Understanding the Effect of Different Polymeric Surfactants on

Mar 3, 2015 - Phone/Fax: +86-25-85891027., *(Z.D.) E-mail: [email protected]. ... The Si/RGO anodes made from these composites act as ideal model...
113 downloads 6 Views 6MB Size
Article pubs.acs.org/JPCC

Understanding the Effect of Different Polymeric Surfactants on Enhancing the Silicon/Reduced Graphene Oxide Anode Performance Xia Liu, Yichen Du, Lingyun Hu, Xiaosi Zhou,* Yafei Li, Zhihui Dai,* and Jianchun Bao Jiangsu Key Laboratory of Biofunctional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, P. R. China S Supporting Information *

ABSTRACT: Silicon-based lithium-ion battery anodes have brought encouraging results to the current state-of-the-art battery technologies due to their high theoretical capacity, but their large-scale application has been hampered by a large volume change (>300%) of silicon upon lithium insertion and extraction, which leads to severe electrode pulverization and capacity degradation. Polymeric surfactants directing the combination of silicon nanoparticles and reduced graphene oxide have attracted great interest as promising choices for accommodating the huge volume variation of silicon. However, the influence of different polymeric surfactants on improving the electrochemical performance of silicon/reduced graphene oxide (Si/RGO) anodes remains unclear because of the different structural configurations of polymeric surfactants. Here, we systematically study the effect of different polymeric surfactants on enhancing the Si/RGO anode performance. Three of the most well-known polymeric surfactants, poly(sodium 4-styrenesulfonate) (PSS), poly(diallydimethylammonium chloride) (PDDA), and polyvinylpyrrolidone (PVP), were used to direct the combination of silicon nanoparticles and RGO through van der Waals interaction. The Si/RGO anodes made from these composites act as ideal models to investigate and compare how the van der Waals forces between polymeric surfactants and GO affect the final silicon anode performance from both experimental observations and theoretical simulations. We found that the capability of these three surfactants in enhancing long-term cycling stability and high-rate performance of the Si/RGO anodes decreased in the order of PVP > PDDA > PSS. on the use of the structure-related properties of silicon.36−45 Ever since the successful demonstration of enhancing Li-ion battery anode performance using silicon nanowires,36 various nanostructured silicon materials, such as Si nanospheres,37 nanotubes,38 nanoparticles,39−41 and porous structure,42−45 have been fabricated and studied as active materials to accommodate the large volume changes. Another effective strategy to alleviate the so-called volume expansion issue is to encapsulate Si by electrically conductive materials.46−54 Such conductive coatings not only buffer the volume changes but also enhance the electrical conductivity of silicon. Hydrothermal growth of polysaccharide and subsequent carbonization,46 conformal generation of a carbon layer via chemical vapor deposition (CVD),47 and electrospinning of the onedimensional carbon network are well-established approaches for the formation of elastic and conductive carbon coatings.48−50 Recently, reduced graphene oxide (RGO) has become one of the most appealing conductive coatings for silicon anodes due to its unique properties, including high

1. INTRODUCTION Lithium-ion batteries (LIBs) presently dominate the portable electronics market due to their long cycle life and high energy and power density.1−13 However, currently commercial LIBs based on a graphite anode with a theoretical specific capacity of 372 mA h g−1 cannot meet the ever-growing energy demand of modern society, especially the requirement of electric vehicles and grid storage for intermittent power sources. Among all the high-capacity anode materials for next-generation LIBs, silicon is especially attractive because of its high theoretical specific capacity (4200 mA h g−1), which is more than 10 times higher than that of graphite.14−16 Moreover, silicon possesses other merits, such as low potential of Li insertion/extraction (300%) during Li uptake/release processes, leading to pulverization of Si particles, loss of electrical contact, and continuous formation of a thick solid electrolyte interphase (SEI) on Si surfaces.29−35 To tackle this problem, two main strategies have been proposed and realized with focus on controlling the enormous volume expansion of silicon anodes. The first strategy is based © 2015 American Chemical Society

Received: December 5, 2014 Revised: February 13, 2015 Published: March 3, 2015 5848

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C electrical conductivity, superior mechanical flexibility, high specific surface area, and excellent chemical stability.55−58 Poly(diallyldimethylammonium chloride) (PDDA) and NH2-terminated molecules have been employed to realize the effective encapsulation of silicon nanostructures by RGO. Notable examples include the work by Ji et al., who reported the use of PDDA for wrapping Si nanoparticles by graphene, which showed stable cycle life at high current density (a capacity of 370 mA h g−1 at 400 mA g−1 after 100 cycles);9 Chen and Sun reported a core−shell structured Si@NH2/GO anode, which delivered a reversible capacity of 1000 mA h g−1 at 420 mA g−1 after 400 cycles.34 Our report demonstrated that silicon nanoparticles encapsulated by RGO through electrostatic attraction exhibited a capacity of 1205 mA h g−1 at a current density of 100 mA g−1 after 150 cycles, but the usage of a low concentration of PDDA-modified silicon nanoparticles and GO aqueous suspension restricts its wide-scale implementation.39 Although electrostatic interaction and hydrogen bonding were investigated, no attempt has been made to systematically study the effect of weak interactions on the electrochemical performance of the Si/RGO anodes due to the varied structural configurations of Si/RGO composites used in previous works. Thus, it would be favorable to have a unified nanostructure as a desirable platform to investigate the influence of van der Waals forces on the Si/RGO electrode from both fundamental and practical perspectives. Recently, our group reported the preparation of the Si/RGO composite through freeze-drying of silicon nanoparticles and graphene oxide aqueous suspension and subsequent thermal reduction as a Li-ion battery anode, which could improve the cycling performance due to the flexible RGO sheets to accommodate the silicon volume changes.52 However, the cycling stability is not satisfactory owing to the presence of aggregated silicon nanoparticles in the composite, which often causes the loss of electrical contact between silicon nanoparticles and the conductive framework during charge/ discharge processes. Herein, we demonstrate stable and highperformance Si/RGO anodes made from polymeric surfactantmodified silicon nanoparticles. Poly(sodium 4-styrenesulfonate) (PSS), PDDA, and polyvinylpyrrolidone (PVP), three of the most well-known polymeric surfactants, were first adsorbed, respectively, onto silicon nanoparticles. Then PDDA-, PVP-, and PSS-coated silicon nanoparticles were assembled with graphene oxide through van der Waals interactions. After heat treatment under argon atmosphere at 700 °C, PSS-, PDDA-, and PVP-directed Si/RGO composites (denoted as S-Si/RGO, D-Si/RGO, and V−Si/RGO, respectively) were finally obtained. Since all of these polymeric surfactant-directed Si/ RGO composites in our research have a similar morphology and structure except using different surfactants, they can act as desired objects to study the effect of polymeric surfactants on improving the Si/RGO anode performance. We also performed ab initio simulations to reveal the difference of adsorption energy between polymeric surfactants and GO in improving the cycling stability of Si/RGO. Comparing the electrochemical performances of the S-Si/RGO, D-Si/RGO, and V-Si/RGO electrodes, we found that the ability of these three polymeric surfactants to achieve long cycle life and high-rate performance decreased in the order of PVP > PDDA > PSS. After 300 cycles at a current density of 0.5 A g−1, the V-Si/RGO electrode still shows high reversible capacity of 1054 mA h g−1. Furthermore, even at high current density of 5 A g−1, a capacity of 242 mA h g−1 can be obtained for the V-Si/RGO electrode.

2. EXPERIMENTAL SECTION 2.1. Preparation of Polymeric Surfactant-Directed Si/ RGO Composites. Graphene oxide was first fabricated from natural graphite flakes through a modified Hummer’s method.59 In a typical synthesis of V-Si/RGO, 150 mg of silicon nanoparticles (>99%, Alfa Aesar) and 150 mg of polyvinylpyrrolidone (PVP; molecular weight of ∼55 000; Aldrich) were dispersed in 15 mL of deionized (DI) water by sonication in a water bath (KQ3200DE, 40 kHz). The excess PVP was removed by repeated centrifugation/wash/redispersion cycles. Then, the resulting PVP-coated silicon nanoparticles were dispersed in 15 mL of DI water, followed by addition into 15 mL of 2.0 mg mL−1 GO aqueous dispersion under sonication. The resultant mixture was frozen with liquid nitrogen and freeze-dried for 2 days. An amount of 180 mg of the freezedried powder was placed in a crucible in a tube furnace, heated to 700 °C, and kept at that temperature for 2 h under argon atmosphere with a heating rate of 10 °C min−1 to obtain V-Si/ RGO. The as-prepared V-Si/RGO was last immersed in 10% HF water/ethanol solution to remove the SiOx surface for the following experiments. For S-Si/RGO and D-Si/RGO, PVP was replaced by poly(sodium 4-styrenesulfonate) (PSS; molecular weight of 70 000; Sigma-Aldrich) and poly(diallydimethylammonium chloride) (PDDA; molecular weight of 100 000−200 000; Sigma-Aldrich), respectively. 2.2. Materials Characterization. Scanning electron microscopy (SEM) was conducted on a JEOL JSM-7600F scanning electron microscope operated at 10 kV. Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) observations were carried out on a JEOL JEM2100F transmission electron microscope operated at 200 kV. Xray photoelectron spectroscopy (XPS) measurements were recorded on an ESCALab250Xi electron spectrometer from VG Scientific using 300 W Al Kα radiation. Nitrogen adsorption and desorption isotherms at 77.3 K were measured on an ASAP 2050 surface area-pore size analyzer. Thermogravimetric analysis (TGA) was performed on a NETZSCH STA 449 F3 under air flow with a heating rate of 10 °C min−1 from room temperature to 800 °C. Auger electron spectroscopy (AES) measurements were carried out on a ULVAC-PHI PHI-700 scanning Auger microscope at 5 keV, and Ar ions were used for step by step ion sputtering of the surface layer. 2.3. Electrochemical Measurements. Electrochemical experiments were performed using CR2032 coin cells. To fabricate working electrodes, S-Si/RGO, D-Si/RGO, or V-Si/ RGO was mixed with Super-P carbon black and carboxymethyl cellulose sodium with a weight ratio of 70:20:10 in water using a mortar and pestle. The resulting slurry was pasted onto pure Cu foil (99.9%, Goodfellow) and then dried in a vacuum oven at 40 °C for 12 h. The mass loading of active material was 1.0− 1.5 mg cm−2. The electrolyte solution for all tests was 1 M LiPF6 in ethylene carbonate/diethyl carbonate/vinylene carbonate (1:1:0.04 v/v/v). Glass fiber mats (GF/D) from Whatman and pure lithium metal foil were utilized as separators and counter electrodes, respectively. The coin cells were assembled in an argon-filled glovebox (H2O, O2 < 0.1 ppm, MBraun, Germany). The charge and discharge measurements of the batteries were carried out on a Land CT2001A multichannel battery testing system in the fixed voltage range of 0.005−1.0 V vs Li+/Li at room temperature. 5849

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C

3. RESULTS AND DISCUSSION Figure 1 schematically illustrates the synthetic procedures for the S-Si/RGO, D-Si/RGO, and V-Si/RGO composites. In a

Figure 1. Schematic illustration for the fabrication processes of polymeric surfactant-directed Si/RGO composites.

typical synthesis, PSS-, PDDA-, or PVP-coated silicon nanoparticles were fabricated through van der Waals interactions by dispersing the silicon nanoparticles in PDDA, PVP, or PSS aqueous solutions. After the removal of unadsorbed polymeric surfactants by repeated centrifugation/wash/redispersion cycles, the PDDA-, PVP-, or PSS-coated silicon nanoparticles were dispersed in deionized water and then encapsulated by graphene oxide sheets via van der Waals forces, which leads to obvious precipitations. Finally, the resulting precipitations were freeze-dried and subsequently pyrolyzed under an argon atmosphere at 700 °C for 2 h, resulting in the formation of S-Si/RGO, D-Si/RGO, and V-Si/RGO, respectively. To clearly reveal the presence of PVP on the surface of silicon nanoparticles, we used XPS to analyze PVP-coated silicon nanoparticles, showing the content of N atoms on the surface of the sample is approximately 4.5% (Supporting Information, Figure S1c). The existence of PSS and PDDA on the surface of PSS- and PDDA-coated silicon nanoparticles was also verified by the XPS, which displays distinct Na and N peaks (Supporting Information, Figure S1a,b). After assembling with graphene oxide sheets and subsequent pyrolysis, the morphology and structure of the as-obtained S-Si/RGO, D-Si/ RGO, and V-Si/RGO were characterized by SEM and TEM. SEM and TEM images show that a large number of silicon nanoparticles are aggregated and exposed on the surface of RGO in the samples of S-Si/RGO and D-Si/RGO (Figure 2a− d). In contrast, the sample of V-Si/RGO shows that most of the silicon nanoparticles are well coated by RGO sheets with only a few nanoparticles exposed on the surface (Figure 2e). In addition, the TEM image (Figure 2f) reveals that the silicon nanoparticles are uniformly distributed between the RGO sheets in the sample of V-Si/RGO. The mixture of silicon nanoparticles and RGO should lead to a higher surface area if they exist separately.60 The Brunauer−Emmett−Teller (BET) surface area decreases from 59.2 m2 g−1 in S-Si/RGO to 55.3 m2 g−1 in D-Si/RGO and 47.8 m2 g−1 in V-Si/RGO (Supporting Information, Figure S2), while the RGO contents in these samples are similar (Supporting Information, Figure

Figure 2. (a, b) SEM and TEM images of S-Si/RGO. (c, d) SEM and TEM images of D-Si/RGO. (e, f) SEM and TEM images of V-Si/ RGO.

S3). This suggests that the RGO nanosheets have better contact with silicon nanoparticles in V-Si/RGO than in S-Si/ RGO or D-Si/RGO. To find out whether the silicon nanoparticles are well wrapped by RGO nanosheets, the surface composition of S-Si/RGO, D-Si/RGO, and V-Si/RGO was analyzed by AES. The AES survey scans show that no apparent Si KLL Auger signal can be identified in the spectrum of V-Si/RGO, but two obvious peaks can be observed for S-Si/ RGO and D-Si/RGO, suggesting that the silicon nanoparticles in the V-Si/RGO composite are better coated by RGO (Supporting Information, Figure S4). Figure 3a shows the first charge/discharge profiles of S-Si/ RGO, D-Si/RGO, and V-Si/RGO. The onset slopes at 0.8 V are attributed to the formation of SEIs. The discharge plateau centered at around 0.1 V corresponds to the generation of LixSi (0 ≤ x ≤ 4.4) alloys.36,61 The charge plateau located at approximately 0.5 V relates to the phase transition from LixSi to amorphous Si.62 Figure 3b displays the cycling performances of the S-Si/RGO, D-Si/RGO, and V-Si/RGO composites tested at a current density of 0.5 A g−1 for 300 cycles. It is noted that the specific capacities are calculated based on the mass of S-Si/ RGO, D-Si/RGO, and V-Si/RGO, respectively. Under 0.5 A g−1, S-Si/RGO, D-Si/RGO, and V-Si/RGO deliver initial charge capacities of 2536, 2235, and 2155 mA h g−1 and discharge capacities of 3295, 2970, and 3175 mA h g−1, 5850

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C

(6%). After 300 cycles, reversible capacities of 104, 483, and 1054 mA h g−1 are still retained for S-Si/RGO, D-Si/RGO, and V-Si/RGO, respectively, corresponding to capacity retentions of 4%, 22%, and 49% of their initial capacities. Additionally, the average Coulombic efficiencies of the three composites for 300 cycles are 99.2% (S-Si/RGO), 99.2% (D-Si/RGO), and 99.5% (V-Si/RGO) (Supporting Information, Figure S5). The second step for making S-Si/RGO, D-Si/RGO, or V-Si/RGO is identical, suggesting that van der Waals forces are effective during the first step between graphene oxide and PSS, PDDA, or PVP-coated silicon nanoparticles, which results in different graphene oxide coating on the silicon surface, and must play a crucial role in the final electrochemical performance of S-Si/ RGO, D-Si/RGO, or V-Si/RGO. To elucidate the weak interactions between graphene oxide and polymeric surfactants (PSS, PDDA, or PVP), we performed ab initio simulations using density functional theory. Figure 4a displays the chemical structures of PSS, PDDA, and PVP, in which n indicates the number of repeat units in these surfactants. Generally, the heteroatoms with lone electron pairs, including sulfur, nitrogen, and oxygen atoms, are prone to form hydrogen bonding with hydroxyl and carboxylic groups of graphene oxide. For simplicity, we used the repeat unit of each polymeric surfactant as the model to calculate the adsorption energy for demonstrating the interaction between these surfactants and graphene oxide. This simulation can provide a qualitative explanation on the effect of adsorption strength on the cycling stability of the Si/RGO anode. The results are shown in Figure 3b. As to PVP, we can observe that both nitrogen and oxygen atoms form hydrogen bonds with the hydrogen atom in functional groups of graphene oxide. This very stable configuration shows a high adsorption energy of 1.58 eV. In comparison, both PSS and PDDA are found to have weaker interaction with GO because less hydrogen bonds can be generated between heteroatoms (sulfur atoms of PSS and nitrogen atoms of PDDA) and graphene oxide. The respective binding energies are much lower, 0.70 eV for PSS and 0.92 eV for PDDA. This stronger adsorption affinity of PVP with graphene oxide is very beneficial to bind silicon nanoparticles with graphene oxide and thus consequently form a robust Si/

Figure 3. (a) Galvanostatic charge−discharge profiles for the first cycles of S-Si/RGO, D-Si/RGO, and V-Si/RGO at a current density of 0.5 A g−1. (b) Cycling performances of S-Si/RGO, D-Si/RGO, and VSi/RGO under 0.5 A g−1. (d) Rate capabilities of S-Si/RGO, D-Si/ RGO, and V-Si/RGO.

respectively, corresponding to respective Coulombic efficiency of 77%, 75%, and 68%. The capacity loss can be ascribed to the electrolyte decomposition and irreversible insertion of the Li ion into silicon nanoparticles and RGO.63−65 After 100 cycles, reversible capacities of 143, 633, and 1434 mA h g−1 are retained for S-Si/RGO, D-Si/RGO, and V-Si/RGO, respectively. Compared to the initial cycle, the capacity retention in the 100th cycle for these electrodes decreased in the following sequence: V-Si/RGO (67%) > D-Si/RGO (28%) > S-Si/RGO

Figure 4. (a) Chemical structures of PSS, PDDA, and PVP. (b) Ab initio simulations displaying the most stable configurations and calculated adsorption energy of PSS, PDDA, or PVP with GO. The corresponding adsorption energies are shown in parentheses. The black, red, white, yellow, purple, blue, and green spheres represent C, O, H, S, Na, N, and Cl, respectively. 5851

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C

Advanced Talents of Nanjing Normal University (2014103XGQ0073), the financial support from the Priority Academic Program Development of Jiangsu Higher Education Institutions, and the Program of Jiangsu Collaborative Innovation Center of Biomedical Functional Materials.

RGO composite, giving rise to a more stable cycling performance compared to PDDA and PSS, in accordance with the results shown in Figure 3. In order to further investigate the electrode kinetics and stability, we tested the S-Si/RGO, D-Si/RGO, and V-Si/RGO electrodes at different current densities. The rate capabilities of these electrodes are demonstrated in Figure 3c. When first cycled at a current density of 0.5 A g−1 for 10 cycles, the halfbatteries made from S-Si/RGO, D-Si/RGO, and V-Si/RGO exhibit reversible capacities of 992, 1453, and 1672 mA h g−1, respectively. Successive battery testing at 1, 2, and 5 A g−1 (each step for 10 cycles) obviously shows that the rate capabilities of the composites are ordered as follows: V-Si/RGO > D-Si/RGO > S-Si/RGO, indicating that V-Si/RGO possesses the fastest reaction kinetics. This is further supported by the cycling performance of these electrodes at 0.5 A g−1 (Figure 3b). In addition, 53%, 59%, and 73% of the former capacities were, respectively, restored for the S-Si/RGO, D-Si/RGO, and V-Si/ RGO electrodes as the current density was suddenly decreased from 5 to 0.5 A g−1 (Figure 3c), indicating the excellent structural stability of the V-Si/RGO electrode.



4. CONCLUSIONS We have successfully prepared high-performance silicon anodes made from PVP-, PDDA-, and PSS-directed Si/RGO composites, which play the roles of rational models to systematically investigate the influence of different van der Waals interactions on the electrochemical performance of Sibased electrodes. Ab initio simulations were performed to disclose the interaction between these polymeric surfactants and graphene oxide for the Si/RGO composites. Comparing the lithium storage properties of the electrodes made from these Si/RGO composites, we found that the weak interactions between the heteroatoms in these surfactants with graphene oxide act critically in enhancing the cycling stability. Moreover, the RGO encapsulation on the silicon surface greatly dominates the rate capabilities of the Si/RGO electrodes. Among the three polymeric surfactants, PVP is found to be the best option that can give stable cycling performance and superior rate capability for the Si/RGO anodes.



ASSOCIATED CONTENT

S Supporting Information *

XPS spectra, nitrogen adsorption/desorption isotherms, TGA curves, AES survey scans, and Coulombic efficiencies of S-Si/ RGO, D-Si/RGO, and V-Si/RGO. This material is available free of charge via the Internet at http://pubs.acs.org.



REFERENCES

(1) Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature 2008, 451, 652−657. (2) Maier, J. Nanoionics: Ion Transport and Electrochemical Storage in Confined Systems. Nat. Mater. 2005, 4, 805−815. (3) Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 2010, 22, 587−603. (4) Xin, S.; Guo, Y.-G.; Wan, L.-J. Nanocarbon Networks for Advanced Rechargeable Lithium Batteries. Acc. Chem. Res. 2012, 45, 1759−1769. (5) Chen, J.; Cheng, F. Combination of Lightweight Elements and Nanostructured Materials for Batteries. Acc. Chem. Res. 2009, 42, 713− 723. (6) 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. (7) Zhang, G.; Wu, H. B.; Song, T.; Paik, U.; Lou, X. W. TiO2 Hollow Spheres Composed of High-crystallinity Nanocrystals Exhibit Superior Lithium Storage Properties. Angew. Chem., Int. Ed. 2014, 53, 12590−12593. (8) Chang, J.; Huang, X.; Zhou, G.; Cui, S.; Hallac, P. B.; Jiang, J.; Hurley, P. T.; Chen, J. Multilayered Si Nanoparticle/Reduced Graphene Oxide Hybrid as a High-Performance Lithium-Ion Battery Anode. Adv. Mater. 2014, 26, 758−764. (9) Ji, J.; Ji, H.; Zhang, L.; Zhao, X.; Bai, X.; Fan, X.; Zhang, F.; Ruoff, R. S. Graphene-Encapsulated Si on Ultrathin-Graphite Foam as Anode for High Capacity Lithium-Ion Batteries. Adv. Mater. 2013, 25, 4673− 4677. (10) Karki, K.; Zhu, Y.; Liu, Y.; Sun, C.-F.; Hu, L.; Wang, Y.; Wang, C.; Cumings, J. Hoop-Strong Nanotubes for Battery Electrodes. ACS Nano 2013, 7, 8295−8302. (11) Ko, M.; Chae, S.; Jeong, S.; Oh, P.; Cho, J. Elastic a-Silicon Nanoparticle Backboned Graphene Hybrid as a Self-Compacting Anode for High-Rate Lithium Ion Batteries. ACS Nano 2014, 8, 8591− 8599. (12) Guan, C.; Wang, X.; Zhang, Q.; Fan, Z.; Zhang, H.; Fan, H. J. Highly Stable and Reversible Lithium Storage in SnO2 Nanowires Surface Coated with a Uniform Hollow Shell by Atomic Layer Deposition. Nano Lett. 2014, 14, 4852−4858. (13) Zhou, X.; Dai, Z.; Liu, S.; Bao, J.; Guo, Y.-G. Ultra-Uniform SnOx/Carbon Nanohybrids toward Advanced Lithium-Ion Battery Anodes. Adv. Mater. 2014, 26, 3943−3949. (14) Yu, Y.; Gu, L.; Zhu, C.; Tsukimoto, S.; van Aken, P. A.; Maier, J. Reversible Storage of Lithium in Silver-Coated Three-Dimensional Macroporous Silicon. Adv. Mater. 2010, 22, 2247−2250. (15) Chen, D.; Mei, X.; Ji, G.; Lu, M.; Xie, J.; Lu, J.; Lee, J. Y. Reversible Lithium-Ion Storage in Silver-Treated Nanoscale Hollow Porous Silicon Particles. Angew. Chem., Int. Ed. 2012, 51, 2409−2413. (16) Liu, X. H.; Zhong, L.; Huang, S.; Mao, S. X.; Zhu, T.; Huang, J. Y. Size-Dependent Fracture of Silicon Nanoparticles During Lithiation. ACS Nano 2012, 6, 1522−1531. (17) Lee, J. K.; Smith, K. B.; Hayner, C. M.; Kung, H. H. Silicon Nanoparticles−Graphene Paper Composites for Li Ion Battery Anodes. Chem. Commun. 2010, 46, 2025−2027. (18) 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. (19) Deng, J.; Ji, H.; Yan, C.; Zhang, J.; Si, W.; Baunack, S.; Oswald, S.; Mei, Y.; Schmidt, O. G. Naturally Rolled-Up C/Si/C Trilayer

AUTHOR INFORMATION

Corresponding Authors

*(X.Z.) E-mail: [email protected]. Phone/Fax: +86-2585891027. *(Z.D.) E-mail: [email protected]. Phone/Fax: +86-2585891051. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (Grant Nos. 21175069 and 21171096), the Natural Science Foundation of Jiangsu Province of China (BK20140915), the Scientific Research Foundation for 5852

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C

spheres for Li-Ion Battery Anodes with High Structure Stability. Adv. Mater. 2014, 26, 6145−6150. (38) Hertzberg, B.; Alexeev, A.; Yushin, G. Deformations in Si-Li Anodes Upon Electrochemical Alloying in Nano-Confined Space. J. Am. Chem. Soc. 2010, 132, 8548−8549. (39) Zhou, X.; Yin, Y.-X.; Wan, L.-J.; Guo, Y.-G. Self-Assembled Nanocomposite of Silicon Nanoparticles Encapsulated in Graphene through Electrostatic Attraction for Lithium-Ion Batteries. Adv. Energy Mater. 2012, 2, 1086−1090. (40) Magasinski, A.; Dixon, P.; Hertzberg, B.; Kvit, A.; Ayala, J.; Yushin, G. High-Performance Lithium-Ion Anodes Using a Hierarchical Bottom-Up Approach. Nat. Mater. 2010, 9, 353−358. (41) Liu, N.; Lu, Z.; Zhao, J.; McDowell, M. T.; Lee, H.-W.; Zhao, W.; Cui, Y. A Pomegranate-Inspired Nanoscale Design for LargeVolume-Change Lithium Battery Anodes. Nat. Nanotechnol. 2014, 9, 187−192. (42) Hassan, F. M.; Chabot, V.; Elsayed, A. R.; Xiao, X.; Chen, Z. Engineered Si Electrode Nanoarchitecture: A Scalable Postfabrication Treatment for the Production of Next-Generation Li-Ion Batteries. Nano Lett. 2014, 14, 277−283. (43) Jia, H.; Gao, P.; Yang, J.; Wang, J.; Nuli, Y.; Yang, Z. Novel Three-Dimensional Mesoporous Silicon for High Power Lithium-Ion Battery Anode Material. Adv. Energy Mater. 2011, 1, 1036−1039. (44) Ge, M.; Lu, Y.; Ercius, P.; Rong, J.; Fang, X.; Mecklenburg, M.; Zhou, C. Large-Scale Fabrication, 3D Tomography, and Lithium-Ion Battery Application of Porous Silicon. Nano Lett. 2014, 14, 261−268. (45) Piper, D. M.; Woo, J. H.; Son, S.-B.; Kim, S. C.; Oh, K. H.; Lee, S.-H. Hierarchical Porous Framework of Si-Based Electrodes for Minimal Volumetric Expansion. Adv. Mater. 2014, 26, 3520−3525. (46) Hu, Y.-S.; Demir-Cakan, R.; Titirici, M.-M.; Müller, J.-O.; Schlögl, R.; Antonietti, M.; Maier, J. Superior Storage Performance of a Si@SiOx/C Nanocomposite as Anode Material for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2008, 47, 1645−1649. (47) Wu, H.; Chan, G.; Choi, J. W.; Ryu, I.; Yao, Y.; McDowell, M. D.; Lee, S. W.; Jackson, A.; Yang, Y.; Hu, L.; Cui, Y. Stable Cycling of Double-Walled Silicon Nanotube Battery Anodes through SolidElectrolyte Interphase Control. Nat. Nanotechnol. 2012, 7, 310−315. (48) Xu, Y.; Zhu, Y.; Han, F.; Luo, C.; Wang, C. 3D Si/C Fiber Paper Electrodes Fabricated Using a Combined Electrospray/Electrospinning Technique for Li-Ion Batteries. Adv. Energy Mater. 2014, 1400753. (49) Wang, B.; Li, X.; Qiu, T.; Luo, B.; Ning, J.; Li, J.; Zhang, X.; Liang, M.; Zhi, L. High Volumetric Capacity Silicon-Based Lithium Battery Anodes by Nanoscale System Engineering. Nano Lett. 2013, 13, 5578−5584. (50) Zhou, X.; Wan, L.-J.; Guo, Y.-G. Electrospun Silicon Nanoparticle/Porous Carbon Hybrid Nanofibers for Lithium-Ion Batteries. Small 2013, 9, 2684−2688. (51) Li, X.; Meduri, P.; Chen, X.; Qi, W.; Engelhard, M. H.; Xu, W.; Ding, F.; Xiao, J.; Wang, W.; Wang, C.; Zhang, J.-G.; Liu, J. Hollow Core−Shell Structured Porous Si−C Nanocomposites for Li-Ion Battery Anodes. J. Mater. Chem. 2012, 22, 11014−11017. (52) Zhou, X.; Yin, Y.-X.; Wan, L.-J.; Guo, Y.-G. Facile Synthesis of Silicon Nanoparticles Inserted into Graphene Sheets as Improved Anode Materials for Lithium-Ion Batteries. Chem. Commun. 2012, 48, 2198−2200. (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. (54) Xiang, H.; Zhang, K. G.; Ji, J.; Lee, Y.; Zou, C.; Chen, X.; Wu, J. Graphene/Nanosized Silicon Composites for Lithium Battery Anodes with Improved Cycling Stability. Carbon 2011, 49, 1787−1796. (55) Yang, S.; Feng, X.; Ivanovici, S.; Mullen, K. Fabrication of Graphene-Encapsulated Oxide Nanoparticles: Towards High-Performance Anode Materials for Lithium Storage. Angew. Chem., Int. Ed. 2010, 49, 8408−8411.

Nanomembranes as Stable Anodes for Lithium-Ion Batteries with Remarkable Cycling Performance. Angew. Chem., Int. Ed. 2013, 52, 2326−2330. (20) Xu, Y.; Zhu, Y.; Wang, C. Mesoporous Carbon/Silicon Composite Anodes with Enhanced Performance for Lithium-Ion Batteries. J. Mater. Chem. A 2014, 2, 9751−9757. (21) He, Y.; Yu, X.; Wang, Y.; Li, H.; Huang, X. Alumina-Coated Patterned Amorphous Silicon as the Anode for a Lithium-Ion Battery with High Coulombic Efficiency. Adv. Mater. 2011, 23, 4938−4941. (22) Koo, B.; Kim, H.; Cho, Y.; Lee, K. T.; Choi, N.-S.; Cho, J. A Highly Cross-Linked Polymeric Binder for High-Performance Silicon Negative Electrodes in Lithium Ion Batteries. Angew. Chem., Int. Ed. 2012, 51, 8762−8767. (23) Jeong, Y. K.; Kwon, T.; Lee, I.; Kim, T.-S.; Coskun, A.; Choi, J. W. Hyperbranched β-Cyclodextrin Polymer as an Effective Multidimensional Binder for Silicon Anodes in Lithium Rechargeable Batteries. Nano Lett. 2014, 14, 864−870. (24) Ma, H.; Chen, 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. (25) Chen, X.; Li, X.; Ding, F.; Xu, W.; Xiao, J.; Cao, Y.; Meduri, P.; Liu, J.; Graff, G. L.; Zhang, J.-G. Conductive Rigid Skeleton Supported Silicon as High-Performance Li-Ion Battery Anodes. Nano Lett. 2012, 12, 4124−4130. (26) Wang, C.; Wu, H.; Chen, Z.; McDowell, M. T.; Cui, Y.; Bao, Z. Self-Healing Chemistry Enables the Stable Operation of Silicon Microparticle Anodes for High-Energy Lithium-Ion Batteries. Nat. Chem. 2013, 5, 1042−1048. (27) Wang, J.; Yu, Y.; Gu, L.; Wang, C.; Tang, K.; Maier, J. Highly Reversible Lithium Storage in Si (Core)−Hollow Carbon Nanofibers (Sheath) Nanocomposites. Nanoscale 2013, 5, 2647−2650. (28) Kohandehghan, A.; Kalisvaart, P.; Kupsta, M.; Zahiri, B.; Amirkhiz, B. S.; Li, Z.; Memarzadeh, E. L.; Bendersky, L. A.; Mitlin, D. Magnesium and Magnesium-Silicide Coated Silicon Nanowire Composite Anodes for Lithium-Ion Batteries. J. Mater. Chem. A 2013, 1, 1600−1612. (29) Liu, B.; Soares, P.; Checkles, C.; Zhao, Y.; Yu, G. ThreeDimensional Hierarchical Ternary Nanostructures for High-Performance Li-Ion Battery Anodes. Nano Lett. 2013, 13, 3414−3419. (30) Nguyen, B. P. N.; Kumar, N. A.; Gaubicher, J.; Duclairoir, F.; Brousse, T.; Crosnier, O.; Dubois, L.; Bidan, G.; Guyomard, D.; Lestriez, B. Nanosilicon-Based Thick Negative Composite Electrodes for Lithium Batteries with Graphene as Conductive Additive. Adv. Energy Mater. 2013, 3, 1351−1357. (31) Yi, R.; Dai, F.; Gordin, M. L.; Sohn, H.; Wang, D. Influence of Silicon Nanoscale Building Blocks Size and Carbon Coating on the Performance of Micro-Sized Si−C Composite Li-Ion Anodes. Adv. Energy Mater. 2013, 3, 1507−1515. (32) Iwamura, S.; Nishihara, H.; Kyotani, T. Effect of Buffer Size around Nanosilicon Anode Particles for Lithium-Ion Batteries. J. Phys. Chem. C 2012, 116, 6004−6011. (33) Zhu, Y.; Liu, W.; Zhang, X.; He, J.; Chen, J.; Wang, Y.; Cao, T. Directing Silicon−Graphene Self-Assembly as a Core/Shell Anode for High-Performance Lithium-Ion Batteries. Langmuir 2013, 29, 744− 749. (34) Sun, C.; Deng, Y.; Wan, L.; Qin, X.; Chen, G. Graphene OxideImmobilized NH2-Terminated Silicon Nanoparticles by Cross-Linked Interactions for Highly Stable Silicon Negative Electrodes. ACS Appl. Mater. Interfaces 2014, 6, 11277−11285. (35) Huang, X.; Yang, J.; Mao, S.; Chang, J.; Hallac, P. B.; Fell, C. R.; Metz, B.; Jiang, J.; Hurley, P. T.; Chen, J. Controllable Synthesis of Hollow Si Anode for Long-Cycle-Life Lithium-Ion Batteries. Adv. Mater. 2014, 26, 4326−4332. (36) 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. (37) Du, F.-H.; Li, B.; Fu, W.; Xiong, Y.-J.; Wang, K.-X.; Chen, J.-S. Surface Binding of Polypyrrole on Porous Silicon Hollow Nano5853

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854

Article

The Journal of Physical Chemistry C (56) Gao, X.; Jang, J.; Nagase, S. Hydrazine and Thermal Reduction of Graphene Oxide: Reaction Mechanisms, Product Structures, and Reaction Design. J. Phys. Chem. C 2010, 114, 832−842. (57) Zhou, X.; Yin, Y.-X.; Cao, A.-M.; Wan, L.-J.; Guo, Y.-G. Efficient 3D Conducting Networks Built by Graphene Sheets and Carbon Nanoparticles for High-Performance Silicon Anode. ACS Appl. Mater. Interfaces 2012, 4, 2824−2828. (58) Xu, Y.; Zhu, X.; Zhou, X.; Liu, X.; Liu, Y.; Dai, Z.; Bao, J. Ge Nanoparticles Encapsulated in Nitrogen-Doped Reduced Graphene Oxide as an Advanced Anode Material for Lithium-Ion Batteries. J. Phys. Chem. C 2014, 118, 28502−28508. (59) Hummers, W. S., Jr.; Offeman, R. E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339−1339. (60) Song, J.; Yu, Z.; Gordin, M. L.; Hu, S.; Yi, R.; Tang, D.; Walter, T.; Regula, M.; Choi, D.; Li, X.; Manivannan, A.; Wang, D. Chemically Bonded Phosphorus/Graphene Hybrid as a High Performance Anode for Sodium-Ion Batteries. Nano Lett. 2014, 14, 6329−6335. (61) Green, M.; Fielder, E.; Scrosati, B.; Wachtler, M.; Moreno, J. S. Structured Silicon Anodes for Lithium Battery Applications. Electrochem. Solid-State Lett. 2003, 6, A75−A79. (62) Baranchugov, V.; Markevich, E.; Pollak, E.; Salitra, G.; Aurbach, D. Amorphous Silicon Thin Films as a High Capacity Anodes for LiIon Batteries in Ionic Liquid Electrolytes. Electrochem. Commun. 2007, 9, 796−800. (63) Liu, N.; Wu, H.; McDowell, M. T.; Yao, Y.; Wang, C.; Cui, Y. A Yolk-Shell Design for Stabilized and Scalable Li-Ion Battery Alloy Anodes. Nano Lett. 2012, 12, 3315−3321. (64) Esmanski, A.; Ozin, G. A. Silicon Inverse-Opal-Based Macroporous Materials as Negative Electrodes for Lithium Ion Batteries. Adv. Funct. Mater. 2009, 19, 1999−2010. (65) 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.

5854

DOI: 10.1021/jp512152f J. Phys. Chem. C 2015, 119, 5848−5854