Six Thousand Electrochemical Cycles of Double-Walled Silicon Nanotube. Anodes for Lithium Ion Batteries

Similar documents
To meet the demands of future portable electronics and

2014 GCEP Report - External

Batteries: Now and Future

To meet the future demands of portable electronics, electric

2015 GCEP Report - external

Lithium-ion Batteries Based on Vertically-Aligned Carbon Nanotubes and Ionic Liquid

Graphene-Wrapped Sulfur Particles as a Rechargeable. Lithium-Sulfur-Battery Cathode Material with High Capacity and

The goal of this project is to enhance the power density and lowtemperature efficiency of solid oxide fuel cells (SOFC) manufactured by atomic layer

Supporting Information

Studying the Kinetics of Crystalline Silicon Nanoparticle Lithiation with In Situ Transmission Electron Microscopy

High capacity Li-ion battery electrode materials such as Si, Crab Shells as Sustainable Templates from Nature for Nanostructured Battery Electrodes

Supporting Information

Supporting Information

The Effect of Low-Temperature Carbon Encapsulation on Si Nanoparticles for Lithium Rechargeable Batteries

Scalable synthesis of silicon-nanolayer-embedded graphite for high-energy lithium-ion batteries

A Scalable Synthesis of Few-layer MoS2. Incorporated into Hierarchical Porous Carbon. Nanosheets for High-performance Li and Na Ion

Supporting Information

Nanoscale Interface Control of High-Quality Electrode Materials for Li-Ion Battery and Fuel Cell

Materials and Structural Design for Advanced Energy Storage Devices

Supporting Information for

Supporting Information

Supplementary Figure 1 XPS, Raman and TGA characterizations on GO and freeze-dried HGF and GF. (a) XPS survey spectra and (b) C1s spectra.

Influence of Silicon Nanoscale Building Blocks Size and Carbon Coating on the Performance of Micro-Sized Si C Composite Li-Ion Anodes

Significant Improvement of LiNi 0.8 Co 0.15 Al 0.05 O 2 Cathodes at 60 C by SiO 2 Dry Coating for Li-Ion Batteries

Si-, Ge-, Sn-Based Anode Materials for Lithium-Ion Batteries: From Structure Design to Electrochemical Performance

High Tap Density Secondary Silicon Particle. Anodes by Scalable Mechanical Pressing for

One dimensional Si/Sn - based nanowires and nanotubes for lithium-ion energy storage materials

Layered reduced graphene oxide with nanoscale interlayer gaps as a stable

Capacity fade studies of Lithium Ion cells

Thierry Djenizian. Fabrication de microbatteries Li-ion à base de nanotubes de TiO2. Department of Flexible Electronics, CMP Gardanne

Multicomponent (Mo, Ni) metal sulfide and selenide microspheres with empty nanovoids as anode materials for Na-ion batteries

Supporting Information

Highly stable and flexible Li-ion battery anodes based on TiO 2 coated

Supplementary Figure 1 a-c, The viscosity fitting curves of high-molecular-weight poly(vinyl alcohol) (HMW-PVA) (a), middle-molecular-weight

Department of Chemical Engineering, Tsinghua University, Beijing , P. R. China

Please do not adjust margins. Electronic supplementary information

Supplementary Figures

Advanced energy storage technologies are critically

Journal of Materials Chemistry A ELECTRONIC SUPPLEMENTARY INFORMATION (ESI )

High Voltage Magnesium-ion Battery Enabled by Nanocluster Mg3Bi2

Supplementary Figure 1 A schematic representation of the different reaction mechanisms

Enhanced Power Systems Through Nanotechnology

Group IV elements are promising anode materials to attain

Nanostrukturphysik (Nanostructure Physics)

Optimized Silicon Electrode Architecture, Interface, and Microgeometry for Next-Generation Lithium-Ion Batteries

High-capacity electrode materials are inherently accompanied. Interfacial Oxygen Stabilizes Composite Silicon Anodes

Supplementary Information

Supplementary Figure 1 Supplementary Figure 2

Topotactically synthesized TiO 2 nanowires as promising anode materials for high-performance lithium-ion batteries

Enhancing Sodium Ion Battery Performance by. Strongly Binding Nanostructured Sb 2 S 3 on

The Importance of Electrochemistry for the Development of Sustainable Mobility

Supporting Information

Tuning the Shell Number of Multi-Shelled Metal Oxide. Hollow Fibers for Optimized Lithium Ion Storage

An Advanced Anode Material for Sodium Ion. Batteries

Low Power Phase Change Memory via Block Copolymer Self-assembly Technology

SiOC Anode Material Derived from Poly(phenyl carbosilane) for Lithium Ion Batteries

Supplementary Figure S1. AFM image and height profile of GO. (a) AFM image

Science and Technology, Dalian University of Technology, Dalian , P. R. China b

Flexible Asymmetrical Solid-state Supercapacitors Based on Laboratory Filter Paper

High-Areal-Capacity Silicon Electrodes with Low-Cost Silicon Particles Based on Spatial Control of Self-Healing Binder

LITHIUM ION BATTERIES

A New Cathode Material for Potassium-Ion Batteries

Supporting Information. Fast Synthesis of High-Performance Graphene by Rapid Thermal Chemical Vapor Deposition

Catalytic role of ge in highly reversible GeO2/Ge/ C nanocomposite anode material for lithium batteries

Supporting Information. Metal-Organic Frameworks Mediated Synthesis of One-Dimensional Molybdenum-Based/Carbon Composites for Enhanced Lithium Storage

MATERIALS FOR SUPERCAPACITORS ELECTRODES: PREFORMANCE AND NEW TRENDS

Inexpensive Colloidal SnSb Nanoalloys as Efficient Anode Materials for Lithium- and Sodium-Ion Batteries

state expose the the positive (electrode 2; top electrode S 1

Electronic Supplementary Information

Carbon scaffold structured silicon anodes for lithium-ion batteries

Honeycomb-like Interconnected Network of Nickel Phosphide Hetero-nanoparticles

Sustainable Li/Na-Ion Batteries

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

High-Performance Silicon Battery Anodes Enabled by

Kinetically-Enhanced Polysulfide Redox Reactions by Nb2O5. Nanocrystal for High-Rate Lithium Sulfur Battery

Tailorable and Wearable Textile Devices for Solar Energy Harvesting and Simultaneous Storage

Supporting Information

Journal of Power Sources

Chemistry: The Central Science. Chapter 20: Electrochemistry

Electronic Supplementary Information. Lithium-Oxygen Batteries: Bridging Mechanistic Understanding and Battery Performance

Supplemental Information. Crumpled Graphene Balls Stabilized. Dendrite-free Lithium Metal Anodes

Lithium Sulfide/Metal Nanocomposite as a High-Capacity Cathode Prelithiation Material

Silicon is a promising high-capacity anode material for Li ion

Novel Substrate-Bound Hybrid Nanomaterials for Anode Electrodes in Lithium-Ion Batteries

High-Performance Flexible Asymmetric Supercapacitors Based on 3D. Electrodes

Supporting Information

Hierarchical MoO 2 /Mo 2 C/C Hybrid Nanowires for High-Rate and. Long-Life Anodes for Lithium-Ion Batteries. Supporting Information

Supporting Information. Co 4 N Nanosheets Assembled Mesoporous Sphere as a Matrix for Ultrahigh Sulfur Content Lithium Sulfur Batteries

Supplementary Materials for

In-Situ Fabrication of CoS and NiS Nanomaterials Anchored on. Reduced Graphene Oxide for Reversible Lithium Storage

Supplementary Figures

Supporting Information. Bi-functional Catalyst with Enhanced Activity and Cycle Stability for. Rechargeable Lithium Oxygen Batteries

5. Building Blocks I: Ferroelectric inorganic micro- and nano(shell) tubes

High-density data storage: principle

A new strategy for achieving a high performance anode for lithium ion batteries-encapsulating germanium nanoparticles in carbon nanoboxes

Fast and reversible thermoresponsive polymer switching materials for safer batteries

Low temperature anodically grown silicon dioxide films for solar cell. Nicholas E. Grant

Electronic Supplementary Information (ESI)

Raman analysis of lithium-ion battery components

Supplementary information

Transcription:

SLAC-PUB-14379 Six Thousand Electrochemical Cycles of Double-Walled Silicon Nanotube Anodes for Lithium Ion Batteries Hui Wu, 1* Gerentt Chan, 2* Jang Wook Choi, 1, 3 Ill Ryu, 1 Yan Yao, 1 Matthew T. McDowell, 1 Seok Woo Lee, 1 Ariel Jackson, 1 Liangbing Hu, 1 Yi Cui 1, 1 Department of Materials Science and Engineering, 2 Department of Chemistry, Stanford University, CA 94305, USA. 3 Current address: Graduate School of EEWS, Korea Advanced Institute of Science and Technology, Daejeon 305-701, Republic of Korea. *These authors contributed equally to this work. To whom correspondence should be addressed. Email: yicui@stanford.edu Work supported in part by US Department of Energy contract DE-AC02-76SF00515. SLAC National Accelerator Laboratory, Menlo Park, CA 94025 1

Abstract Despite remarkable progress, lithium ion batteries still need higher energy density and better cycle life for consumer electronics, electric drive vehicles and large-scale renewable energy storage applications. Silicon has recently been explored as a promising anode material for high energy batteries; however, attaining long cycle life remains a significant challenge due to materials pulverization during cycling and an unstable solid-electrolyte interphase. Here, we report double-walled silicon nanotube electrodes that can cycle over 6000 times while retaining more than 85% of the initial capacity. This excellent performance is due to the unique double-walled structure in which the outer silicon oxide wall confines the inner silicon wall to expand only inward during lithiation, resulting in a stable solid-electrolyte interphase. This structural concept is general and could be extended to other battery materials that undergo large volume changes. Among all the rechargeable battery technologies, lithium-ion batteries (LIBs) offer superior performance, and are the main power source for portable electronics. LIBs are also the most promising power source for electric vehicles and are projected to be enablers of smart grids based on renewable energy technologies (1-3). For many of these applications, energy density and cycle life stand out as two important technical parameters which need significant improvements. For example, in 2010 the U.S. Department of Energy has put forth a goal to create a LIB with double the energy density of current batteries and a cycle life of 5000 cycles with 80% capacity retention for electric vehicles, in comparison, the typical high energy batteries used in portable electronics only have a cycle life of 500-1000 cycles (4). Increasing energy density in LIBs requires developing electrode materials with higher charge capacity or higher voltage (5). Improving cycle life involves stabilizing two critical components of battery electrodes: the active electrode materials and their interface with the electrolyte the so-called solid-electrolyte interphase (SEI). 2

Recently, silicon (Si) has emerged as one of the most promising electrode materials for next-generation high energy LIBs. It offers a suitable low voltage for an anode and a high theoretical specific capacity of ~4,200 mah/g based on the formation of the Li 4.4 Si alloy, which is 10 times higher than that of conventional carbon anodes (~372 mah/g corresponding to the formation of LiC 6 ) (6-11). However, Si expands volumetrically by up to 400% upon full lithium insertion (lithiation) to form the Li 4.4 Si alloy, and it can contract significantly upon lithium extraction (delithiation) (6, 11), creating two critical challenges associated with silicon-based anode materials: degradation of the mechanical integrity of Si electrodes and the stability of SEI. The stress induced by the large volume changes causes cracking and pulverization of Si anodes, which leads to loss of electrical contact and eventual capacity fading. This was considered to be the main reason for the rapid capacity loss in early studies of Si anodes (6-11). Recently, there have been successes in addressing these materials stability issues by designing nanostructured materials, including nanowires, nanotubes, nanoporous films and Si nanoparticle/carbon composites (12-21). The strain in nanostructures can be easily relaxed without mechanical fracture due to their small size and surrounding free space. This nanostructuring strategy has greatly increased the cycle life of silicon anodes up to a few hundred cycles with 80% capacity retention (18, 19); however, it is still far from the desired cycle life of thousands of cycles. The SEI stability at the interface between Si and the liquid electrolyte is another critical factor in obtaining long cycle life; it is a very challenging issue and has not been effectively addressed for materials with large volume changes, as shown in the schematic in Fig. 1. Electrolyte decomposition occurs due to the low potential of the anode and forms a passivating SEI layer on the electrode surface during battery charging. The SEI layer is an electronic insulator but a lithium ion conductor; this results in the SEI eventually terminating growth at a certain thickness (22, 23). Even though Si nanostructures have largely overcome mechanical fracture issues, their interface with the electrolyte is not static due to repetitive volume expansion and contraction (6, 8, 21, 24). As schematically shown in Fig. 1 A and B, 3

both solid and hollow Si structures expand out towards the electrolyte upon lithiation and contract during delithiation. The SEI formed in the lithiated expanded state can be broken as the nanostructure shrinks during delithiation. This re-exposes the fresh Si surface to the electrolyte and the SEI forms again, resulting in thicker and thicker SEI upon charge/discharge cycling. This results in battery performance degradation through: a) the consumption of electrolyte and of lithium ions during continuous SEI formation, b) the electrically insulating nature of the SEI weakening electrical contact between the current collector and anode material, c) the long lithium diffusion distance through the thick SEI and d) electrode material degradation caused by mechanical stress from the thick SEI The formation of a stable SEI is critical for realizing long cycle life in Si anodes, which also holds in general for other battery electrode materials that undergo large volume changes. Here we design a novel double-walled Si nanotube (DWSiNT) anode, in which the inner wall is active Si material and the outer wall is a confining material that allows lithium ions to pass through (see the cross-sectional view in Fig. 1C). In this design, the electrolyte only contacts the outer surface and cannot enter the inner hollow space. During lithiation, lithium ions penetrate through the outer wall and react with the inner Si wall. The outer wall is mechanically rigid, so the inner Si wall expands inward into the hollow space. This inward expansion is possible because silicon is considerably softened upon significant lithium insertion (25). During delithiation, the inner surface of the Si wall shrinks back. Overall, the interface with electrolyte is mechanically constrained and remains static during both lithiation and delithiation. Only the inner surface moves back and forth, and it does not contact the electrolyte (Fig. 1C). The DWSiNT structure provides two attractive points as an anode material: first, the static outer surface allows for the development of a stable SEI; second, the inner space allows for free volume expansion of Si without mechanical breaking. We note that our proposed working mechanisms are substantially different from previous Si nanotube studies (15, 19, 21), in which no static interface between silicon and the electrolyte have been designed. To experimentally realize the DWSiNTs, we have developed a templating method, as schematically 4

shown in Fig. 2A. Polymer nanofibers (green) were first produced by electrospinning, a well-known technique for making continuous ultralong nanofibers at low-cost and large quantity (26, 27). The polymer nanofiber networks were then annealed in inert gas at 500 o C to form carbon nanofiber mats (black in Fig. 2A and Fig. S1). Using a SiH 4 chemical vapor deposition (CVD) method, a thin layer (~ 30 nm) of silicon (blue Fig. 2A) was coated onto the carbon nanofibers at 490 o C (Fig. S2). The samples were then heated in air at 500 o C for 2 hrs. Oxygen gas can diffuse through the broken ends of the nanofibers, and the carbon cores were oxidized into carbon dioxide gas and removed completely. Fig. 2B shows a scanning electron microscopy (SEM) image of the synthesized materials demonstrating that smooth and uniform hollow tubes were fabricated with full removal of their inner cores. The low magnification SEM image in Fig. 2C indicates that the synthesized tubes were continuous over long distances (28). This morphology can be attributed to two primary reasons: the extremely high aspect ratio of the template electrospun nanofiber template and the conformal coating of silicon in the CVD process. The continuous tube structures is vital for preventing electrolyte from wetting the interior and thus protect the tubes from SEI interior growth, as will be discussed later. Another advantage of this continuous network of nanotubes is to provide mechanical and electrical interconnects along the whole network, thus making it possible to directly use free-standing nanotube mats as battery anodes without any conducting additives or binders. A transmission electron microscopy (TEM) image of synthesized tubes (Fig. 2D) shows that both the inside and outside surfaces are smooth and the wall thickness is 30 nm. The selected area electron diffraction (SAED) pattern shows diffuse scattering (insert of Fig. 2D), suggesting that the synthesized nanotubes are amorphous. It is important to notice that the outside surface of the tubes, which were heated in air at 500 o C during the carbon-removal process, has been naturally oxidized, forming a thin, external coating of silicon oxide. During the oxidation of the existing carbon core, there is very little oxidation of the inner surface of the Si tubes since the carbon core is a sacrificial material which consumes oxygen first. 5

Therefore, we expect that the proposed DWSiNT structure can be formed, where the outer wall is SiO 2 and the inner wall is Si. To confirm this, Auger electron spectroscopy sputter depth profiling was performed to assess the elemental distribution along the cross section of DWSiNTs (Fig. 2E). During the initial sputtering both Si and O are present, corresponding to the outer SiO 2 wall of the nanotubes. Working through the inner silicon wall from t=1-3min, little O intensity is observed. From t=4-5min, the sputtering has bored through the first Si wall and has started on the Si wall on the other side, working outwards. The thickness of the outer SiO 2 wall was determined to be ~10 nm and the inner Si wall is ~20 nm, which is consistent with the TEM study (Figure S3). The results of the Auger electron spectroscopy depth profiling clearly show that our proposed DWSiNTs are successfully synthesized. We hypothesize that the outer SiO 2 wall can serve as a mechanical constraining layer to hold the external surface of the DWSiNTs static while forcing the Si expansion inwards into the hollow core. SiO 2 has also been known to allow lithium ions to penetrate through it so that the lithiation process can take place (12, 29). To understand the mechanical constraining effects of the outer oxide wall, we have conducted finite element simulations to model the stress evolution and deformation behavior during lithiation. In the model, the amorphous Si wall is the active material that reacts with the lithium ions. Lithium can also react with SiO 2, although its reaction is not reversible and its products, such as lithium silicates and oxides, are mechanically strong materials. For simplicity, only the mechanical properties of SiO 2 are taken into account, and the reaction between Li ions and the oxide is ignored in the model. Fig. 2 F, G and Fig. S6 show the hoop stress evolution during lithiation in Si nanotubes (SiNTs) with and without the outer oxide constraining layer. In SiNTs without the oxide wall, the outer region is in compression and the inner region is in tension during lithiation (Fig. 2F, fig. S6A). In this case, the main source of stress generation is the volumetric strain gradient. However, the addition of the oxide layer causes a strong compressive stress due to the constraining effect from the outer oxide layer. Therefore, stresses mainly result from the amount of volumetric strain itself, not from its gradient (Fig. 2G, fig. S6B). This different mechanism for stress generation is revealed by the fact that the maximum 6

stress occurs at a very early stage in SiNTs, while DWSiNTs show maximum stress at 12.5% of the fully lithiated state (Fig. S6). As a result, the oxide-free SiNTs expand outward during lithiation, while the oxide wall of DWSiNTs forces the Si to only expand inward (See Movie S1, S2. The details of the simulations are given in the supplementary information section). This mechanical constraining effect was confirmed experimentally by tracking a single nanotube before and after lithiation by TEM observation (see SOM for experimental details). As shown in Fig. 2 H and I, a DWSiNT with inner diameter (ID) of 209±2 nm and outer diameter (OD) of 294±2 nm before lithiation has ID of 156±4 nm and OD of 293±3 nm after full lithiation. Thus, only inward expansion in DWSiNTs was observed after full lithiation of silicon. We have confirmed this general observation through TEM experiments of more individual DWSiNTs (Table S1). The mechanical constraining effect in DWSiNTs affords a very exciting opportunity to control the SEI. To demonstrate this experimentally, we have prepared battery cells in which three different types of Si nanostructured electrodes are paired with lithium foil in half cells: 1) solid silicon nanowires (Fig. 3A); 2) hollow silicon nanotubes without the SiO 2 clamping layer (Fig. 3D) and 3) hollow DWSiNTs with the silicon oxide clamping layer (Fig. 3G). These Si electrodes were subjected to multiple deep battery charge/discharge cycles (1 V - 0.01 V) and then extracted for observation with SEM (200 cycles for Si nanowire and single-walled Si nanotube without the clamping layer, 2000 cycles for DWSiNTs). After 200 cycles, both the Si nanowires and Si nanotubes without the clamping layer were buried under SEI layers, indicating that the SEI had grown very thick (Fig. 3 B and E). However, the SEI in DWSiNTs is very different. Even after many more cycles (2000), the DWSiNTs only have a very thin SEI layer and individual DWSiNTs with SEI coating can be clearly resolved (Fig. 3H). The statistical analysis shows that the outer diameters of DWSiNTs only increase slightly after cycling (Fig. 3J), and the SEI layer thickness is calculated to be only ~110 nm. This thickness is similar to the thickness of SEI formed on an excellent carbon anode. To confirm that the electrolyte does not wet the inside of 7

DWSiNTs and cause SEI formation there, we have used focused ion beam to cut open the DWSiNTs after cycling. Fig. S5 shows SEM images of cycled DWSiNTs after cutting and clearly confirms that there is no SEI formation inside the hollow space of DWSiNTs. In addition, the experimental results on solid nanowires and hollow tubes with and without constraining oxide layers confirm our suggested SEI breaking/reconstruction mechanism (Fig. 1A-C). The Si material stability is also important for battery cycling, and we compared the Si morphology changes for these three types of nanostructures. We have previously developed an acid etching method to remove the SEI without damaging the Si electrode materials (see SOM for details) (29). Upon SEI removal using this method, the nanowires and nanotubes without the oxide layer can be seen to have very rough and highly porous surfaces (Fig. 3 C and F, respectively), indicating significant material displacement and morphological changes (29). However, after removing the SEI from the DWSiNTs, they still appear smooth, continuous and uniform, similar to those before cycling (Fig. 3I). This further suggests that the Si material is stable and the outer surface undergoes very little change during cycling and is static. The controlled thin SEI growth on DWSiNTs can be beneficial for the cell impedance in batteries. In silicon-based anode systems, the fast SEI growth can easily result in an increase in the impedance in electrodes and a decrease in capacity with cycling (30, 31). Figure 3K shows the cell impedance of DWSiNTs before and after cycling 1000 and 2000 times. No obvious impedance increase was detected, indicating no significant growth of SEI during cycling. The stable materials and SEI of DWSiNTs should afford excellent cycle life in batteries. The battery performance of the DWSiNT anodes was evaluated using deep charge/discharge cycles from 1V to 0.01V (Fig. 4 A to E). As shown in Fig. 4A, the specific reversible lithium extraction capacity of the DWSiNTs reached 1,780 mah/g at a rate of C/5 (1C=2A/g), which is four times higher than that of graphitic anodes. The gravimetric capacity of silicon alone is estimated to be very high, ~2970 mah/g. The cycling performance of three different types of Si electrodes with different SEI formation behavior is compared using C/5 deep charge/discharge cycling (Fig. 3): 1) hollow DWSiNTs with the 8

SiO 2 clamping layer (Black), 2) hollow silicon nanotubes without the SiO 2 constraining layer (Blue) and 3) solid silicon nanowires (Red). Under the deep charge/discharge, fast capacity fading for anodes made from silicon nanowires and nanotubes without the constraining oxide can be observed in Fig. 4A. For DWSiNT samples, no capacity decay was detected after 300 cycles, capacity retention after 500 cycles was 94%, and 76% of the initial discharge capacity remained after 900 cycles. Note that it takes more than one year to complete 900 cycles at the C/5 rate (Fig. 4A), and the fading here might have resulted from electrolyte leakage of the battery over such a long time. To obtain ultralong cycling results within a reasonable time period, faster charge and discharge tests were also carried out. Significantly, the DWSiNT anode cycled at 10C rate retains 93 % of its initial capacity after 4000 cycles, and 88% after 6000 cycles (Fig. 4B). The voltage profiles of the different cycles are shown in Fig. 4C. The lithiation potential shows a sloping profile between 0.1-0.01 V, consistent with the behavior of amorphous Si (13, 14). No obvious change in charge/discharge profile can be found after 6000 cycles for the DWSiNT anode, indicating superior and stable cycling performance. This extremely long cycle life, which already exceeds the DOE requirements for HEV and EV battery systems, can be attributed to the stable SEI and materials in the DWSiNTs. Coulombic efficiency (CE) is another important concern for silicon as anode material. The stability of the SEI layer helps to achieve high CE of >99.5% after the first cycle. The CE of the first cycle is 76 % since the constraining SiO 2 layer and the initial SEI formation consume some electrolyte, although this can be improved by pre-lithiation in our future studies. In addition, the thin and stable SEI and the nanoscale thickness of the walls in DWSiNTs can also enable outstanding high power rate capability. At fast charge/discharge rates between 1 C and 20 C, high and stable capacities between 2000 mah/g to 900 mah/g in the DWSiNTs were demonstrated (Fig. 4D and E). Clearly even at very high C rates, Li ions can rapidly pass through the thin SEI layer and constraining oxide to reach the silicon active material even at very high C rates. In summary, we have designed a DWSiNT structure to successfully address the Si material and SEI 9

stability issues, and we have demonstrated its use as a high performance anode with long cycle life (6000 cycles with 88% capacity retention), high specific charge capacity (~2970 mah/g at C/5, ~1000 mah/g at 12C) and fast charging/discharging rates (up to 20C). These performance parameters can enable the next generation of high performance batteries for portable electronics, electric vehicles and grid-scale applications. This successful materials design for silicon-based anodes could also be extended to other high capacity anode and cathode materials systems that undergo large volume expansion. 10

Fig. 1. Schematic of SEI formation on Si surfaces. (A) A solid Si nanowire expands upon lithiation. A thin layer of SEI forms in this lithiated and expanded state. During delithiation, the silicon structures shrink, and the SEI can break down into separate pieces, exposing fresh Si surface to the electrolyte. In later cycles, new SEI continue to form on newly exposed Si surfaces, and this finally results in a very thick SEI on the outside of the silicon nanowires. (B) Similarly, a thick SEI grows outside the Si nanotube without a mechanical constraining layer, which also has a varying and unstable interface with electrolyte. (C) Designing a mechanical constraining layer on the hollow Si nanotubes can prevent Si from expanding outside towards the electrolyte during lithiation; as a result, a thin and stable SEI can be built. 11

Fig. 2. (A) Schematic of the fabrication process of DWSiNTs. Polymer nanofibers (green) were first made by electrospinning. The polymer fibers were then carbonized and coated with Si (blue) by a CVD method. By heating the sample in air at 500 C, the inner carbon templates (black) were selectively removed, leaving continuous Si tubes with a SiO 2 mechanical constraining layer (red). (B) and (C) SEM images of synthesized DWSiNTs at high and low magnification. (D) TEM image of DWSiNTs, showing the uniform hollow structure with smooth tube walls. (E) Auger depth profiling of DWSiNTs, showing the double layer structure with the oxide shell on the outside. (F) and (G) Theoretical 12

modeling of hoop stress distribution before and after lithiation in single-walled Si nanotubes without surface coating (F) and DWSiNTs (G). Left column: Initial geometry of nanotubes. Right column: Hoop stress distribution through the wall of the nanotubes in the lithiated state. The unit of stress is MPa. (H) and (I) TEM images of an individual DWSiNT before and after lithiation, respectively. They show the tube wall expanded towards the inside, while the outside diameter of the tube remains constant. Fig. 3. (A)-(I) SEM images of different Si nanostructures before and after cycling. (A), (D) and (G): SEM images of Si nanowires, Si nanotubes and DWSiNTs, respectively. (B) and (E): SEM images of Si nanowires and Si nanotubes after 200 cycles; a thick SEI layer covering the nanostuructures can be seen. (H) DWSiNTs after 2000 cycles. The tubes are coated with a uniform SEI layer. (J) Statistical analysis of DWSiNTs before and after 2000 cycles. The average diameter of the initial tubes and the cycled tubes is 400 nm and 620 nm, indicating an average SEI thickness of 110 nm. (C) (F) and (I) SEM images of nanowires, nanotubes and DWSiNTs after cycling for 200, 200 and 2000 times, respectively. In these images, the SEI was selectively removed through chemical etching. (K) Impedance measurements of DWSiNTs after different numbers of cycles. 13

Fig. 4. Electrochemical characteristics of DWSiNTs tested between 1V and 0.01V. (A) Capacity retention of different Si nanostructures. All the samples were cycled at the same charge/discharge rate of C/5. The calendar life and delithiation capacity of DWSiNTs can also be seen in this figure. (B) Lithiation/delithiation capacity and Coulombic efficiency of DWSiNTs cycled at 12C for 6000 cycles. There is no significant capacity fading after 6000 cycles. (C) Voltage profiles were plotted for the 1 st, 1000 th, 2000 th, 3000 th and 6000 th cycles. (D) and (E): Galvanostatic charge/discharge profiles (D) and capacity (E) of DWSiNTs cycled at various rates from 1C to 20C. All electrochemical measurements (A)-(E) were carried out at room temperature in two-electrode 2032 coin-type half-cells. Capacity is calculated based on the weight of silicon. (32) 14

References and Notes: 1. M. Armand, J. M. Tarascon, Nature 451, 652 (2008). 2. J. B. Goodenough, Y. Kim, Chem. Mat. 22, 587 (2009). 3. M. S. Whittingham, Chem. Rev. 104, 4271 (2004). 4. Annual Progress Report on Energy Stroge Research and Develoment, U.S. Department of Energy (2010). 5. J. M. Tarascon, M. Armand, Nature 414, 359 (2001). 6. L. Y. Beaulieu, K. W. Eberman, R. L. Turner, L. J. Krause, J. R. Dahn, Electrochem. Solid State Lett. 4, A137 (2001). 7. J. O. Besenhard, J. Yang, M. Winter, J. Power Sources 68, 87 (1997). 8. T. D. Hatchard, J. R. Dahn, J. Electrochem. Soc. 151, A838 (2004). 9. P. R. Raimann et al., Ionics 12, 253 (2006). 10. W. J. Weydanz, M. Wohlfahrt-Mehrens, R. A. Huggins, J. Power Sources 81, 237 (1999). 11. X. W. Zhang et al., J. Power Sources 125, 206 (2004). 12. C. K. Chan et al., Nat. Nanotechnol. 3, 31 (2008). 13. L. F. Cui, R. Ruffo, C. K. Chan, H. L. Peng, Y. Cui, Nano Lett. 9, 491 (2009). 14. L. F. Cui, Y. Yang, C. M. Hsu, Y. Cui, Nano Lett. 9, 3370 (2009). 15. B. Hertzberg, A. Alexeev, G. Yushin, J. Am. Chem. Soc. 132, 8548 (2010). 16. H. Kim, B. Han, J. Choo, J. Cho, Angew. Chem.-Int. Edit. 47, 10151 (2008). 17. H. Kim, M. Seo, M. H. Park, J. Cho, Angew. Chem.-Int. Edit. 49, 2146 (2008). 18. A. Magasinski et al., Nat. Mater. 9, 353 (2010). 19. M. H. Park et al., Nano Lett. 9, 3844 (2009). 20. K. Q. Peng, J. S. Jie, W. J. Zhang, S. T. Lee, Appl. Phys. Lett. 93, 3 (2008). 21. T. Song et al., Nano Lett. 10, 1710 (2010). 22. D. Aurbach, J. Power Sources 89, 206 (2000). 23. P. Verma, P. Maire, P. Novak, Electrochimica Acta 55, 6332 (2010). 24. L. Y. Beaulieu, T. D. Hatchard, A. Bonakdarpour, M. D. Fleischauer, J. R. Dahn, J. Electrochem. Soc. 150, A1457 (2003). 25. Our preliminary experiment of nanoindentation on a fully lithiated silicon wafer surface showed dramatic change of yield strength down to 60MPa which is soft enough for inward expansion due to a mechanical clamping. 26. A. Greiner, J. H. Wendorff, Angew. Chem.-Int. Edit. 46, 5670 (2007). 27. D. Li, Y. N. Xia, Adv. Mater. 16, 1151 (2004). 28. The breaking ends of nanotubes shown in the SEM images (Fig. 2B and Fig. 3) were made by out force while preparing sample. 29. J. W. Choi et al., Nano Lett. 10, 1409 (2010). 30. C. K. Chan, R. Ruffo, S. S. Hong, Y. Cui, J. Power Sources 189, 1132 (2009). 31. R. Ruffo, S. S. Hong, C. K. Chan, R. A. Huggins, Y. Cui, J. Phys. Chem. C 113, 11390 (2009). 32. This work was partially supported by the Assistant Secretary for Energy Efficiency and Renewable Energy, Office of Vehicle Technologies of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231, Subcontract NO. 6951379 under the Batteries for Advanced Transportation Technologies (BATT) Program. This work is also partially supported by the Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under contract DE-AC02-76SF0051, through the SLAC National Accelerator Laboratory LDRD project. Y.C. acknowledges support from the King Abdullah University of Science and Technology (KAUST) Investigator Award (No. KUS-l1-001-12). S.W.L. also acknowledges support from KAUST (Award No. KUK-F1-038-02). H.W. acknowledges helpful disucssion with Prof. William D. Nix, Mr. Lucas Alexander Berla and Mr. Tom Carney, G. C. 15

acknowledges support from the Agency of Science, Technology and Research Singapore (A*STAR) National Science Scholarship. I.R. was supported by the Office of Science, Office of Basic Energy Sciences, of the US Department of Energy under Contract No. DE-FG02-04ER46163. M.T.M. acknowledges support from the Stanford Graduate Fellowship, the National Science Foundation Graduate Fellowship, and the National Defense Science and Engineering Graduate Fellowship. 16