P were presented in Figure S1. Charge transfer resistance was found to increase with a decrease in P concentration. Nevertheless, charge transfer resistance did not vary significantly and remained almost unchanged with P concentration more than 1%. Therefore, 1% of P was chosen for further analysis. Figure 1a shows the XPS of Si 2p of P-

Scheme 1. Synthesis Process of Nanostructured P-Doped Si/ Graphite Composite

life. This low-cost material can be easily scaled-up for largescale manufacturing.



EXPERIMENTAL SECTION

Material Synthesis. Silicon powder (1−3 μm, 99.9%, Aladdin), red phosphorus (AR, 98.5%, Aladdin), and graphite power (99.99%, Aladdin) were used as starting materials. First ball milling was carried out to mix Si and P powders at a weight ratio of 99:1 in a stainless-steel vial (350 rpm, 6 h) and followed by ball milling the obtained product and graphite in another stainless-steel vial (400 rpm, 12 h). The vials were filled with argon, and planetary ball-milling (MITR-YXQM-0.4L) was carried out at normal temperature. The as-prepared P-doped Si/ graphite composites [P-doped Si/graphite at weight ratios of 3:7 (PSG37), 5:5 (PSG55), and 7:3 (PSG73)] were investigated as anodes systematically. Sample Characterization. The crystalline phase of primary Si and PSG composites were characterized by a Bruker D8 advanced diffractometer (XRD) using Cu Kα radiation (λ = l.540596 Å). Surface analysis was conducted on an X-ray photoelectron spectrometer (XPS, Kratos Axis Ultra) using 1486.6 eV Al Kα Xrays. Scanning electron microscopy (SEM, S-4800, Hitachi) and transmission electron microscopy (TEM; FEI, Tecnai F20, 200 kV) were used to characterize the morphology of Si. Operando AFM experiments were conducted by a Bruker icon AFM. The AFM was put in an argon-filled glovebox (MBRAUN) with H2O and O2 concentration both ≤0.1 ppm at room temperature. The Li−Si and Li-PSG55 cells were composed of Si or PSG55 electrode as the working electrode (WE) and Li wire as both reference and counter electrodes (RE and CE). Electrolyte solution used was 1 M LiPF6 dissolved in a mixture of dimethyl carbonate (DMC), ethyl methyl carbonate (EMC), and fluoroethylene carbonate (FEC) with a 1:1:1 (v/v/v) ratio (Shanshan Corporation). AFM images were collected simultaneously in ScanAsyst mode. Cyclic voltammetry (CV) was measured at a scanning rate of 0.6 mV/s between 2 and 0.01 V. A 2032-type coin cell system was used to carry out the electrochemical properties. The composite anodes were prepared by mixing carboxymethyl cellulose sodium (CMC), conductive carbon black (super P) and active materials at a weight ratio of 1:2:7. Slurry was formed on copper foil (current collectors) using deionized water as solvent. A vacuum oven was used to dry the obtained anodes at 100 °C for 12 h. Typical mass of active materials are ∼0.5 mg/cm2. The counter and reference electrodes were both metallic lithium, and the separator was Celgard2400 polypropylene. The electrolytes consisted of 1 M LiPF6 in a mixture of FEC/DMC/EMC (volume ratio of 1:1:1). A LAND CT2000 battery test system was used to conduct galvanostatic charge−discharge measurements of the cells in the voltage range of 0.01−2 V (vs Li/Li+). A Solartron 1470E Electrochemical Interface (Solartron Analytical, U.K.) electrochemical workstation was used to measure the cyclic voltammetry (CV) curves and electrochemical impedance spectroscopy (EIS) spectra. For the CV, the voltage sweep rate was 0.1 mV/s between potential window from 0.01 to 2 V versus Li/Li+. An AC signal of 10 mV in amplitude was used as the perturbation for EIS measurements, which were performed in the frequency range from 0.01 to 1 M Hz.

Figure 1. (a) XPS spectrum for Si 2p of P-doped Si, (b) XRD patterns of pristine Si and P-doped Si, and SEM images of (c) pristine Si and (d) P-doped Si.

doped Si, and three peaks at around 103.5, 100, and 99.4 eV corresponded to SiO2, n-type Si, and Si−Si, respectively.56 A new peak at 100 eV was found which was new in comparison to intrinsic Si (Figure S2a).56,57 XRD patterns of Si and P-doped Si were presented in Figure 1b. Both samples show peaks at 2θ values of 28.5, 47.5, 56.2, 69.4, and 76.4°, corresponding to (111), (220), (311), (400), and (331) planes of Si phase (JCPDS no. 27-1402), respectively. Thus, the ball-milling process did not destroy the Si structure, and P-doped Si was able to maintain the same crystal structure as pristine Si. Figure 1c,d shows the SEM images of pristine Si and P-doped Si. The size of pristine Si ranged from 1 to 3 μm; meanwhile the Pdoped Si was composed of small particles with the size of 100− 1000 nm. The charge capacity and corresponding Coulombic efficiency (CE) of Si and P-doped Si anodes were evaluated during 50 cycles in a potential window of 0.01−2 V versus Li/Li+ and at a current density of 200 mA/g (Figures 2a and S3). The discharge capacities of Si and P-doped Si were 3105.2 and 3044 mAh/g at the first cycle, and the corresponding CEs were 92.8 and 89.1%, respectively. After 50 cycles, the discharge capacity of Si and P-doped Si decreased to 146.5 and 587.7 mAh/g, with capacity retentions of 4.7 and 19.3%, respectively. The average capacity fading rates of Si and P-doped Si were 1.9 and 1.6%. Pdoped Si and Si achieved stable CE at 20th and 32th cycles (≥98%). Figure 2b presents the electrochemical impedance plots of Si and P-doped Si anodes before cycling. Rb represents bulk resistance of battery components (electrolyte, electrodes, and the separator); charge-transfer resistance is indicated by Rct, and Cdl represents double-layer capacitance. A straight sloping line in the low frequency region of electrochemical impedance plots is the Warburg impedance (W), which gives a combination of the interfacial diffusion resistance of Li+ ions 23673

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

Research Article

ACS Applied Materials & Interfaces

Figure 2. (a) Cycling performance and Coulombic efficiency of Si and P-doped Si electrodes at 200 mA/g. (b) Electrochemical impedance plots of Si and P-doped Si electrodes before cycling. The inset shows the equivalent circuit.

3.1 Å, corresponding to the Si (111) plane. Dark field scanning TEM (STEM) images of Si, C, and P proved that the amorphous phase covered on the Si was graphite and the homogeneous distribution of P element in Si phase. Such uniformly distributed P formed the n-type Si which enhanced the electrical conductivity. Coated graphite skin not only improved the electrical conductivity but also provided a buffer area to restrain the huge volume expansion of Si, at the same time; such a graphite skin can prevent Si from direct contacting with the electrolyte resulting in the formation of stable SEI. As shown in Figures 4, S4, and S5, the surface structural evolutions of Si and PSG55 anodes were very different. Figure 4a,c,e shows the 3D in situ AFM images of Si anode before and after the first and third cycle at a scanning rate of 0.6 mV/s between 2.0 and 0.01 V. It was clear that after the first cycle, the SEI film had covered the entire surface of Si (Figures 4c and S4b). The structure of pristine Si started to break during the second cycle. Despite that, the SEI film was still able to cover the Si surface and prevent the Si surface from contacting with the electrolyte (Figure S4c). After the third cycle, the Si structure continues to break (Figures 4e and S4d), and a massive new SEI film continued to form on the fresh surface at the forth cycle (Figure S4e). With further cycling, Si particles became isolated (Figure S7a). In contrast, the SEI film formed on the PSG55 anode was stable (Figures 4b,d,f and S5) and homogeneous (Figure S7b). We can see from line profiles in Figure S6 that the outlines of PSG55 remained almost the same. Meanwhile, the rough outline of Si was smoother after the first cycle and became inhomogeneous after the third cycle. Such a robust SEI film complemented by graphite skin truly suppressed the violent volume change and prevented the negative effects of continual growth of the thick SEI film on the Si surface during the charge/discharge process and resulting in excellent performance of LIBs.60 Figure 5a illustrates the first, second, and fiftieth charge and discharge curves of the PSG55 anode at a current density of 200 mA/g. PSG55 had a discharge and charge capacity of 2204.9 and 1427.1 mAh/g during the first cycle and a discharge and charge capacity of 1761.8 and 1640.2 mAh/g during the second cycle. The initial CE was 64.7%. The large irreversible capacity was caused by formation of the SEI film on the PSG55 anode. Nevertheless, 93.1% of the capacity can be reached at the second cycle which remained stable in the following cycles. First three CV profiles of the PSG55 anode were presented in Figure 5b. During the first CV scan, reductive peaks appeared at 1−1.5 V which corresponded to the formation of the SEI film. One cathodic peak at ∼0.01 V was attributed to the lithiation process of crystal Si with Li+, and two anodic peaks (0.32 and 0.51 V) were attributed to the delithiation process of Li−Si

across the interface of electrolyte and anode.58 The P-doped Si anode showed a much smaller semicircle than the Si anode, indicating that P-doped Si had a lower charge transfer resistance. On the other hand, P-doping suppressed the phase transition and surface morphology change of Si anode during charging/discharging processes, which can reduce the violent volume change of Si and prevent the Si anode from pulverization, besides that, such a structure might be helpful to relieve stress. All of these improved the poor cycle performance of the Si anode. Figure 3a shows the XRD patterns of PSG37, PSG55, and PSG73. The peak at 2θ values of 26.5° can be assigned to the

Figure 3. (a) XRD patterns of PSG37, PSG55, and PSG73; (b) SEM image; (c) HRTEM image; (d) STEM image of PSG55.

(002) plane of graphite phase (JCPDS no. 25-0284), which disappeared when there was less than 50% graphite. Graphite had become amorphous with less graphite additive. As for Si, three distinct peaks (28.5, 47.5, and 56.2°) assigned to (111), (220), and (311) planes existed at all three samples; thus, the crystal structure of the Si phase did not change at different content. The morphology of PSG55 is shown in Figure 3b, and the particle size had decreased to 100−300 nm following ball milling. The surfaces of these particles were rough with small particles gathered around on the bigger ones. This structure was reported to form during ball-milling by repeated cold welding/fracturing processes.59 Figure 3c,d presents more detailed morphologies and structure. High-resolution TEM (HRTEM) image demonstrated that the crystal structure dispersed in the amorphous phase, and the layer spacing was 23674

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

Research Article

ACS Applied Materials & Interfaces

Figure 4. 3D in situ AFM images of Si anode (a) before the cycle, (c) after the first cycle, and (e) after the third cycle and the PSG55 anode (b) before the cycle, (d) after the first cycle, and (f) after the third cycle at a scanning rate of 0.6 mV/s between 2.0 and 0.01 V.

Figure 5. (a) Charge and discharge profiles of the PSG55 anode for the first, second, and fiftieth cycles at a current density of 200 mA/g. (b) First three CV profiles of the PSG55 anode between 0.01 and 2 V at a scanning rate of 0.1 mV/s. (c) Cycling performance of PSG37, PSG55, and PSG55 anodes at a current density of 200 mA/g. (d) Rate performance for the PSG55 anode at different current densities.

phases which formed amorphous Si.28,61 The cathodic peak at ∼0.2 V during subsequent cycles was the characteristic peak of the lithiation process of amorphous Si with Li+. For the activation of crystal Si, the intensities of both anodic and

cathodic peaks gradually increased with further cycles.28,61 Figure 5c compares the cycling performance of the PSG37, PSG55, and PSG73 anodes. The capacity retentions of the PSG55 anode after 200 and 500 cycles were 883.4 and 600 23675

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

ACS Applied Materials & Interfaces



ACKNOWLEDGMENTS We thank the National key research and development program (Grant No. 2016YFB0100106), Ningbo 2015 graphene industrialization application development program (Development and application of high performance silicon/graphene anode materials for lithium ion batteries, Grant No. 2015S1005), and the Youth Innovation Promotion Association, CAS for financial support.

mAh/g, respectively (Figure S8), which showed the best cycling stability. The first discharge capacity and CE of the PSG37 anode were 1699.8 mAh/g and 53.6%. After the 200th cycle, the discharge capacity gradually decreased to 521.7 mAh/g and the capacity retention was only 31%. Nevertheless, it was much better than that of P free SG37, whose capacity faded to 190 mAh/g after the 200th cycle (Figure S9). As for the PSG73 anode, although it had the highest first discharge capacity and CE (2885.8 mAh/g and 80.4%), the capacity rapidly reduced to 532.4 mAh/g at the 30th cycle, and at the 200th cycle it remained at 224.3 mAh/g which cannot even compare with a commercial graphite anode (372 mAh/g). Here, we used the second discharge capacity as a reference, because the SEI film formed at the first cycle consumed a lot of Li+. Figure 5d shows the rate performance of the PSG55 anode. PSG55 achieved high discharge capacities of 1681.3, 1565, 1401, 1133.7, and 680.1 mAh/g at the current densities of 50, 100, 200, 500, and 1000 mA/g, respectively. The discharge capacity recovered when the current density changed from 1000 to 50 mA/g. Such excellent capacity recovery proved the outstanding rate performance of the PSG55 anode. It was also interesting to note that the PSG55 anode exhibited almost the same cycling performance regardless of the choice of CMC or PVDF as binder; meanwhile, both PSG37 and PSG73 anodes showed better capacity retention using PVDF and CMC binders, respectively (Figure S10).



CONCLUSION We developed nanostructured P-doped Si/graphite (PSG55) composite as anodes for LIBs by a two-step planetary ballmilling process. The as-prepared PSG composite showed good electrochemical performance. The ball milling process created n-type Si which had a lower charge transfer resistance and provided significant electrons and lithium ions transport channels in the Si bulk and created a higher capacity. A graphite skin formed after the second ball milling improved the conductivity and also prevented the direct contact between Si and the electrolyte. The graphite shell combined with the uniform SEI film can restrain the volume expansion of Si and thus maintain the structural integrity of the PSG55 anode. Cost-effective materials and the scalable preparation methods made it feasible for large-scale application of nanostructured PSG55 composite as anodes for Li-ion batteries. ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b04361.



REFERENCES

(1) Van Noorden, R. The Rechargeable Revolution: A Better Battery. Nature 2014, 507, 26−28. (2) Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the Development of Advanced Li-Ion Batteries: a review. Energy Environ. Sci. 2011, 4, 3243−3262. (3) Kang, B.; Ceder, G. Battery Materials for Ultrafast Charging and Discharging. Nature 2009, 458, 190−193. (4) Arico, A. S.; Bruce, P.; Scrosati, B.; Tarascon, J. M.; Van Schalkwijk, W. Nanostructured Materials for Advanced Energy Conversion and Storage Devices. Nat. Mater. 2005, 4, 366−377. (5) Ji, L.; Lin, Z.; Alcoutlabi, M.; Zhang, X. Recent Developments in Nanostructured Anode Materials for Rechargeable Lithium-Ion Batteries. Energy Environ. Sci. 2011, 4, 2682−2699. (6) Tang, Y.; Zhang, Y.; Deng, J.; Wei, J.; Hong Le, T.; Chandran, B. K.; Dong, Z.; Chen, Z.; Chen, X. Mechanical Force-Driven Growth of Elongated Bending TiO2-based Nanotubular Materials for Ultrafast Rechargeable Lithium Ion Batteries. Adv. Mater. 2014, 26, 6111−6118. (7) Tang, Y.; Zhang, Y.; Li, W.; Ma, B.; Chen, X. Rational Material Design for Ultrafast Rechargeable Lithium-Ion Batteries. Chem. Soc. Rev. 2015, 44, 5926−5940. (8) Tang, Y.; Zhang, Y.; Deng, J.; Qi, D.; Leow, W. R.; Wei, J.; Yin, S.; Dong, Z.; Yazami, R.; Chen, Z.; Chen, X. Unravelling the Correlation between the Aspect Ratio of Nanotubular Structures and Their Electrochemical Performance To Achieve High-Rate and LongLife Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2014, 53, 13488− 13492. (9) Tang, Y.; Zhang, Y.; Rui, X.; Qi, D.; Luo, Y.; Leow, W. R.; Chen, S.; Guo, J.; Wei, J.; Li, W.; Deng, J.; Lai, Y.; Ma, B.; Chen, X. Conductive Inks Based on a Lithium Titanate Nanotube Gel for HighRate Lithium-Ion Batteries with Customized Configuration. Adv. Mater. 2016, 28, 1567−1576. (10) Idota, Y.; Kubota, T.; Matsufuji, A.; Maekawa, Y.; Miyasaka, T. Tin-Based Amorphous Oxide: A High-Capacity Lithium-Ion-Storage Material. Science 1997, 276, 1395−1397. (11) Reddy, M. V.; Yu, T.; Sow, C.-H.; Shen, Z. X.; Lim, C. T.; Rao, G. V. S.; Chowdari, B. V. R. Alpha-Fe2O3 Nanoflakes as an Anode Material for Li-Ion Batteries. Adv. Funct. Mater. 2007, 17, 2792−2799. (12) Wang, H.; Cui, L.-F.; Yang, Y.; Casalongue, H. S.; Robinson, J. T.; Liang, Y.; Cui, Y.; Dai, H. Mn3O4-Graphene Hybrid as a HighCapacity Anode Material for Lithium Ion Batteries. J. Am. Chem. Soc. 2010, 132, 13978−13980. (13) Wu, Z.-S.; Ren, W.; Wen, L.; Gao, L.; Zhao, J.; Chen, Z.; Zhou, G.; Li, F.; Cheng, H.-M. Graphene Anchored with Co3O4 Nanoparticles as Anode of Lithium Ion Batteries with Enhanced Reversible Capacity and Cyclic Performance. ACS Nano 2010, 4, 3187−3194. (14) Chang, W.-S.; Park, C.-M.; Kim, J.-H.; Kim, Y.-U.; Jeong, G.; Sohn, H.-J. Quartz (SiO2): A New Energy Storage Anode Material for Li-Ion Batteries. Energy Environ. Sci. 2012, 5, 6895−6899. (15) Mcdowell, M. T.; Lee, S. W.; Nix, W. D.; Cui, Y. 25th anniversary article: Understanding the Lithiation of Silicon and other Alloying Anodes for Lithium-Ion Batteries. Adv. Mater. 2013, 25, 4966−4985. (16) Zhang, Y.; Rui, X.; Tang, Y.; Liu, Y.; Wei, J.; Chen, S.; Leow, W. R.; Li, W.; Liu, Y.; Deng, J.; Ma, B.; Yan, Q.; Chen, X. Wet-Chemical Processing of Phosphorus Composite Nanosheets for High-Rate and High-Capacity Lithium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1502409.





Research Article

EIS, XPS, AFM, SEM, and electrochemical performance of Si and different kinds of P-doped Si. (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Cai Shen: 0000-0001-5825-4028 Notes

The authors declare no competing financial interest. 23676

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

Research Article

ACS Applied Materials & Interfaces (17) Zhang, Y.; Tang, Y.; Li, W.; Chen, X. Nanostructured TiO2Based Anode Materials for High-Performance Rechargeable LithiumIon Batteries. ChemNanoMat 2016, 2, 764−775. (18) Ren, W.; Zhang, Z.; Wang, Y.; Tan, Q.; Zhong, Z.; Su, F. Preparation of Porous Silicon/Carbon Microspheres as High Performance Anode Materials for Lithium Ion Batteries. J. Mater. Chem. A 2015, 3, 5859−5865. (19) Kasavajjula, U.; Wang, C.; Appleby, A. J. Nano- and BulkSilicon-Based Insertion Anodes for Lithium-Ion Secondary Cells. J. Power Sources 2007, 163, 1003−1039. (20) Lin, L.; Xu, X.; Chu, C.; Majeed, M. K.; Yang, J. Mesoporous Amorphous Silicon: A Simple Synthesis of a High-Rate and Long-Life Anode Material for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2016, 55, 14063−14066. (21) Song, T.; Hu, L.; Paik, U. One-Dimensional Silicon Nanostructures for Li Ion Batteries. J. Phys. Chem. Lett. 2014, 5, 720−731. (22) Xue, L.; Xu, G.; Li, Y.; Li, S.; Fu, K.; Shi, Q.; Zhang, X. CarbonCoated Si Nanoparticles Dispersed in Carbon Nanotube Networks As Anode Material for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2013, 5, 21−25. (23) Ashuri, M.; He, Q.; Shaw, L. L. Silicon As A Potential Anode Material for Li-Ion Batteries: Where Size, Geometry and Structure Matter. Nanoscale 2016, 8, 74−103. (24) Yi, R.; Dai, F.; Gordin, M. L.; Chen, S.; Wang, D. Micro-sized Si-C Composite with Interconnected Nanoscale Building Blocks as High-Performance Anodes for Practical Application in Lithium-Ion Batteries. Adv. Energy Mater. 2013, 3, 295−300. (25) Chen, Y.; Zeng, S.; Qian, J.; Wang, Y.; Cao, Y.; Yang, H.; Ai, X. Li+-Conductive Polymer-Embedded Nano-Si Particles as Anode Material for Advanced Li-ion Batteries. ACS Appl. Mater. Interfaces 2014, 6, 3508−3512. (26) Wu, H.; Zheng, G.; Liu, N.; Carney, T. J.; Yang, Y.; Cui, Y. Engineering Empty Space between Si Nanoparticles for Lithium-Ion Battery Anodes. Nano Lett. 2012, 12, 904−909. (27) Agyeman, D. A.; Song, K.; Lee, G.-H.; Park, M.; Kang, Y.-M. Carbon-Coated Si Nanoparticles Anchored between Reduced Graphene Oxides as an Extremely Reversible Anode Material for High Energy-Density Li-Ion Battery. Adv. Energy Mater. 2016, 6, 1600904. (28) Yang, Y.; Chen, D.; Liu, B.; Zhao, J. Binder-Free Si Nanoparticle Electrode with 3D Porous Structure Prepared by Electrophoretic Deposition for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 7497−7504. (29) Yao, Y.; McDowell, M. T.; Ryu, I.; Wu, H.; Liu, N.; Hu, L.; Nix, W. D.; Cui, Y. Interconnected Silicon Hollow Nanospheres for Lithium-Ion Battery Anodes with Long Cycle Life. Nano Lett. 2011, 11, 2949−2954. (30) Ma, H.; Cheng, F.; Chen, J.; Zhao, J.; Li, C.; Tao, Z.; Liang, J. Nest-Like Silicon Nanospheres for High-Capacity Lithium Storage. Adv. Mater. 2007, 19, 4067−4070. (31) Chan, C. K.; Peng, H.; Liu, G.; McIlwrath, K.; Zhang, X. F.; Huggins, R. A.; Cui, Y. High-Performance Lithium Battery Anodes Using Silicon Nanowires. Nat. Nanotechnol. 2008, 3, 31−35. (32) Wen, Z.; Stark, J.; Saha, R.; Parker, J.; Kohl, P. A. Fabrication and Electrochemical Performance of Interconnected Silicon Nanowires Synthesized from AlCu Catalyst. J. Phys. Chem. C 2013, 117, 8604−8610. (33) Chockla, A. M.; Klavetter, K. C.; Mullins, C. B.; Korgel, B. A. Tin-Seeded Silicon Nanowires for High Capacity Li-Ion Batteries. Chem. Mater. 2012, 24, 3738−3745. (34) Song, T.; Xia, J.; Lee, J.-H.; Lee, D. H.; Kwon, M.-S.; Choi, J.M.; Wu, J.; Doo, S. K.; Chang, H.; Park, W. I.; Zang, D. S.; Kim, H.; Huang, Y.; Hwang, K.-C.; Rogers, J. A.; Paik, U. Arrays of Sealed Silicon Nanotubes As Anodes for Lithium Ion Batteries. Nano Lett. 2010, 10, 1710−1716. (35) Yoo, J.-K.; Kim, J.; Jung, Y. S.; Kang, K. Scalable Fabrication of Silicon Nanotubes and their Application to Energy Storage. Adv. Mater. 2012, 24, 5452−5456.

(36) Epur, R.; Hanumantha, P. J.; Datta, M. K.; Hong, D.; Gattu, B.; Kumta, P. N. A Simple and Scalable Approach to Hollow Silicon Nanotube (h-SiNT) Anode Architectures of Superior Electrochemical Stability and Reversible Capacity. J. Mater. Chem. A 2015, 3, 11117− 11129. (37) Kim, W.-S.; Hwa, Y.; Shin, J.-H.; Yang, M.; Sohn, H.-J.; Hong, S.-H. Scalable synthesis of silicon nanosheets from sand as an anode for Li-ion batteries. Nanoscale 2014, 6, 4297−4302. (38) Ryu, J.; Hong, D.; Choi, S.; Park, S. Synthesis of Ultrathin Si Nanosheets from Natural Clays for Lithium-Ion Battery Anodes. ACS Nano 2016, 10, 2843−2851. (39) Wada, T.; Ichitsubo, T.; Yubuta, K.; Segawa, H.; Yoshida, H.; Kato, H. Bulk-Nanoporous-Silicon Negative Electrode with Extremely High Cyclability for Lithium-Ion Batteries Prepared Using a TopDown Process. Nano Lett. 2014, 14, 4505−4510. (40) Kim, H.; Han, B.; Choo, J.; Cho, J. Three-Dimensional Porous Silicon Particles for Use in High-Performance Lithium Secondary Batteries. Angew. Chem., Int. Ed. 2008, 47, 10151−10154. (41) Deng, H.; Chu, G.; Luo, F.; Li, H.; Chen, L.; Aifantis, K. E. Si Micropyramid Patterned Anodes that Can Suppress Fracture and Solid Electrolyte Interface Formation During Electrochemical Cycling. J. Power Sources 2016, 329, 372−378. (42) Zhang, H.; Braun, P. V. Three-Dimensional Metal Scaffold Supported Bicontinuous Silicon Battery Anodes. Nano Lett. 2012, 12, 2778−2783. (43) Wu, H.; Yu, G.; Pan, L.; Liu, N.; McDowell, M. T.; Bao, Z.; Cui, Y. Stable Li-Ion Battery Anodes by In-Situ Polymerization of Conducting Hydrogel to Conformally Coat Silicon Nanoparticles. Nat. Commun. 2013, 4, 1943. (44) Luo, L.; Zhao, P.; Yang, H.; Liu, B.; Zhang, J.-G.; Cui, Y.; Yu, G.; Zhang, S.; Wang, C.-M. Surface Coating Constraint Induced SelfDischarging of Silicon Nanoparticles as Anodes for Lithium Ion Batteries. Nano Lett. 2015, 15, 7016−7022. (45) Ng, S.-H.; Wang, J.; Wexler, D.; Konstantinov, K.; Guo, Z.-P.; Liu, H.-K. Highly Reversible Lithium Storage in Spheroidal CarbonCoated Silicon Nanocomposites as Anodes for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2006, 45, 6896−6899. (46) Zhang, L.; Rajagopalan, R.; Guo, H.; Hu, X.; Dou, S.; Liu, H. A Green and Facile Way to Prepare Granadilla-Like Silicon-Based Anode Materials for Li-Ion Batteries. Adv. Funct. Mater. 2016, 26, 440−446. (47) Hwang, T. H.; Lee, Y. M.; Kong, B.-S.; Seo, J.-S.; Choi, J. W. Electrospun Core-Shell Fibers for Robust Silicon Nanoparticle-Based Lithium Ion Battery Anodes. Nano Lett. 2012, 12, 802−807. (48) Zhang, R.; Du, Y.; Li, D.; Shen, D.; Yang, J.; Guo, Z.; Liu, H. K.; Elzatahry, A. A.; Zhao, D. Highly Reversible and Large Lithium Storage in Mesoporous Si/C Nanocomposite Anodes with Silicon Nanoparticles Embedded in a Carbon Framework. Adv. Mater. 2014, 26, 6749−6755. (49) Guo, J.; Yang, Z.; Archer, L. A. Mesoporous Silicon@carbon Composites via Nanoparticle-Seeded Dispersion Polymerization and Their Application as Lithium-Ion Battery Anode Materials. J. Mater. Chem. A 2013, 1, 5709−5714. (50) Yu, J.; Zhan, H.; Wang, Y.; Zhang, Z.; Chen, H.; Li, H.; Zhong, Z.; Su, F. Graphite Microspheres Decorated with Si Particles Derived From Waste Solid of Organosilane Industry As High Capacity Anodes for Li-Ion Batteries. J. Power Sources 2013, 228, 112−119. (51) Ryou, M.-H.; Kim, J.; Lee, I.; Kim, S.; Jeong, Y. K.; Hong, S.; Ryu, J. H.; Kim, T.-S.; Park, J.-K.; Lee, H.; Choi, J. W. Mussel-Inspired Adhesive Binders for High-Performance Silicon Nanoparticle Anodes in Lithium-Ion Batteries. Adv. Mater. 2013, 25, 1571−1576. (52) Kovalenko, I.; Zdyrko, B.; Magasinski, A.; Hertzberg, B.; Milicev, Z.; Burtovyy, R.; Luzinov, I.; Yushin, G. A Major Constituent of Brown Algae for Use in High-Capacity Li-Ion Batteries. Science 2011, 334, 75−79. (53) Lee, W. J.; Hwang, T. H.; Hwang, J. O.; Kim, H. W.; Lim, J.; Jeong, H. Y.; Shim, J.; Han, T. H.; Kim, J. Y.; Choi, J. W.; Kim, S. O. N-Doped Graphitic Self-Encapsulation For High Performance Silicon Anodes in Lithium-Ion Batteries. Energy Environ. Sci. 2014, 7, 621− 626. 23677

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678

Research Article

ACS Applied Materials & Interfaces (54) Wang, B.; Li, X.; Zhang, X.; Luo, B.; Zhang, Y.; Zhi, L. ContactEngineered and Void-Involved Silicon/Carbon Nanohybrids as Lithium-Ion-Battery Anodes. Adv. Mater. 2013, 25, 3560−3565. (55) Domi, Y.; Usui, H.; Shimizu, M.; Kakimoto, Y.; Sakaguchi, H. Effect of Phosphorus-Doping on Electrochemical Performance of Silicon Negative Electrodes in Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 7125−7132. (56) Kim, J. S.; Choi, W.; Byun, D.; Lee, J. K. Electrochemical Characteristics of Phosphorus Doped Silicon for the Anode Material of Lithium Secondary Batteries. Solid State Ionics 2012, 212, 43−46. (57) Kong, M.-h.; Noh, J.-h.; Byun, D.-j.; Lee, J.-k. Electrochemical Characteristics of Phosphorus Doped Silicon and Graphite Composite for the Anode Materials of Lithium Secondary Batteries. J. Electroceram. 2009, 23, 376−381. (58) Ahmed, B.; Anjum, D. H.; Gogotsi, Y.; Alshareef, H. N. Atomic Layer Deposition of SnO 2 on MXene for Li-Ion Battery Anodes. Nano Energy 2017, 34, 249−256. (59) Gauthier, M.; Mazouzi, D.; Reyter, D.; Lestriez, B.; Moreau, P.; Guyomard, D.; Roue, L. A Low-Cost and High Performance BallMilled Si-based Negative Electrode for High-Energy Li-Ion Batteries. Energy Environ. Sci. 2013, 6, 2145−2155. (60) Luo, F.; Chu, G.; Xia, X.; Liu, B.; Zheng, J.; Li, J.; Li, H.; Gu, C.; Chen, L. Thick Solid Electrolyte Interphases Grown on Silicon Nanocone Anodes During Slow Cycling and Their Negative Effects on the Performance of Li-Ion Batteries. Nanoscale 2015, 7, 7651−7658. (61) Obrovac, M. N.; Krause, L. J. Reversible Cycling of Crystalline Silicon Powder. J. Electrochem. Soc. 2007, 154, A103−A108.

23678

DOI: 10.1021/acsami.7b04361 ACS Appl. Mater. Interfaces 2017, 9, 23672−23678