Amorphous Tin Oxide Nanohelix Structure Based Electrode for Highly

7 days ago - Vertically aligned individual NHs that are laterally separated from each other with equidistant empty space and not irregularly tangled u...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV AUTONOMA DE COAHUILA UADEC

Article

Amorphous Tin Oxide Nanohelix Structures Based Electrode for Highly Reversible Na-Ion Batteries Il Yong Choi, Changshin Jo, Won-Gwang Lim, Jong-Chan Han, Byeong-Gyu Chae, Chan Gyung Park, Jinwoo Lee, and Jong Kyu Kim ACS Nano, Just Accepted Manuscript • Publication Date (Web): 09 May 2019 Downloaded from http://pubs.acs.org on May 9, 2019

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 32 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 Nano

Amorphous Tin Oxide Nanohelix Structures Based Electrode for Highly Reversible Na-Ion Batteries Il Yong Choi,†,§ Changshin Jo,‡,||,§ Won-Gwang Lim,‡,⊥ Jong-Chan Han,† Byeong-Gyu Chae,† Chan Gyung Park,† Jinwoo Lee,*,⊥ and Jong Kyu Kim*,†



Department of Materials Science and Engineering, Pohang University of Science and

Technology (POSTECH), Pohang 37673, Republic of Korea ‡

Department of Chemical Engineering, Pohang University of Science and Technology

(POSTECH), Pohang 37673, Republic of Korea ⊥

Department of Chemical and Biomolecular Engineering, Korea Advanced Institute of Science

and Technology (KAIST), Daejeon 34141, Republic of Korea

ACS Paragon Plus Environment

1

ACS Nano 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

Page 2 of 32

Abstract

An array of amorphous tin oxide (a-SnOx) nanohelixes (NHs) was fabricated on copper foil as an electrode for Na-ion batteries via oblique angle deposition method, solution-and-surfactantfree process. The combination of the amorphous phase SnOx with a low oxidation number and its vertically-aligned NH geometry with a large surface area and high porosity, which facilitate Naion dynamics and accommodate the volume changes, enabled a reversible capacity ranging up to 915 mA h g-1 after 50 cycles, fast rate capability with 48.1% retention at 2 A g-1, and high stability, which are superior to those of crystalline nanoparticle-based electrodes.

Keywords amorphous tin oxide, nanohelix structures, oblique angle deposition, high reversibility, sodiumion batteries

ACS Paragon Plus Environment

2

Page 3 of 32 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 Nano

Sodium-ion batteries (SIBs) have been identified as promising alternatives to lithium-ion batteries (LIBs) owing to the shortage and high price of lithium (Li) as well as the earthabundance and low cost of sodium (Na).1,2 In order to demonstrate the economic and technological feasibility of SIBs, an improvement in the performance of both the anode and cathode is necessary because the energy and power densities of SIBs are not yet comparable with those of commercial LIBs. In particular, a lack of anode material with appropriate micro/nanostructure for sufficient Na storage, a large reversible capacity, and high structural stability remains an obstacle in developing high performance SIBs. For example, graphite, a typical anode material used in LIBs, has shown insufficient Na insertion/extraction capacities when used in SIBs.3–6 Therefore, researchers have made substantial efforts to find new materials with appropriate micro/nanostructures for highly reversible and stable anodes for SIBs. Tin oxide (SnOx) has been considered as a promising anode material for SIBs owing to its earth-abundance, low cost as well as a low operation potential.7,8 Theoretically, tin dioxide (SnO2) can deliver a high capacity up to 1378 mA h g-1, considering its conversion as well as alloying reactions with Na ions.9 Although many studies on tin oxides as anode materials for SIBs have been reported, there are several issues that should be considered and addressed. First, it is important to improve the reversibility of the electrochemical reaction between tin oxide and Na ions, which involves conversion and alloying reactions.10,11 Thus, tin oxides with low crystallinity, or even amorphous phase, would be beneficial for the insertion of Na ions into the electrode, preferentially through extended defects, resulting in an easy initiation of the breaking and rearrangement of tin (Sn) and oxygen (O) bonds.12–14 Despite such potential advantages of an amorphous anode, it is difficult to synthesize amorphous tin oxides by conventional solutionbased or surfactant-combined processes that require thermal annealing to remove remnant

ACS Paragon Plus Environment

3

ACS Nano 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

Page 4 of 32

solvents and organic surfactants.15,16 Furthermore, high capacity anodes for SIBs suffer from severe volume expansion and hence structural degradation during repetitive Na ion insertion/extraction reactions.9,17–20 Previous studies have reported improved cycle-life and reversibility of anodes by applying nanostructured materials such as randomly distributed nanoparticles or nanowires, however, these structures require large amounts of a polymeric binder and carbon additive for maintaining the electrode integrity and electrical contact, which degrades both the volumetric and gravimetric capacities of the electrode.20,21 Consequently, the development of an amorphous tin oxides-based anode with an appropriate nanostructure for effectively accommodating the volume changes during the charge-discharge processes via a surfactant-free fabrication process is strongly desired for the technological advancement of SIBs. In this study, we present an array of amorphous tin oxide (a-SnOx) nanohelixes (NHs) on copper (Cu) foil fabricated by an oblique angle deposition (OAD) method as an efficient anode for SIBs. Compared to electrodes based on crystalline tin dioxide (c-SnO2) and tin monoxide (cSnO) nanoparticles (NPs), a-SnOx NH with high porosity and high aspect ratio shows superior electrochemical performance, including reversible specific capacity and stability even without the use of the carbon additive and polymeric binder.

ACS Paragon Plus Environment

4

Page 5 of 32 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 Nano

Results/Discussion Synthesis of Amorphous SnOx Nanohelixes on Cu Foil An array of SnOx NH was fabricated on Cu foil by OAD using an electron beam evaporator. During OAD, the vapor flux of SnOx has an incident angle (θ) with respect to the substrate surface-normal direction by tilting the substrate, as schematically demonstrated in Figure 1a. At the initial stage of deposition, nanoscale nuclei form on the substrate, creating a self-shadowed region where subsequent vapor flux can no longer deposit. (Figure 1b) With the progression of the deposition process, the SnOx vapor deposits preferentially on top of the nuclei, forming a porous film composed of an array of slanted nanorods when the tilted substrate is held static or vertically aligned NHs when it is rotated. By adjusting the rotation speed and the deposition rate, one can control the geometrical shape of the NHs from vertically aligned nanorods to spring-like NHs with different diameters. Figure 1c shows a schematic illustration of the changes occurring in the structure of SnOx NHs after charge and discharge processes. Note that individual NHs in the vertically aligned array are laterally equidistant from each other and have empty spaces, which can enable great tolerance for the volume changes during the charge/discharge processes. The a-SnOx NH materials achieved better electrochemical performance compared with crystalized materials with same nanostructure (see further discussion). Therefore, in order to investigate the reaction mechanism, commercially available tin oxides, crystalline SnO2 (c-SnO2) NPs and crystalline SnO (c-SnO) NPs were also fabricated to serve as control electrodes. Figure 2a shows a scanning electron micrograph (SEM) of an array of SnOx NHs. Vertically aligned array consisting of 6-turn NHs with a high aspect ratio and large surface area is wellformed on the Cu substrate, with direct contact between the NHs and the current collector. This

ACS Paragon Plus Environment

5

ACS Nano 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

Page 6 of 32

contact can facilitate a more facile electron transport as compared with the electrodes obtained by conventional coating of a slurry of NPs on the substrate. Vertically aligned individual NHs which are laterally separated from each other with equidistant empty space and not irregularly tangled up, which can enable facilitated diffusion of Na ions, promoting the reaction with large surface area of NHs, as well as substantial tolerance for isotropic volume changes during the charge/discharge processes. In addition, spring-like NH shape would be beneficial for mechanical stability during repeated operations.22 The empty spaces between the laterally equidistant NHs can be very beneficial not only for the diffusion of Na ions but also for accommodating the volume changes during charging/discharging. High-resolution transmission electron micrograph (HR-TEM) shown in Figure 2b and the selected area electron diffraction pattern acquired from the region enclosed in a yellow box reveal the amorphous nature of the SnOx NHs formed by OAD via electron-beam evaporation at room temperature.23 The X-ray diffraction (XRD) pattern of the as-deposited SnOx NHs on the Cu foil (Figure 2c) shows only a sharp peak of Cu (200) (JCPDS No. 04-0836) from the Cu substrate, indicating that the asdeposited SnOx NHs are amorphous. In contrast, the c-SnO2 NP and c-SnO NP electrodes show a number of XRD peaks (JCPDS No. 41-1445 and No. 06-0395, respectively) from the crystalline phase of SnO2 and SnO, respectively. The Raman spectra in Figure 2d show Raman peaks at 476.5 cm-1 (Eg), 634 cm-1 (A1g), and 775 cm-1 (B2g) for c-SnO2 and at 211 cm-1 (A1g) for c-SnO, whereas a very broad peak ranging from 300 to 800 cm-1 appears for the a-SnOx sample.24 The TEM, XRD, and Raman spectroscopic analyses confirm that the SnOx NHs grown on Cu foil by OAD are amorphous. Note that, with bottom-up approaches including solution coating of NPs, obtaining an amorphous phase metal oxide thin-film is difficult because a high-temperature annealing process is necessary for removing unwanted materials, such as solvents, surfactants, or

ACS Paragon Plus Environment

6

Page 7 of 32 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 Nano

structure directing agents, incorporated during the synthetic process. In contrast, with OAD, a physical vapor deposition process using an electron beam evaporator, the array of amorphous SnOx NHs can be easily fabricated directly on virtually any substrate at room temperature in a simple and easy way that is compatible with current micro-electronics fabrication processes. Further, atomic-selective X-ray absorption spectroscopy of the tin oxide samples was performed using the 10C beam line of the Pohang Accelerator Laboratory. In the extended X-ray absorption fine structure (EXAFS) profiles shown in Figure 3a, the differences in the Fourier transform (FT) magnitudes in the second shell interaction, which is dominated by the Sn-Sn interaction (ranging from a radial distance of 3 to 4 Å), between c-SnO2, c-SnO, and a-SnOx clearly shows the existence of anisotropy in the a-SnOx owing to a lower local structural symmetry, further confirming the amorphous nature of the a-SnOx NHs.12,25 The oxidation state of the transition metal in a transition metal oxide affects the electrochemical reaction for Na ion storages, and therefore, can play an important role in the electrical conductivity, reversibility, and stability of SIBs. The average oxidation number of Sn atoms in the a-SnOx NH electrode is estimated to be 2.2 (Figure S1), which is between those of c-SnO2 and c-SnO, based on the X-ray absorption near edge structure (XANES) profiles shown in Figure 3b.26,27 The oxidation state of Sn in a-SnOx NHs was determined by correlating the oxidation states of Sn metal, c-SnO, and c-SnO2 references and their edge energies, which were decided based on the points at which the second derivatives are zero in the XANES profiles.26,27 Metal oxides formed by physical vapor deposition under a high vacuum condition (an oxygendeficient condition) typically have a lower oxidation number. It has been reported that such reduced form of transition metal oxides shows enhanced electrical conductivity and high electrochemical activity, resulting in a low electrode resistance, high reversibility, and high

ACS Paragon Plus Environment

7

ACS Nano 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

Page 8 of 32

capacity.28–30 Therefore, the combination of the microstructural properties, i.e., the amorphous phase with a low oxidation number and the geometrical structure with a large surface area, high porosity for facilitating Na-ion dynamics and accommodating the volume changes, enabled by the array of a-SnOx NHs is expected to result in highly enhanced electrochemical performance of the SIBs.

Electrochemical Performance of a-SnOx NH Electrodes Cyclic voltammetry (CV) measurements of the a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes were performed to investigate their electrochemical performances and the relevant mechanism of each electrode. The CV profiles of the different electrodes show quite different discharge-charge characteristics. The a-SnOx NH electrode exhibits an intense cathodic peak at 0.65 V at the first discharge (sodiation) process. (Figure 4a) This peak is related to the initiation of the conversion reaction and the formation of a solid-electrolyte interphase (SEI) along the electrode surface, which will be discussed with the EXAFS analysis. A cathodic peak related to formation of Sn-Na alloy is observed at the end of the cathodic scan. During the charging (desodiation) process, there are two distinct peaks at 0.27 V and 0.95 V, respectively, related to the de-alloying and re-oxidation of Sn to SnOx phase. As the discharge-charge process continues, the cathodic and the anodic peaks get closer and reach a much higher reversible electrochemical reaction. However, the c-SnO2 NP electrode shows broad peaks during both discharging and charging processes. (Figure 4b) Although reversible conversion and alloying reactions occur in the c-SnO2 NP electrode in the initial cycle, the cathodic and anodic (re-oxidation at 1.0 V) peaks suffer from continuous decline in the intensity, indicating that the sodiation/desodiation

ACS Paragon Plus Environment

8

Page 9 of 32 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 Nano

processes are not occurring effectively in the c-SnO2 NP electrode. Owing to the lower oxidation state, the sodiation process in the c-SnO electrode (see CV profile in Figure 4c) begins at a lower potential compared with that of the c-SnO2 electrode. The c-SnO NP electrode exhibits a similar profile as the a-SnOx NH electrode during the desodiation processes. However, peaks at ~1.0 V for the c-SnO NP electrode during the charging processes are not stable with cycling. The reason for the enhanced reversibility of all the redox peaks in the CV profile of the a-SnOx NH electrode, compared with those of the crystalline electrodes, is probably due to the larger amount of defects including vacancies in the amorphous phase.31 Galvanostatic discharge-charge curves of the a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes, shown in Figure 4d, 4e, and 4f, respectively, were obtained at 20 mA g-1 current density in the potential range of 2.5–0.002 V (vs. Na/Na+). During the discharge process, the aSnOx NH electrode shows a plateau at 1.0 V, followed by a sloped curve. During the charging process, it exhibits two capacity contributions in the 0.1–0.3 V and 0.7–1.3 V regions. The capacity of the 0.1–0.3 V region gradually increases with cycling, as shown in Figure 4a. Discharge and charge capacities in the first cycle for the a-SnOx NH electrode are 1414 and 751 mA h g-1, respectively (initial Coulombic efficiency (ICE) value is 53.1%). Compared with the aSnOx NH, the c-SnO2 NP electrode shows a quite different discharge-charge profile including a steep decrease in the potential, a sloped curve up to 0.1 V, and a plateau, as shown in Figure 4b. On one hand, the discharge and charge capacities in the first cycle for the c-SnO2 NP electrode are 537 and 124 mA h g-1, respectively (ICE: 23.1%). On the other hand, the Na ion insertion plateau at 0.4–0.25 V contributes to most of the discharge capacity of the c-SnO NP electrode. The c-SnO NP electrode shows similar discharge-charge profiles as the a-SnOx after the initial cycle (531 and 327 mA h g-1, ICE: 61.6%) and its reversible capacities increase with the number

ACS Paragon Plus Environment

9

ACS Nano 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

Page 10 of 32

of cycles, consistent with the CV results. The relatively low ICE of the a-SnOx NH electrode is attributed to the formation of an SEI layer on the NH structures with a large surface area. In order to trace the chemical reaction of the three electrodes during the discharge-charge processes, Sn K-edge EXAFS profiles of the a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes were analyzed ex-situ at each key stage of the reaction. (Figure 5a, 5d, and 5g) The a-SnOx NH electrode shows Sn-O peak at the open-circuit voltage (OCV). During the discharge process, the Sn-O peak declines and the Sn-Sn peak grows, which is the evidence of the conversion reaction. (Figure 5b) This Sn-Sn peak becomes bigger at fully sodiated state, indicating further conversion reaction process occur at lower potential. There is no direct evidence of the formation of Na-Sn alloy during the sodiation process in the EXAFS profiles. This is likely caused by competitive processes between the formation of Sn metal by the conversion reaction and the formation of NaSn alloy phase, decreasing the Sn-Sn peak. (see TEM data in Figure S2) Yoon and co-workers reported that SnO2 with Li+ reaction makes growing Sn-Sn peaks during conversion process while the intensity of the peak decreases during further lithiation (alloying) process.8 During the charging process, there is a small difference in the intensity of the Sn-O and Sn-Sn peaks at 0.002 V (point 2) and 0.5 V (point 3) (Figure 5c) despite the contribution of a high capacity, ~290 mA h g-1 (Figure 5a), which is correlated to the de-alloying process. The Sn-O peak recovers whereas the Sn-Sn peak decreases during the charging process, with the EXAFS profile becoming similar to that of the OCV state, suggesting a highly reversible dischargecharge process. In the EXAFS profiles of the c-SnO2 NP electrode (Figure 5e and 5f), the first and second Sn-O peaks decrease during the discharging process, but the Sn-Sn peak in the profiles is negligible. Although the Sn-O bond is partially recovered, this reaction had low reversibility, as confirmed in the discharge-charge curves of the c-SnO2 NP electrode. (Figure

ACS Paragon Plus Environment

10

Page 11 of 32 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 Nano

5c) The disappearance of the second Sn-O peak after a full discharge-charge cycle indicates that the highly crystalline grain structure could not be recovered very well. In contrast, the EXAFS profiles of the c-SnO NP electrode (Figure 5h and 5i) exhibits a similar trend as those of the aSnOx NH electrode. The results imply that the low oxidation state of Sn atoms in tin oxide, which can make less Na2O phase, is required for both conversion and alloying reactions to achieve a highly efficient Na-ion storage. The ex-situ EXAFS results showing the conversion and alloying reactions of a-SnOx with Na ions are further supported by TEM and XRD analyses. (Figure S2 and S3) Selective area electron diffraction (SAED) patterns of the a-SnOx NH electrode in Figure S2b show the existence of Na2O, Sn and Na15Sn4 when a-SnOx NH electrode is discharged to 0.35 V, which indicates the conversion and alloying reactions of the a-SnOx with Na ions. The SAED patterns of the a-SnOx NH electrode charged to 2.5 V in Figure S2d mainly show Sn and SnO phases. In the XRD study (Figure S3b), we found that amorphous SnOx was converted to Sn metallic phase after full discharge and converted to Sn and SnO phases after charge process, which is coincident with the SAED pattern of a-SnOx NH electrode in Figure S2d. In order to investigate the effect of the amorphous phase on electrochemical performance, the a-SnOx NH was annealed under Ar (inert) ambient and H2/Ar (reductive) ambient at 350 oC for 2 h, respectively. Discharge-charge curves and cycle life of the SIBs with the a-SnOx NH electrode and the annealed a-SnOx NH electrode are shown in Figure S4. The annealed a-SnOx NH electrode shows lower reversible capacities than bare a-SnOx NH electrode, indicating that the amorphous phase is more advantageous than the crystalline phase for electrochemical reaction of tin oxide with Na ions. The a-SnOx NH electrode annealed under H2/Ar ambient exhibits improved capacity compared with the same material annealed under Ar ambient, which may

ACS Paragon Plus Environment

11

ACS Nano 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

Page 12 of 32

imply that crystalline Sn based oxide with low oxidation state has better electrochemical reaction with Na ions. It is coincident with better electrochemical performance of c-SnO NP electrode than that of c-SnO2 NP electrode as shown in Figure 6a. On the other hand, in order to investigate the effect of geometrical shape, NHs structure, on the electrochemical performance, a-SnOx thin-film electrode, i.e., non-nanostructured electrode, was prepared by electron beam evaporation. The thin-film a-SnOx showed a similar initial sodiation profile but a poor reversibility compared to the a-SnOx NH electrode as shown in Figure S5. The capacity contribution from re-oxidation process (> 0.5 V) disappeared from second cycle and de-alloying capacity (< 0.5 V) decreased with cycle numbers. This indicates that appropriate geometrical parameters of electrodes, such as porosity, aspect ratio, and surface area, are important factors for stable cycling of a-SnOx material. The electrochemical performances of the a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes, such as cycle-life and rate capability, were investigated. All electrodes were cycled at 25 mA g-1 for 5 cycles and then at 100 mA g-1 for further cycles. (Figure 6a) On one hand, the charge capacity of the a-SnOx NH electrode increased continuously with cycling, reaching 915 mA h g-1 after 50 cycles. The increment, as confirmed by the CV profiles in Figure 4a and dischargecharge curves in Figure 4d, is due to activation process of both the alloying and conversion processes, as shown in previous studies on high capacity anodes.32,33 On the other hand, the reversible capacities of the c-SnO2 NP and c-SnO NP electrodes decline with cycling: the charge capacities are 32 and 92 mA h g-1 after the 50th cycle, respectively. The capacity fading observed for the c-SnO2 NP and c-SnO NP electrodes may be attributed to the formation of large Na2O domains in the NP structures or the relatively large particle/crystal sizes of the crystalline materials, resulting in high electrical impedance.21 The rate capabilities of all the electrodes were

ACS Paragon Plus Environment

12

Page 13 of 32 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 Nano

investigated under different current densities. (Figure 6b) The a-SnOx NH electrode reaches a maximum capacity of 904 mA h g-1 at 40 mA g-1 and 428 mA h g-1 (48.1 % of the maximum capacity) is maintained at 2000 mA g-1. The SIB with the a-SnOx NH electrode has a reversible specific capacity (Figure 6c), more than 900 mA h g-1, which is an excellent electrochemical performance for SIBs with tin oxidebased anodes without any carbon additives.9,11,20,34–38 Such excellent electrochemical performance could be attributed to the combination of the following factors: i) intrinsic properties, i.e., an amorphous phase with a low oxidation state of a-SnOx deposited by e-beam evaporation under a high vacuum condition, which lowers the kinetic barrier for efficient electrochemical reactions and ii) the geometrical structure, i.e., vertically aligned NH array with a large surface area and high porosity obtained by OAD, which facilitates Na-ion dynamics and effectively accommodates the volume change. It is worth noting that such an excellent performance was obtained without using any additional carbon-based materials, and it was realized by a solution-free, organic surfactant-free fabrication process.

ACS Paragon Plus Environment

13

ACS Nano 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

Page 14 of 32

Conclusions An array of amorphous SnOx NH was fabricated directly on Cu foil by OAD using an electron beam evaporator at room temperature, which is not only a simple and scalable process compatible with micro-electronics, but also a solution-free and organic surfactant-free fabrication process. The amorphous phase of the SnOx NH with a low oxidation number and its geometrical structure with high porosity, high aspect ratio, and large surface area led to facile Na ion dynamics and accommodation of the volume change. The SIB with the amorphous SnOx NH electrode has superior reversible capacity, which continuously increases with cycling, reaching up to 915 mA h g-1 after 50 cycles, and is highly stable, thus exhibiting an excellent electrochemical performance compared to SIBs with tin oxide-based anodes.

ACS Paragon Plus Environment

14

Page 15 of 32 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 Nano

Figure 1. (a) Schematic illustration of the oblique angle deposition method using an electron beam evaporator. (b) Nuclei of the deposited material forms a self-shadowed region where subsequent incident vapor flux can no longer deposit, resulting in the formation of an array of slanted nanorods or vertically aligned NHs when the tilted substrate is rotated. (c) Schematic showing the changes occurring in the structure of SnOx NHs fabricated directly on a Cu substrate after discharging and charging processes. The empty spaces between the individual NHs that are laterally equidistant from each other enable great tolerance for the volume change during charging/discharging processes.

ACS Paragon Plus Environment

15

ACS Nano 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

Page 16 of 32

Figure 2. (a) SEM image of the array of SnOx NH structures. (b) HR-TEM image of a-SnOx NH and SAED pattern acquired from the yellow square region in (b). (c) XRD patterns and (d) Raman spectra of the as-deposited SnOx NH, c-SnO2 NP, and c-SnO NP electrodes. (scale bar: 1 μm for (a), 20 nm for (b))

ACS Paragon Plus Environment

16

Page 17 of 32 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 Nano

Figure 3. (a) EXAFS and (b) XANES profiles of the a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes, and Sn foil used as a reference.

ACS Paragon Plus Environment

17

ACS Nano 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

Page 18 of 32

Figure 4. (a, b, c) CV and (d, e, f) galvanostatic discharge-charge curves of SIBs with the aSnOx NH electrode, c-SnO2 NP electrode, and c-SnO NP electrode, respectively. (CV scan rate: 0.1 mV s-1, current density: 20 mA g-1)

ACS Paragon Plus Environment

18

Page 19 of 32 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 Nano

Figure 5. Discharge-charge curves for the first cycle: (a) a-SnOx NH, (d) c-SnO2 NP and (g) cSnO NP electrodes. (b,c) Sn K-edge EXAFS profiles of the a-SnOx NHs at key reaction stages during the discharge and charge processes. (e,f) Sn K-edge EXAFS profiles of the c-SnO2 NPs and (h,i) Sn K-edge EXAFS profiles of the c-SnO NP electrode for discharge and charge processes.

ACS Paragon Plus Environment

19

ACS Nano 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

Page 20 of 32

Figure 6. (a) Cycle-life (25 mA g-1 for the first 5 cycles and 100 mA g-1 for further cycles) and (b) rate capability profiles of the SIBs with a-SnOx NH, c-SnO2 NP, and c-SnO NP electrodes. (c) Comparison of the electrochemical performance of the electrode examined in this study with previously reported tin oxide-based anodes for SIBs.9,11,20,34-38

ACS Paragon Plus Environment

20

Page 21 of 32 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 Nano

Methods/Experimental Fabrication of a-SnOx NH electrode. Cu foil (20 um thickness) was first cleaned by sonicating in acetone and isopropyl alcohol for 10 min and rinsed in distilled water for 10 min. The array of a-SnOx NH structures was directly grown on the Cu foil by an OAD method using an electron-beam evaporator. SnO2 precursor powder (Taewon Scientific Co., Korea) was used for electron-beam deposition. In the OAD process, the deposition rate and base pressure were maintained at 2.5 Å s-1 and 7.0 × 10-7 Torr, respectively. The oblique angle with respect to the normal axis of the substrate was 80° and the substrate was rotated every 2 s at 1 rpm and stopped for 12 s. For comparison, a dense, non-nanostructured a-SnOx film electrode was also fabricated on the same Cu foil using the electron beam evaporator without tilting and rotating the substrate. The electrodes were pressed and cut into a round shape with 14 mm diameter. Fabrication of c-SnO2 NP and c-SnO NP electrodes. SnO2 NP and SnO NP electrodes were fabricated on carbon-coated aluminum (Al) foils by a conventional doctor blade method. c-SnO2 NP and c-SnO NP electrodes were prepared as control samples by mixing c-SnO2 and c-SnO NPs with conductive carbon (Super P) and a polymeric organic surfactant (carboxymethyl cellulose) at 8:1:1 weight ratio in DI water. The slurry was coated on the carbon-coated Al current collector and dried at 110 °C in a vacuum oven. The electrodes were pressed and cut into a round shape with 14mm diameter. Characterization. The micro-structure of the electrodes was characterized by field emission scanning electron microscopy (FE-SEM, Philips, XL30S FEG) and field emission transmission electron microscopy (FE-TEM, Jeol, JEM-ARM200F). XRD was carried out on a X-ray Diffractometer (Rigaku, D/MAX-2500) with Cu Kα radiation (λ = 1.5418 Å). Raman spectra were acquired using a Horiba Jobin Yvon LabRam Aramis with an Ar-ion laser beam at an

ACS Paragon Plus Environment

21

ACS Nano 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

Page 22 of 32

excitation wavelength of 514.5 nm. XAS data were collected at the nanoprobe XAFS beamline of the Pohang Light Source (PLS-II) in the 3.0 GeV storage ring, with a ring current of 360 mA. The X-ray beam was monochromatized by a Si (111) double crystal, whereby the beam intensity was reduced by 30% to eliminate the higher-order harmonics. The X-ray beam was then delivered to a secondary source aperture where the beam size was adjusted to be 0.3 nm (v) × 1 mm (h). A high voltage (3000 V) was applied to the ionization chambers filled with N2/Ar gas mixture to measure the X-ray intensity. XAFS analysis was conducted in the fluorescence mode and the obtained spectra were processed using the Demeter software. Electrochemical characterization. Electrochemical analysis of tin oxide materials was conducted using two-electrode coin cells (CR2032 type). As a working electrode, each a-SnOx NH, c-SnO2 NP, and c-SnO NP electrode was directly assembled with metallic sodium, which acts as both counter and reference electrodes. The electrodes were assembled in a glove box using an electrolyte (1.0 M sodium hexafluorophosphate (NaPF6) in ethylene carbonate (EC)/dimethyl carbonate (DMC) containing 5 wt% fluoroethylene carbonate (the volume ratio of EC:DMC = 1:1)) and a separator (GF/F glass microfiber filters). The mass loadings of a-SnOx NH, c-SnO2 NP, and c-SnO, NP electrodes were 0.2 to 0.32 mg cm-2, 1.9 mg cm-2, and 3.2 mg cm-2. The galvanostatic discharge-charge analysis in the range of current densities from 20 to 2000 mA g-1 and CV at 0.1 mV s-1 scan rate were performed in the range of 0.002 to 2.5 V (vs. Na/Na+). All electrochemical experiments were performed using a WBCS-3000 battery cycler (WonA Tech, Korea).

ACS Paragon Plus Environment

22

Page 23 of 32 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 Nano

ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: Figure S1 indicates the average oxidation number of Sn in the a-SnOx NH electrode estimated from the XANES profiles. Figure S2 shows TEM images and corresponding SAED patterns of aSnOx NH electrode at different states. Figure S3 shows XRD patterns of a-SnOx NH, c-SnO2 NP and c-SnO NP electrodes at certain states. Figure S4 shows discharge-charge curves and cycle performance of a-SnOx NH electrode and annealed a-SnOx NH electrode. Figure S5 shows discharge-charge curves of a-SnOx thin-film electrode with the same height as the a-SnOx NHs. (PDF)

ACS Paragon Plus Environment

23

ACS Nano 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

Page 24 of 32

AUTHOR INFORMATION Corresponding Author Jinwoo Lee* E-mail: [email protected] Jong Kyu Kim* E-mail: [email protected]

Present Addresses ||

Institute for Manufacturing, Department of Engineering, University of Cambridge, Cambridge,

United Kingdom

Author Contributions §

I.Y.C. and C.J. contributed equally to this work. I.Y.C., C.J., J.L. and J.K.K. conceived the idea

and designed the experiments. I.Y.C. performed sample fabrication and characterization. C.J. and W.-G.L. conducted the electrochemical studies. I.Y.C., C.J. and W.-G.L. analyzed the experimental data. I.Y.C., B.-G.C., C.-G.P. performed TEM analysis. All authors discussed the experimental results and contributed to the manuscript.

ACS Paragon Plus Environment

24

Page 25 of 32 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 Nano

ACKNOWLEDGMENT This work was supported by the Basic Research Lab (BRL) Project (NRF2017R1A4A1015811) and Basic Science Research Program (NRF-2017R1A2B3004648) through the National Research Foundation of Korea (NRF) funded by the Ministry of Education. This work was supported by NRF (National Research Foundation of Korea) Grant funded by the Korea government (MSIT) (NRF-2018R1A2B6005410). I.Y.C. acknowledges the support by the Global Ph.D. Fellowship through the NRF (NRF-2013H1A2A1034492).

ACS Paragon Plus Environment

25

ACS Nano 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

Page 26 of 32

REFERENCES (1) Xu, Y.; Zhou, M.; Lei, Y. Nanoarchitectured Array Electrodes for Rechargeable Lithium‐ and Sodium‐Ion Batteries. Adv. Energy Mater. 2016, 6, 1502514. (2) Vaalma, C.; Buchholz, D.; Weil, M.; Passerini, S. A Cost and Resource Analysis of SodiumIon Batteries. Nat. Rev. Mater. 2018, 3, 18013. (3) Doeff, M. M.; Ma, Y.; Visco, S. J.; Jonghe, L. C. D. Electrochemical Insertion of Sodium into Carbon. J. Electrochem. Soc. 1993, 140, 12, L169-L170. (4) Thomas, P.; Billaud, D. Electrochemical Insertion of Sodium into Hard Carbons. Electrochim. Acta 2002, 47, 3303-3307. (5) Jo, C.; Park, Y.; Jeong, J.; Lee, K. T.; Lee, J. Structural Effect on Electrochemical Performance of Ordered Porous Carbon Electrodes for Na-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 11748-11754. (6) Kim, H.; Hong, J.; Yoon, G.; Kim, H.; Park, K. -Y.; Park, M. -S.; Yoon, W. -S.; Kang, K. Sodium Intercalation Chemistry in Graphite. Energy Environ. Sci. 2015, 8, 2963-2969. (7) Kim, Y.; Ha, K. -H.; Oh, S. M.; Lee, K. T. High‐Capacity Anode Materials for Sodium‐Ion Batteries. Chem. Eur. J. 2014, 20, 11980-11992. (8) Kim, H.; Park, G. O.; Kim, Y.; Muhammad, S.; Yoo, J.; Balasubramanian, M.; Cho, Y. –H.; Kim, M. –G.; Lee, B.; Kang, K.; Kim, H.; Kim, J. M.; Yoon, W. –S. New Insight into the Reaction Mechanism for Exceptional Capacity of Ordered Mesoporous SnO2 Electrodes via Synchrotron-Based X-Ray Analysis. Chem. Mater. 2014, 26, 6361-6370.

ACS Paragon Plus Environment

26

Page 27 of 32 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 Nano

(9) Ao, X.; Jiang, J.; Ruan, Y.; Li, Z.; Zhang, Y.; Sun, J.; Wang, C. Honeycomb-Inspired Design of Ultrafine SnO2@C Nanospheres Embedded in Carbon Film as Anode Materials for High Performance Lithium- and Sodium-Ion Battery. J. Power Sources 2017, 359, 340-348. (10) Cui, J.; Xu, Z. –L.; Yao, S.; Huang, J.; Huang, J. –Q.; Abouali, S.; Garakani, M. A.; Ning, X.; Kim, J. –K. Enhanced Conversion Reaction Kinetics in Low Crystallinity SnO2/CNT Anodes for Na-Ion Batteries. J. Mater. Chem. A 2016, 4, 10964-10973. (11) Fan, L.; Li, X.; Yan, B.; Feng, J.; Xiong, D.; Li, D.; Gu, L.; Wen, Y.; Lawes, S.; Sun, X. Controlled SnO2 Crystallinity Effectively Dominating Sodium Storage Performance. Adv. Energy Mater. 2016, 6, 1502057. (12) Ku, J. H.; Ryu, J. H.; Kim, S. H.; Han, O. H.; Oh, S. M. Reversible Lithium Storage with High Mobility at Structural Defects in Amorphous Molybdenum Dioxide Electrode. Adv. Funct. Mater. 2012, 22, 3658–3664. (13) Chae, O. B.; Kim, J.; Park, I.; Jeong, H.; Ku, J. H.; Ryu, J. H.; Kang, K.; Oh, S. M. Reversible Lithium Storage at Highly Populated Vacant Sites in an Amorphous Vanadium Pentoxide Electrode. Chem. Mater. 2014, 26, 5874−5881. (14) Ren, Y.; Armstrong, A. R.; Jiao, F.; Bruce, P. G. Influence of Size on the Rate of Mesoporous Electrodes for Lithium Batteries. J. Am. Chem. Soc. 2010, 132, 996–1004. (15) Kim, G.; Jo, C.; Kim, W.; Chun, J.; Yoon, S.; Lee, J.; Choi, W. TiO2 Nanodisks Designed for Li-Ion Batteries: a Novel Strategy for Obtaining an Ultrathin and High Surface Area Anode Material at the Ice Interface. Energy Environ. Sci. 2013, 6, 2932-2938.

ACS Paragon Plus Environment

27

ACS Nano 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

Page 28 of 32

(16) Kim, J.; Cho, J. Rate Characteristics of Anatase TiO2 Nanotubes and Nanorods for Lithium Battery Anode Materials at Room Temperature. J. Electrochem. Soc. 2007, 154, A542-A546. (17) Bian, H.; Zhang, J.; Yuen, M.-F.; Kang, W.; Zhan, Y.; Yu, D. Y. W.; Xu, Z.; Li, Y. Y. Anodic Nanoporous SnO2 Grown on Cu Foils as Superior Binder-Free Na-Ion Battery Anodes. J. Power Sources 2016, 307, 634-640. (18) Liu, Y,; Fang, X,; Ge, M,; Rong, J,; Shen, C,; Zhang, A,; Enaya, H. A.; Zhou, C. SnO2 Coated Carbon Cloth with Surface Modification as Na-Ion Battery Anode. Nano Energy 2015, 16, 399–407. (19) Dirican, M.; Lu, Y.; Ge, Y.; Yildiz, O.; Zhang, X. Carbon-Confined SnO2-Electrodeposited Porous Carbon Nanofiber Composite as High-Capacity Sodium-Ion Battery Anode Material. ACS Appl. Mater. Interfaces 2015, 7, 18387−18396. (20) Su, D.; Wang, C.; Ahn, H.; Wang, G. Octahedral Tin Dioxide Nanocrystals as High Capacity Anode Materials for Na-Ion Batteries. Phys. Chem. Chem. Phys. 2013, 15, 1254312550. (21) Gu, M.; Kushima, A.; Shao, Y.; Zhang, J. –G.; Liu, J.; Browning, N. D.; Li J.; Wang, C. Probing the Failure Mechanism of SnO2 Nanowires for Sodium-Ion Batteries. Nano Lett. 2013, 13, 5203-5211. (22) Chun, S.; Choi, I. Y.; Son, W. K.; Bae, G. Y.; Lee, E. J.; Kwon, H.; Jung, J.; Kim, H. S.; Kim, J. K.; Park, W. A Highly Sensitive Force Sensor with Fast Response Based on Interlocked Arrays of Indium Tin Oxide Nanosprings toward Human Tactile Perception. Adv. Funct. Mater. 2018, 28, 1804132.

ACS Paragon Plus Environment

28

Page 29 of 32 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 Nano

(23) Singh, J. P.; Karabacak, T.; Ye, D. -X.; Liu, D. -L.; Picu, C.; Lu, T. -M.; Wang, G. -C. Physical Properties of Nanostructures Grown by Oblique Angle Deposition. J. Vac. Sci. Technol. B 2005, 23, 2114-2121. (24) Vijayarangamuthu, K.; Rath, S. Nanoparticle Size, Oxidation State, and Sensing Response of Tin Oxide Nanopowders Using Raman Spectroscopy. J. Alloys Compd. 2014, 610, 706-712. (25) Zhu, Q.; Ma, Q.; Buchholz, D. B.; Chang, R. P. H.; Bedzyk M. J.; Mason, T. O. Structural Anisotropy in Amorphous SnO2 Film Probed by X-Ray Absorption Spectroscopy. Appl. Phys. Lett. 2013, 103, 031913. (26) Kau, L. –S.; Spira-Solomon, D. J.; Penner-Hahn, J. E.; Hodgson, K. O.; Solomon, E. I. XRay Absorption Edge Determination of the Oxidation State and Coordination Number of Copper. Application to the Type 3 Site in Rhus Vernicifera Laccase and Its Reaction with Oxygen. J. Am. Chem. Soc. 1987, 109, 6433-6442. (27) Dau, H.; Liebisch, P.; Haumann, M. X-Ray Absorption Spectroscopy to Analyze Nuclear Geometry and Electronic Structure of Biological Metal Centers—Potential and Questions Examined with Special Focus on the Tetra-Nuclear Manganese Complex of Oxygenic Photosynthesis. Anal. Bioanal. Chem. 2003, 376, 562-583. (28) Hwnag, J.; Jo, C.; Kim, M. G.; Chun, J.; Lim E.; Kim, S.; Jeong, S.; Kim, Y.; Lee, J. Mesoporous Ge/GeO2/Carbon Lithium-Ion Battery Anodes with High Capacity and High Reversibility. ACS Nano 2015, 9, 5, 5299-5309. (29) Yoon, S.; Jo, C.; Noh, S. Y.; Lee, C. W.; Song, J. H.; Lee, J. Development of a HighPerformance Anode for Lithium Ion Batteries Using Novel Ordered Mesoporous Tungsten

ACS Paragon Plus Environment

29

ACS Nano 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

Page 30 of 32

Oxide Materials with High Electrical Conductivity. Phys. Chem. Chem. Phys. 2011, 13, 1106011066. (30) Jo, C.; Hwnag, I.; Lee, J.; Lee, C.; Yoon, S. Investigation of Pseudocapacitive ChargeStorage Behavior in Highly Conductive Ordered Mesoporous Tungsten Oxide Electrodes. J. Phys. Chem. C 2011, 115, 11880-11886. (31) Xu, Y.; Zhou, M.; Zhang, C.; Wang, C.; Liang, L.; Fang, Y.; Wu, M.; Cheng, L.; Lei, Y. Oxygen Vacancies: Effective Strategy to Boost Sodium Storage of Amorphous Electrode Materials. Nano Energy 2017, 38, 304-312. (32) Zhang, Y.; Xie, J.; Zhang, S.; Zhu, P.; Cao, G.; Zhao, X. Ultrafine Tin Oxide on Reduced Graphene Oxide as High-Performance Anode for Sodium-Ion Batteries. Electrochim. Acta 2015, 151, 8-15. (33) Zhao, X.; Zhang, Z.; Yang, F.; Fu, Y.; Lai, Y.; Li, J. Core–Shell Structured SnO2 Hollow Spheres–Polyaniline Composite as an Anode for Sodium-Ion Batteries. RSC Advances 2015, 5, 31465. (34) Zhang, F.; Zhu, J.; Zhang, D.; Schwingenschlögl, U.; Alshareef, H. N. Two-Dimensional SnO Anodes with a Tunable Number of Atomic Layers for Sodium Ion Batteries. Nano Lett. 2017, 17, 1302-1311. (35) Zheng, Y.; Zhou, T.; Zhang, C.; Mao, J.; Liu, H.; Guo, Z. Boosted Charge Transfer in SnS/SnO2 Heterostructures: Toward High Rate Capability for Sodium‐Ion Batteries. Angew. Chem. Int. Ed. 2016, 55, 3408-3413.

ACS Paragon Plus Environment

30

Page 31 of 32 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 Nano

(36) Lu, Y. C.; Ma, C.; Alvarado, J.; Kidera, T.; Dimov, N.; Meng, Y. S.; Okada, S. Electrochemical Properties of Tin Oxide Anodes for Sodium-Ion Batteries. J. Power Sources 2015, 284, 287-295. (37) Patra, J.; Chen, H. –C.; Yang, C. –H.; Hsieh, C. –T.; Su, C. –Y.; Chang, J. –K. High Dispersion of 1-nm SnO2 Particles between Graphene Nanosheets Constructed Using Supercritical CO2 Fluid for Sodium-Ion Battery Anodes. Nano Energy 2016, 28, 124-134. (38) Liu, Z.; Song, T.; Kim, J. H.; Li, Z.; Xiang, J.; Lu, T.; Paik, U. Partially Reduced SnO2 Nanoparticles Anchored on Carbon Nanofibers for High Performance Sodium-Ion Batteries. Electrochem. Commun. 2016, 72, 91-95.

ACS Paragon Plus Environment

31

ACS Nano 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

Page 32 of 32

For Table of Contents Only

ACS Paragon Plus Environment

32