Utilizing van der Waals slippery interfaces to enhance the

6 days ago - High specific capacity anode materials such as Silicon (Si) are increasingly being explored for next-generation, high performance Lithium...
0 downloads 4 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Energy, Environmental, and Catalysis Applications

Utilizing van der Waals slippery interfaces to enhance the electrochemical stability of Silicon film anodes in Lithium-ion batteries Swastik Basu, Shravan Suresh, Kamalika Ghatak, Stephen F Bartolucci, Tushar Gupta, Prateek Hundekar, Rajesh Kumar, Toh-Ming Lu, Dibakar Datta, Yunfeng Shi, and Nikhil A. Koratkar ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b00258 • Publication Date (Web): 05 Apr 2018 Downloaded from http://pubs.acs.org on April 6, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Utilizing van der Waals slippery interfaces to enhance the electrochemical stability of Silicon film anodes in Lithium-ion batteries *Swastik Basu1, *Shravan Suresh1, Kamalika Ghatak2, Stephen F. Bartolucci3, Tushar Gupta1, Prateek Hundekar4, Rajesh Kumar5, Toh-Ming Lu6, Dibakar Datta2, **Yunfeng Shi4, and **Nikhil Koratkar1,4 1

Department of Mechanical, Aerospace and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY, 12180, USA 2

Department of Mechanical and Industrial Engineering, Newark College of Engineering, New Jersey Institute of Technology (NJIT), Newark, NJ, 07102, USA

3

US Army Armaments Research Development and Engineering Center, Watervliet, NY, 12189, USA 4

Department of Materials Science and Engineering, Rensselaer Polytechnic Institute, Troy, NY, 12180, USA

5

University School of Basic & Applied Sciences, Guru Gobind Singh Indraprastha University, New Delhi, 110078, India 6

Department of Physics, Applied Physics and Astronomy, Rensselaer Polytechnic Institute, Troy, NY, 12180, USA

Keywords: Lithium-ion battery, silicon film anode, van der Waals interface, Interfacial slip, stable cycle life

*Equal Contribution **Correspondence should be addressed to Y.S. ([email protected]) and N.K. ([email protected])

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT High specific capacity anode materials such as Silicon (Si) are increasingly being explored for next-generation, high performance Lithium (Li) – ion batteries. In this context, Si films are advantageous compared to Si nano-particle based anodes, since in films the free volume between nano-particles is eliminated, resulting in very high volumetric energy density. However Si undergoes a volume expansion (contraction) under lithiation (delithiation) of up to 300%. This large volume expansion leads to stress build-up at the interface between the Si film and the current collector, leading to delamination of Si from the surface of the current collector. To prevent this, adhesion promotors (such as chromium interlayers) are often used to strengthen the interface between the Si and the current collector. Here we show that such approaches are in fact counter-productive and far better electrochemical stability can be obtained by engineering a van der Waals “slippery” interface between the Si film and the current collector. This can be accomplished by simply coating the current collector surface with graphene sheets. For such an interface, the Si film slips with respect to the current collector under lithiation/delithiation, while retaining electrical contact with the current collector. Molecular dynamics simulations indicate: (i) less stress build-up and (ii) less stress ‘cycling’ on a van der Waals slippery substrate as opposed to a fixed interface. Electrochemical testing confirms more stable performance and much higher coulombic efficiency for Si films deposited on graphene-coated Nickel (i.e. slippery interface) as compared to conventional Nickel current collectors.

ACS Paragon Plus Environment

Page 2 of 21

Page 3 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

INTRODUCTION Lithium-ion batteries (LIBs) have been extensively utilized in portable electronics because of their relatively high energy density and long life cycle.1 The growth in demand for these batteries is increasing with proliferation of consumer electronics products. In fact, the high storage capacity and better efficiency of Li-ion batteries may allow their use in various electric grid applications, which can improve the reliability of energy harvested from renewable sources.2 Also, Li-ion batteries will significantly reduce greenhouse gas emissions, if a substantial part of the gasoline powered transportation system is replaced by electric vehicles (EVs).3 As the theoretical capacity of traditional graphite anodes (~370 mAh/g), cannot satisfy the increasing requirements for high capacity LIBs, much attention has been focused on higher capacity anode materials. Of the various possible materials currently being explored, silicon (Si) exhibits the highest gravimetric (~4200 mAh/g ) and volumetric (~9800 mAh/cm3) capacities.4-8 One of the first attempts to deploy Si anodes in LIBs was as films deposited on Cu foil current collector substrates.4 However mechanical instabilities have consistently led to poor electrochemical performance of such Si film anodes. The principal reason for mechanical failure is stress build up at the interface between the expanding/contracting Si film and the rigid (fixed) current collector substrate.5 The most popular research direction to counter this challenge has been focused on studying various kinds of nanostructured Si materials. Various studies have been done by i) synthesizing sophisticated nanostructured Si, including nanowires6,7, nanotubes8, nanopillars, nanospheres9,10, etc.; ii) embedding of nanostructured Si particles into an active or inactive matrix11–15, iii) hybridization with carbonaceous materials to electrochemically improve the performance16–20, iv) exploring new polymer binding materials21,22, and v) core-shell Si nanoparticles.23–25 Perhaps the most successful approach to date has involved conventional carbon coating on nano Si materials26-31. This can be accomplished by directly mixing Si nanopowder with graphite, mesocarbon microbeads or hard carbon to achieve better cycle life.26 Another variation of this approach is a yolk-shell structure, where Si nanoparticles were sealed inside carbon shells with the void space allowing the Si to expand and contract, leading to a more stable solid electrolyte interface.27 Many such core shell structures have been further designed using different forms of carbon.28–31 Even though improvement in the design of nanostructured Si has led to better performance and cycle life, several disadvantages persist. Si nanoparticle based anodes offer much less packing density than Si films, which results in lower volumetric energy density.32 Moreover, yolk-shell type nano Si architectures are complex in design and expensive to manufacture. Further, it should be noted that portable electronics and even batteries for automotive applications are “volume” limited rather than weight limited.32 Therefore, if Si film anodes can be rendered stable in a LIB then this has important implications for development of batteries with high volumetric performance. Here we demonstrate a pathway to achieving such an outcome by using graphene sheets to engineer a van der Waals “slippery” interface between the Si-film anode and the metal current collector substrate.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

RESULTS AND DISCUSSIONS I. Stress Evolution The evolution of the normal and shear stress profile is analyzed using molecular dynamics (MD) simulations upon lithiation and delithiation of an amorphous Si (a-Si) film deposited on different substrates. In order to compute the profiles of the normal and shear stress, the system was divided into thin slabs with equal thickness of 0.55 nm along the x-direction of the simulation box. For computation of normal stress, the height of the slabs along the z direction is the height of the system (2.15 nm) excluding the fixed substrate, while for computation of the shear stress τ slabs of height 1 nm above the substrate are chosen. Figure 1 shows the profiles of the stress tensor components along the normal (σ

) and shear (τ ) directions in two systems with no-slip substrate (fixed a-Si substrate) and slip substrate (monolayer graphene). Here positive (negative) values of the normal stress components indicate compressive (tensile) stresses that act to contract (expand) the film, while positive (negative) values of the shear stress components indicate shear stress tending to turn the element counter-clockwise (clockwise). It is observed from the σ

profiles that at a highly lithiated stage (Li/Si = 3.5) there is a considerable build-up of compressive stress in the central region of the film where the Li-Si is in contact with the fixed substrate (Fig. 1- left panel). This is because the a-Si network tries to expand in volume upon lithiation, while the fixed substrate constrains this volume expansion, resulting in considerable shear stress distribution between the substrate and the a-Si. The shear stress from the top and bottom substrate together lead to overall compressive stress in the a-Si in the middle. This rather high compressive stress from the substrate acts to counter the thermodynamic driving force towards lithiation, such that the a-Si interior region contains much less lithium atoms than the outer region. The stress distribution roughly reverses its sign during delithiation. This is because the expanded a-Si network tries to revert to its original density while the sticky substrate prevents the a-Si from shrinking. Thus, the shear distribution between the no-slip substrate and the a-Si changes sign, while the overall stress along the x direction becomes tensile at the center of the a-Si film. It should be noted that both the compressive stress reached during lithiation and the tensile stress reached during delithiation are significant. Such large stress cycling over extended lithiation/delithiation cycles will invariably lead to fatigue damage, leaving the Si film susceptible to fracture and pulverization. On the other hand, one can see that the graphene substrate creates a slippery van der Waals interface with the a-Si layer, which allows the a-Si network to expand freely upon lithiation and shrink freely upon delithiation (Fig.1 – right). Thus there is considerably lower build-up of normal and shear stress in comparison to the no-slip substrate. Consequently, the resulting stress cycling during lithiation/delithiation cycling (Fig.1 – right) is much less, which would greatly alleviate fatigue induced crack propagation and failure. These results suggest that controlled mechanical failure (i.e. interfacial slippage) at the interface between the Si film and the current

ACS Paragon Plus Environment

Page 4 of 21

Page 5 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

collector can greatly alleviate stress build-up and stress cycling in the Si film, which is expected to enhance its stability. It should be noted that such interfacial slip would not result in electrical disconnection of the Si film as it would remain in electrical contact with the substrate.

II. Interface Properties Having established that the structural integrity of the a-Si electrode is sensitive to the aSi/substrate interface (strong adhesion for no-slip substrate, while low adhesion for slippery substrate), we will examine the interfacial adhesion properties of a number of a-Si/substrate combinations. In general, the work of separation W provides a useful measure of the strength with which two materials adhere to each other, and larger the strength of adhesion, higher will be the stress buildup at the interface due to lithiation induced volume expansion, and vice versa. The standard definition of W for an interface between two materials is, W =

 E + E − E A

 E  is the total energy of slab i, E is the total energy of the interface system with the slab materials, and A is the total area of the interface.33 We have used a-Si/a-Si, a-Si/Cu, a-Si/Ni and a-Si/Graphene vacuum interface models (Figure 2) to calculate W for a-Si as a function of different substrates using first-principles density functional theory (DFT) calculations. The number of metal layers used in the calculations were determined after studying the convergence of W as a function of slab thickness (Fig. S1 in supporting information).

The results for W for the different interfaces are summarized in Figure 3 and indicate that the interface energy for the a-Si/graphene system is around one fifth of that for the Cu and Ni substrates and less than half of that for the a-Si/a-Si substrate. This indicates that compared to traditional metal current collector films such as Cu and Ni, graphene offers a much weaker (slippery) interface with the amorphous Si film. The greater work of separation of a-Si/Cu and aSi/Ni relative to a-Si/Si indicates that the “non-slip” condition is more valid for a-Si/Cu and aSi/Ni relative to a-Si/Si. Since MD results (Figure 1) indicate lower stress build-up during cycling of a-Si/Graphene interface compared to the a-Si/a-Si interface, one can conclude that the performance of the a-Si/Graphene interface would be even better when compared to the nonslipping (i.e. fixed) a-Si/Cu and a-Si/Ni interfaces during cycling. Therefore slippage during lithiation/delithiation of the amorphous Si film with respect to the graphene substrate should greatly alleviate stress build-up and stress cycling at the interface and therefore prolong the cycle life of the a-Si film when compared to traditional metal current collectors such as Ni and Cu.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 6 of 21

Page 7 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 1: (a1-a2) show snapshots of the simulation systems with a rigid non-slip substrate (a1) and a rigid slip substrate (a2) during a lithiation and delithiation cycle. The snapshots are taken prior to lithiation, at a highly lithiated stage (Li/Si = 3.5) and a highly delithiated stage (Li/Si = 0.3), from top to bottom. The red atoms are silicon while the blue atoms are lithium. (b1-b2) show the normal stress σxx profile at Li/Si=3.5 (black lines) and Li/Si=0.3 (orange lines) for the systems with non-slip (b1) and slip substrate (b2), respectively. σxx is averaged over the simulation system excluding the substrate every 0.55 nm along the x-direction. (c1-c2) show the shear stress τxz profile at Li/Si=3.5 (black lines) and Li/Si=0.3 (orange lines) for the systems with non-slip (c1) and slip substrate (c2), respectively. τxz is averaged over the region within 1nm above the bottom substrate (as indicated by the box) every 0.55 nm along the x-direction. The error bar shows the stress fluctuation over ten independent stress measurements.

Figure 2: Setup for the interface system of a-Si on a-Si, Cu (111), Ni (111) and monolayer graphene substrates used for carrying out the first principles (DFT) calculations.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: Interfacial work of separation for relaxed a-Si/a-Si, a-Si/Cu, a-Si/Ni and a-Si/graphene interfaces.

III. Electrochemical Testing To validate our prediction that graphene substrates could be used to alleviate stress build up and prolong the cycle life of Si films, we performed constant current charge-discharge cycling of a-Si films deposited on metal (Ni) and graphene-coated Ni current collector substrates. Since the work of separation of a-Si is similar for both Ni and Cu substrates, one can expect similar results on Cu as well and this should be the subject of future study. The graphene-coated Ni foils used in this study were purchased from ACS Materials Inc. Cross-sectional SEM imaging of the graphene coated Ni current collectors (Fig. S2A) was carried out. The SEM imaging indicates that the multilayer graphene coating is ~1 µm thick. The graphene film is of high quality as indicated by its Raman spectra. From Fig. S2B, it is evident that the Raman D peak, related to disorder, is miniscule compared to the G peak (ID/IG = 0.0014). In addition, there is no D’ peak (another disorder related peak), nor is the D+D’ peak present in the spectrum. All these signs indicate that the defect density in the graphene film is negligibly low. Figures 4a and 4b show schematics of the Si sputter deposition on bare Ni foil and graphenecoated Ni foil. It should be noted that Si films deposited by sputtering are amorphous in nature.32 Scanning electron microscopy (SEM) imaging of the Si film deposited on the Ni foil (Figure 4c) shows that the Si film mimics the roughness of the Ni foil. On the other hand, Si films deposited on graphene-coated Ni (Si-Gr) in Fig. 4d show graphene-like morphology with grain boundaries. For the theoretical calculations and molecular dynamics (MD) simulations, we used a monolayer graphene sheet at the interface between the Si film and the Ni current collector. This is because van der Waals forces are short range and hence there is negligible interaction between Si and the second graphene layer (i.e. the layer below the nearest graphene layer). Therefore purely from a

ACS Paragon Plus Environment

Page 8 of 21

Page 9 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

“theoretical” perspective, a monolayer graphene sheet is sufficient to capture the interfacial stress evolution and illustrate the principle of van der Waals slip. While theoretically, we can use a defect-free and perfect graphene sheet, experimentally this is not the case. Monolayer graphene grown by chemical vapor deposition (CVD) is defective and imperfect with a multitude of topological and vacancy defects32 not to mention wrinkles/folds and tears. Therefore for the experiments, the graphene film that was deployed was multilayered. The use of such a multilayered and relatively defect-free (Fig. S2) graphene film ensures that there is minimal interaction between the Ni substrate and the Si film and therefore a “slippery” interface between the Si layer and the graphene film is ensured. The cyclic voltammograms (CV) of Si and Si-Gr films were acquired at a scan rate of 0.5 mV s-1 and are shown in Figure 5a. The peak at ~0.2 V during the cathodic scan and ~0.56 V during the anodic scan represent the electrochemical lithiation/delithiation reactions of Si. This indicates that only Si is the active material and the contribution of graphene is negligible. To verify the capacity contribution of graphene, we also performed electrochemical tests on pristine graphene films and compared the capacity of Si-Gr with the pristine graphene films at a current density of 1.8 A g-1. As shown in Figure 5b, the capacity of the pristine graphene film was negligible as compared to the Si-Gr films. The CV curves in Figure 5a also show that the area under the curve of Si-Gr films is higher, and the redox peaks are sharper. The area under the CV curves indicates charge storage capacity and the intensity of the CV peaks is proportional to the charge transfer due to the electrochemical reactions. Thus, the enhancement of area under the curve and the increase in intensity of the currents is due to the improved charge transfer kinetics because of the graphene buffer layer at the interface. Further, due to the presence of graphene, the lithiation/delithiation reactions of Si films are also enhanced by reducing the interfacial charge transfer resistance. Hence, although the direct capacity contribution of the graphene buffer layer is negligible, it indirectly increases the charge storage capacity due to an enhanced “electrical” interface with the Si film. Such an enhanced electrical interface has also been reported for Si nanowire anodes deposited on graphene substrates.33

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4: Schematics of Si sputter deposition on Bare Ni foil (Si-Ni) (a) and multilayer graphene coated Ni foil (Si-Gr) (b). SEM images of (c) Si-Ni electrode and (d) Si-Gr electrode before electro-chemical cycling, mimicking the surface roughness of Ni and Graphene respectively.

The galvanostatic charge/discharge cycling test results are shown in Figures 5d-e. The baseline Si-Ni electrode exhibited a rapid drop in capacity after ~25 charge/discharge steps. To verify this several coin cells were tested to check for repeatability (Fig. S3-S4). Interestingly, it can also be seen in Figure 5c that the coulombic efficiency (CE) followed the same trend as the curve in Figure 5e, i.e. after ~25 cycles it dropped rapidly for Si films (average CE of ~98% after 50 cycles). On the other hand, the Si-Gr films exhibited an excellent average CE of ~99.2%. As seen in Figure 5e, the specific capacity of Si films decayed rapidly, retaining only ~190 mAh g-1 after 150 cycles. The Si-Gr films were far more stable with a capacity retention of ~800 mAh g-1 after 150 cycles. The volumetric capacity of the Si-Gr film averaged over 150 cycles is ~5462 mAh cm-3 which is far superior to that of conventional32 graphitic anodes (~400 mAh cm-3). The electrochemical characterizations confirm the superior cycle life of Si films due to the presence of a graphene buffer layer. The graphene layer creates a slippery interface, which plays a key role in preventing fracture due to volume expansion of Si films. After the first few cycles, Si films form a mud-cracked surface which is typically observed during their electrochemical cycling. This results in the formation of isolated Si islands, which will expand/contract during lithiation/delithiation. In the case of Si films deposited on Ni foil, there is a strong adhesion at the Si-Ni interface. During lithiation of Si films, the volume expansion is restrained by the

ACS Paragon Plus Environment

Page 10 of 21

Page 11 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

interfacial adhesion, which results in build-up of stress in the Si films. This causes fracture near the Si-Ni interface and accelerates the delamination of Si film (Figure 6a). In the case of Si-Gr films, the graphene buffer layer allows the Si film to slip. Hence the expansion of Si in the presence of a graphene buffer is relatively unconstrained and the fracture due to adhesion stresses is mitigated (Figure 6b). Thus, we can see that due to strong adhesion on metal current collectors, Si films would be eviscerated rapidly due to mechanical stresses. On the other hand, a slippery van der Waals interface due to its low interfacial adhesion mitigates fracture. Our results indicate that such low energy interfaces can be engineered using graphene buffer layers and this concept could also be extended to other high specific capacity anode materials (such as As, Sb, P, Al etc.) that undergo large volume expansion on lithiation. Electron microprobe analysis was performed after electrochemical cycling to confirm the delamination of Si films. The composition maps of Si and Si-Gr films exhibited a striking difference. In Si films, the Ni current collector can be clearly seen after cycling which means the Si film has delaminated during electrochemical cycling. Contrastingly, a thick Si film can be seen in Si-Gr films samples after 150 charge/discharge cycles. As shown in Figure 6c, the wavelength dispersive spectra acquired at ~10 keV show a strong presence of Ni in the Si film, which is unlike the Si-Gr film. This indicates that the Si film which remains on the graphenecoated Ni after cycling is thicker than the Si film that remains on the Ni foil. Furthermore, the SEM images in Figures 6d-e that were acquired post electrochemical cycling (after 150 chargedischarge steps) indicate the presence of pulverized debris on the Si films on Ni, whereas mudcracked Si films that are still intact can be seen in the case of cycled Si-Gr films. This was also confirmed by high resolution SEM imaging (Fig. S5) of the cycled Si-Ni and Si-Gr electrodes. Mud-cracked Si films that are still intact can be seen in the case of cycled Si-Gr films, whereas the Si film is seen to be delaminated and exposing the Ni substrate in the case of cycled Si-Ni films (Fig. S5). The presence of the graphene layer below the mud-cracked Si film was confirmed by Raman spectroscopy acquired after the electrochemical cycling (150 chargedischarge cycles) showing the characteristic D (~1360 cm-1) and G (~1580 cm-1) band peaks of graphene in the Si-Gr film (Fig. S6). Both the Si and Si-Gr films are unprotected from the top which implies that there will be losses due to pulverization and an unstable solid electrolyte interface (SEI). Hence, a capacity fade can be observed in both Si and Si-Gr film. The reduction of specific capacity and CE at ~25 cycles can be ascribed to the loss of active Si due to electrical disconnections. Further, the fractured surfaces near the interface of Si-Ni also re-expose the surface of Si film to the electrolyte, thereby augmenting the growth of SEI, which further drops the CE. On the contrary, Si-Gr films which slip at the interface remain unconstrained and therefore, the stresses at the interface are minimized. Hence, in the case of Si-Gr films the fade of the specific capacity and the CE is significantly reduced. Thus, we see that the bottom interface (with the current collector) plays a major role in the electrochemical stability of Si films in LIB anodes. Further improvements in electrochemical stability and coulombic efficiency can be expected as the top interface with the electrolyte is also protected by draping a graphene film over the a-Si film as shown in Ref. 32.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5: Electrochemical characterization of Si-Gr electrode. (a) Cyclic voltammograms acquired at a scan rate of 0.5mVs-1. The CVs show that Si is the active anode material in both SiNi and Si-Gr. (b) Comparison of the areal average capacity of the graphene coating with and without the Si film indicating negligible contribution of the graphene coating to the charge capacity of the anode. (c) Plots indicating an improvement in coulombic Efficiency of the Si-Gr. (d) Galvanostatic charge-discharge curves. (e) Electrochemical cycling performance of Si-Gr showed an enhanced capacity of ~800 mAh/g compared to a capacity of ~190 mAh/g exhibited by Si-Ni. The galvanostatic cycling was performed at a current density of ~1.8 A/g.

ACS Paragon Plus Environment

Page 12 of 21

Page 13 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 6: (a) Schematic representation of fracture near the interface due to electrochemical cycling of Si film on bare Ni foil. (b) Slip at the interface due to the graphene coating on Ni suppresses Si film delamination near the interface on electrochemical cycling. (c) Wavelength Dispersive Spectra acquired at ~10 KeV show a strong presence of Ni in Si-Ni film in comparison to the Si-Gr film after cycling. (d) Postcycling SEM image of Si-Gr electrode after 150 cycles indicating the toughened mud-cracked structure. (e) Postcycling SEM image of Si-Ni electrode after 150 charge/discharge cycles. Si has delaminated exposing the bare Ni surface with large debris on the surface.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

CONCLUSIONS To summarize, we have used MD and DFT simulations to explore the fundamental reasons behind the failure of bulk Si on a typical metal (i.e. rigid or fixed) substrate. Further, a graphene draped metal substrate has been introduced for the a-Si film which serves to “unconstrain” the Si film and has been shown to significantly reduce the stress build-up and stress cycling at the interface. The concept has been proven experimentally where Si has been deposited on multilayer graphene coated Ni, which provides a better specific capacity and a more stable coulombic efficiency as an anode than Si deposited on Ni foil. Such a Si film with a slippery van der Waals interface has been shown to be stabilized for up to 150 cycles with a capacity of ~800 mAh g-1, whereas the Si film deposited on Ni could retain only 190 mAh g-1. With such densely packed Si films, an average volumetric capacity of ~5462 mAh cm-3 was achieved with the Si film deposited on graphene, which is ~10-fold higher than conventional graphitic anodes. Graphene was used as the substrate on top of the current collector and no protective layer on the Si-electrolyte interface was used in order to demonstrate that even without attempting to stabilize the SEI with a protective graphene drape, it is possible to improve the performance and cyclic stability by preventing delamination at the Si/current-collector interface, as predicted by the MD and DFT simulations. In the MD simulations, the fixed substrate was taken to be Si, rather than the typical Ni and Cu substrates which are used in experiments. This is because in MD we are yet to develop a reactive potential involving Cu/Ni, Si and Li, but the strong Si-Si interaction essentially demonstrates the same physical restraints of a fixed (or rigid) interface. Comparative study of the evolution of stress profile and the interface properties give quantitative evidence of the fundamental reasons that lead to better performance of the Si anode when lithiated on a substrate with weak van der Waals interaction. Our results indicate: i) “slip” is observed between the graphene layer and the expanding Silicon, ii) Less stress build-up, and iii) Less stress ‘cycling’ is observed when Si is supported on graphene. These predictions were validated by experiments and indicate that Si films with superior volumetric capacity, coulombic efficiency and cycling stability can be engineered by exploiting van der Waals slippery interfaces between the Si films and the current collector substrates. In this work, we have focused on thin Si films (mass loading of ~0.05 mg/cm2), to show the proof-of-principle. For thicker Si films a layer-by-layer deposition strategy with multiple graphene films as interlayers might be necessary and this should be the focus of future work. The concept of using van der Waals slippery interfaces to enhance electrochemical stability could also in principle be extended to other high specific capacity anode materials (such as As, Sb, P, Al etc.) that undergo large volume expansion on lithiation.

ACS Paragon Plus Environment

Page 14 of 21

Page 15 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

METHODS I. Molecular Dynamics Simulation To simulate the lithiation process in amorphous silicon, MD simulation with the ReaxFF potential developed for Li-Si systems by Jung34 is employed. The total system energy is the sum of contributions from the electron lone pairs, over-coordination, under-coordination, valence and torsion angles, conjugation, hydrogen bonding, and van der Waals and Coulomb interactions.35 Previously this force field has been successfully used to determine the mechanical properties of amorphous Li Si alloys.36 For simplicity, the interaction between carbon and lithium/silicon is modeled using a very weak Lennard-Jones potential. The 12−6 Lennard-Jones potential (V(r)) for two non-bonded atoms is given by V"r) = −Ar $% + Br $, where A and B denote the attractive and repulsive Hamaker constants, respectively. From the work by Chan and coworkers,37 we choose an A value of 3959 eV Å% and repulsive constant B of 904.438 eV Å , respectively. We used the LAMMPS software, with a Verlet integration time step of 1 Femtosecond. In our simulation, amorphous Si structure is generated by melting and subsequent quenching of single-crystal silicon. The amorphous silicon film tested is periodic in the transverse directions and has two surfaces in the horizontal direction as in Figure 1. In the case of no-slip substrate, a thin a-Si layer with a thickness of 0.54 nm is kept fixed. In the case of slip substrate, a rigid graphene sheet is placed next to the a-Si, acting as the substrate. In both situations, due to periodic boundary conditions, the active a-Si anode material is essentially sandwiched by the substrate. The system is first equilibrated under a canonical NVT ensemble at 300 K for 100 ps. The lithiation process is simulated by depositing lithium towards the free surface of the a-Si film with a driving force along the horizontal direction. A three step process is implemented; i) first, four lithium atoms are introduced in regions defined close to each of the two surfaces of the amorphous silicon, ii) followed by application of a small external force of 0.7 nN to these lithium atoms to drive them into the a-Si system in a process analogous to the voltage-driven lithiation of a-Si in an electrolytic cell.32 iii) Freshly lithiated system is allowed to relax under NVT ensemble for 10 ps. These steps are repeated for lithiation of the Si film to the desired extent. The delithiation process is carried out by selecting four lithium atoms closest to each of the two surfaces, which are then pulled out of the bulk by application of an external force of 0.7 nN, in a direction opposite to that during lithiation. Once the lithium atoms taken out are in vacuum, they are deleted, and the incrementally delithiated system is allowed to relax under NVT ensemble for 10 ps. Due to the computational expense associated with the ReaxFF force field, the initial system size is limited to around 1000 atoms. The temperature of 300 K is maintained by a Nose Hoover thermostat with a damping parameter of 0.01 fs $ . II. DFT Calculations First principle calculations were carried out to study the interface properties of a-Si with different substrates. In order to create the a-Si, ab Initio Molecular Dynamics (AIMD) simulation were

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

performed at a very high temperature of 1200 K in a NVT ensemble with a Nose thermostat for around 5000 MD time steps, with an interval of 1 femtosecond. The heating process was followed by the rapid cooling process to 298 K (room temperature) within 5000 MD time steps. Upon the end of the final RT AIMD calculation, the structure was further relaxed using DFT in order to achieve energy minimized structure. The mass density and structural signatures such as the radial distribution function (RDF) of the a-Si were observed with the dominant features consistent with previous quantum calculations.38 To calculate the interface energy, individual slab models of a-Si and the different substrates were created by adding ~12 Å vacuum to the zdirection in the simulation cell. The supercell containing the interface is constructed by placing the a-Si and different substrates together and adding ~14 Å of vacuum along the z-direction. Both the interface models were relaxed using DFT without any cell relaxation. This methodology was adapted from a previous reported study.39 In this study, four different types of substrates have been considered namely, a-Si, Cu (111), Ni (111) and Graphene. We generated a-Si, graphene, Ni and Cu substrates first with similar cell dimensions and then we accordingly created a-Si samples of 64 Si atoms each with slightly different dimensions according to the different substrates. This gives rise to distinctive a-Si samples for a-Si, graphene, Ni (111) and Cu (111). After the end of the DFT relaxation, each of the a-Si’s RDF was checked and verified in order to get the confirmations of their amorphous nature. An accurate estimate of long range van der Waals (vdW) interactions is important to study the aSi/graphene interface, which is not properly calculated by conventional DFT methods, using the standard (semi)local exchange correlation functionals such as the local density approximation (LDA)40 or the generalized gradient approximation (GGA)41. An important vdW-correction approach is the vdW-density functional (vdW-DF)42 which combines nonlocal correlations directly within a DFT functional. All DFT based calculations in this study have been performed using the optPBE functional within the vdW-DF family, implemented in the VASP package by Klimes and coworkers.43-45 The plane-wave energy cutoff for different interfaces (for all the calculations) was taken as 550 eV. The convergence tolerance for the electronic relaxation was 10$/ eV/cell, and the total energy was calculated with the linear tetrahedron method with Blochl corrections. All structural relaxations employed conjugate gradient methods to minimize the total energy of the structure, and the required Hellmann-Feynman force on each atom was less than 0.02 eV/Å. For all our interface calculations, 3×3×1 Γ-centered k-meshes were employed. In this study, we relaxed all atoms and have not fixed any atomic positions. III. Experiment Silicon films of ~200 nm thickness were deposited on bare Ni foils and graphene-coated Ni foils by DC sputtering in an AJA sputter system. An N-type Silicon (Plasmaterials Inc.) target (~2 inches in diameter) was used. The base pressure of the chamber was ~1×10-7 Torr and the working pressure was ~3 mTorr. Scanning electron microscopy was performed on an FEI

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Nanolab 600i Dual Beam SEM operating at 10 kV. The mass loading of the Silicon electrode is ~0.05 mg/cm2. The deposited Si films were tested as anodes in 2032 type coin cells. The cells were assembled inside a glove box (Mbraun Labstar) with an oxygen and moisture content of < 1ppm. Lithium foil was used as the counter electrode and a Celgard 2340 polypropylene membrane was used as the separator. 1M LiPF6 in a 1:1 volumetric mixture of ethylene carbonate (EC) and diethyl carbonate (DEC) was used as the electrolyte. The lithium metal foil and electrolyte were purchased from MTI Corp. and Sigma Aldrich, respectively. An Arbin BT2000 Battery Tester was used to perform the charge/discharge tests and a potentiostat (Gamry Instruments Inc.) was used to obtain the cyclic voltammograms.

Acknowledgements: N.K. acknowledges support from the USA National Science Foundation (Awards 1435783, 1510828, 1608171) and the John A. Clark and Edward T. Crossan endowed Chair Professorship at the Rensselaer Polytechnic Institute. D.D. and K.G. acknowledge the support of the Extreme Science and Engineering Discovery Environment (XSEDE) for the computational facilities (Start Up Allocation - DMR170065 & Advanced Allocation – MSS170034, DMR180013)

Supporting Information Available: Test of convergence for work of separation as a function of substrate thickness, cross-section SEM and Raman spectroscopy characterization of the graphene coated Ni current collector prior to electrochemical cycling, repeatability tests for Si-Ni and SiGr anodes, high resolution SEM imaging of the Si-Ni and Si-Gr electrodes in the post-cycled state and Raman spectroscopy analysis of the Si-Ni and Si-Gr electrodes in the post-cycled state.

REFERENCES (1)

Jeong, G.; Kim, Y.-U.; Kim, H.; Kim, Y.-J.; Sohn, H.-J. Prospective Materials and Applications for Li Secondary Batteries. Energy Environ. Sci. 2011, 4, 1986-2002.

(2)

Tarascon, J.-M.; Armand, M. Issues and Challenges Facing Rechargeable Lithium Batteries. Nature 2001, 414, 359–367.

(3)

Pacala, S.; Socolow, R. Stabilization Wedges: Solving the Climate Problem for the next 50 Years with Current Technologies. Science 2004, 305, 968–972.

(4)

Maranchi, J. P.; Hepp, A. F.; Kumta, P. N. High Capacity, Reversible Silicon Thin-Film Anodes for Lithium-Ion Batteries. Electrochem. Solid-State Lett. 2003, 6, A198-A201.

(5)

Liang, B.; Liu, Y.; Xu, Y. Silicon-Based Materials as High Capacity Anodes for next Generation Lithium Ion Batteries. J. Power Sources 2014, 267, 469−490.

(6)

Chan, C. K.; Peng, H.; Liu, G.; McIlwrath, K.; Zhang, X. F.; Huggins, R. a; Cui, Y. High-

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Performance Lithium Battery Anodes Using Silicon Nanowires. Nat. Nanotechnol. 2008, 3, 31–35. (7)

Cui, L.; Yang, Y.; Hsu, C.; Cui, Y. Carbon - Silicon Core - Shell Nanowires as High Capacity Electrode for Lithium Ion Batteries. Nano Lett. 2009, 9, 1–5.

(8)

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. Il; 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.

(9)

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.

(10)

Liu, B.; Soares, P.; Checkles, C.; Zhao, Y.; Yu, G. Three-Dimensional Hierarchical Ternary Nanostructures for High-Performance Li-Ion Battery Anodes. Nano Lett. 2013, 13, 3414–3419.

(11)

Saint, J.; Morcrette, M.; Larcher, D.; Laffont, L.; Beattie, S.; Pérès, J. P.; Talaga, D.; Couzi, M.; Tarascon, J. M. Towards a Fundamental Understanding of the Improved Electrochemical Performance of Silicon-Carbon Composites. Adv. Funct. Mater. 2007, 17, 1765–1774.

(12)

Liu, W.-R.; Wang, J.-H.; Wu, H.-C.; Shieh, D.-T.; Yang, M.-H.; Wu, N.-L. Electrochemical Characterizations on Si and C-Coated Si Particle Electrodes for LithiumIon Batteries. J. Electrochem. Soc. 2005, 152, A1719-A1725.

(13)

Yang, X.; Wen, Z.; Xu, X.; Lin, B.; Lin, Z. High-Performance Silicon/Carbon/Graphite Composites as Anode Materials for Lithium Ion Batteries. J. Electrochem. Soc. 2006, 153, A1341-A1344.

(14)

Kim, I. S.; Blomgren, G. E.; Kumta, P. N. Si-SiC Nanocomposite Anodes Synthesized Using High-Energy Mechanical Milling. J. Power Sources 2004, 130, 275–280.

(15)

Dimov, N.; Kugino, S.; Yoshio, M. Carbon-Coated Silicon as Anode Material for Lithium Ion Batteries: Advantages and Limitations. Electrochim. Acta 2003, 48, 1579–1587.

(16)

Jang, S. M.; Miyawaki, J.; Tsuji, M.; Mochida, I.; Yoon, S. H. The Preparation of a Novel Si-CNF Composite as an Effective Anodic Material for Lithium-Ion Batteries. Carbon 2009, 47, 3383–3391.

(17)

Martin, C.; Alias, M.; Christien, F.; Crosnier, O.; Bélanger, D.; Brousse, T. GraphiteGrafted Silicon Nanocomposite as a Negative Electrode for Lithium-Ion Batteries. Adv. Mater. 2009, 21, 4735–4741.

(18)

Wang, W.; Kumta, P. N. Nanostructured Hybrid Silicon/Carbon Nanotube Heterostructures: Reversible High-Capacity Lithium-Ion Anodes. ACS Nano 2010, 4, 2233–2241.

(19)

Xiang, H.; Zhang, K.; Ji, G.; Lee, J. Y.; Zou, C.; Chen, X.; Wu, J. Graphene/nanosized Silicon Composites for Lithium Battery Anodes with Improved Cycling Stability. Carbon 2011, 49, 1787–1796.

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(20)

Xiao, Q.; Fan, Y.; Wang, X.; Susantyoko, R. A.; Zhang, Q. A Multilayer Si/CNT Coaxial Nanofiber LIB Anode with a High Areal Capacity. Energy Environ. Sci. 2014, 7, 655– 661.

(21)

Liu, G.; Zheng, H.; Kim, S.; Deng, Y.; Minor, A. M.; Song, X.; Battaglia, V. S. Effects of Various Conductive Additive and Polymeric Binder Contents on the Performance of a Lithium-Ion Composite Cathode. J. Electrochem. Soc. 2008, 155, A887-A892.

(22)

Liu, G.; Zheng, H.; Simens, A. S.; Minor, A. M.; Song, X.; Battaglia, V. S. Optimization of Acetylene Black Conductive Additive and PVDF Composition for High-Power Rechargeable Lithium-Ion Cells. J. Electrochem. Soc. 2007, 154, A1129-A1134.

(23)

Zhang, C.; Gu, L.; Kaskhedikar, N.; Cui, G.; Maier, J. Preparation of silicon@Silicon Oxide Core-Shell Nanowires from a Silica Precursor toward a High Energy Density LiIon Battery Anode. ACS Appl. Mater. Interfaces 2013, 5, 12340–12345.

(24)

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.

(25)

Kim, C.; Ko, M.; Yoo, S.; Chae, S.; Choi, S.; Lee, E.-H.; Ko, S.; Lee, S.-Y.; Cho, J.; Park, S. Novel Design of Ultra-Fast Si Anodes for Li-Ion Batteries: Crystalline Si@amorphous Si Encapsulating Hard Carbon. Nanoscale 2014, 6, 10604–10610.

(26)

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.

(27)

Liu, N.; Wu, H.; McDowell, M. T.; Yao, Y.; Wang, C.; Cui, Y. A Yolk-Shell Design for Stabilized and Scalable. Nano Lett. 2012, 12, 3315-3321.

(28)

Zhang, L.; Hao, W.; Wang, H.; Zhang, L.; Feng, X.; Zhang, Y.; Chen, W.; Pang, H.; Zheng, H. Porous Graphene Frame Supported Silicon@graphitic Carbon via in Situ SolidState Synthesis for High-Performance Lithium-Ion Anodes. J. Mater. Chem. A 2013, 1, 7601-7611.

(29)

Yu, W. J.; Liu, C.; Hou, P. X.; Zhang, L.; Shan, X. Y.; Li, F.; Cheng, H. M. Lithiation of Silicon Nanoparticles Confined in Carbon Nanotubes. ACS Nano 2015, 9, 5063–5071.

(30)

Zhou, J.; Qian, T.; Wang, M.; Xu, N.; Zhang, Q.; Li, Q.; Yan, C. Core-Shell Coating Silicon Anode Interfaces with Coordination Complex for Stable Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 5358–5365.

(31)

Lu, C.; Fan, Y.; Li, H.; Yang, Y.; Tay, B. K.; Teo, E.; Zhang, Q. Core-Shell CNT-Ni-Si Nanowires as a High Performance Anode Material for Lithium Ion Batteries. Carbon 2013, 63, 54–60.

(32)

Suresh, S.; Wu, Z. P.; Bartolucci, S. F.; Basu, S.; Mukherjee, R.; Gupta, T.; Hundekar, P.; Shi, Y.; Lu, T.-M.; Koratkar, N. Protecting Silicon Film Anodes in Lithium-Ion Batteries Using an Atomically Thin Graphene Drape. ACS Nano 2017, 11, 5051-5061.

(33)

Xia, F.; Kwon, S.; Lee, W.; Liu, Z.; Kim, S.; Song, T.; Choi, K. J.; Paik, U.; Park, W. I.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Graphene as an Interfacial Layer for Improving Cycling Performance of Si Nanowires in Lithium-Ion Batteries. Nano Lett. 2015, 15, 6658–6664. (34)

Jung, H.; Lee, M.; Yeo, B. C.; Lee, K. R.; Han, S. S. Atomistic Observation of the Lithiation and Delithiation Behaviors of Silicon Nanowires Using Reactive Molecular Dynamics Simulations. J. Phys. Chem. C 2015, 119, 3447–3455.

(35)

Ostadhossein, A.; Cubuk, E. D.; Tritsaris, G. A.; Kaxiras, E.; Zhang, S.; van Duin, A. C. T. Stress Effects on the Initial Lithiation of Crystalline Silicon Nanowires: Reactive Molecular Dynamics Simulations Using ReaxFF. Phys. Chem. Chem. Phys. 2015, 17, 3832–3840.

(36)

Fan, F.; Huang, S.; Yang, H.; Raju, M.; Datta, D.; Shenoy, V. B.; van Duin, A. C. T.; Zhang, S.; Zhu, T. Mechanical Properties of Amorphous Li X Si Alloys: A Reactive Force Field Study. Model. Simul. Mater. Sci. Eng. 2013, 21, 1-15.

(37)

Chan, Y.; Hill, J. M. Lithium Ion Storage between Graphenes. Nanoscale Res. Lett. 2011, 6, 1-6.

(38)

Huang, S.; Zhu, T. Atomistic Mechanisms of Lithium Insertion in Amorphous Silicon. J. Power Sources 2011, 196, 3664–3668.

(39)

Stournara, M. E.; Xiao, X.; Qi, Y.; Johari, P.; Lu, P.; Sheldon, B. W.; Gao, H.; Shenoy, V. B. Li Segregation Induces Structure and Strength Changes at the Amorphous Si/Cu Interface. Nano Lett. 2013, 13, 4759–4768.

(40)

Ceperley, D. M.; Alder, B. J. The Ground State of the Electron Gas by a Stochastic Method. Phys. Rev. Lett. 1980, 45, 566-569.

(41)

Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

(42)

Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Density Functional for General Geometries. Phys. Rev. Lett. 2004, 92, 22–25.

(43)

Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van Der Waals Density Functional. J. Phys. Condens. Matter 2010, 22, 1-5.

(44)

Klimeš, J.; Bowler, D. R.; Michaelides, A. Van Der Waals Density Functionals Applied to Solids. Phys. Rev. B 2011, 83, 1–13.

(45)

Klimeš, J.; Michaelides, A. Perspective: Advances and Challenges in Treating van Der Waals Dispersion Forces in Density Functional Theory. J. Chem. Phys. 2012, 137, 1-12.

ACS Paragon Plus Environment

Page 20 of 21

Page 21 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

TABLE OF CONTENTS GRAPHIC

ACS Paragon Plus Environment