Superior Sodium Storage of Carbon-coated NaV6O15 Nanotubes

Feb 25, 2019 - Xuexia Song , Jicheng Li , Zhaohui Li , Qizhen Xiao , Gangtie Lei , Zhongliang Hu , Yanhuai Ding , Hirbod Maleki Kheimeh Sari , and Xif...
0 downloads 0 Views 1MB Size
Subscriber access provided by Washington University | Libraries

Energy, Environmental, and Catalysis Applications

Superior Sodium Storage of Carbon-coated NaV6O15 Nanotubes Cathode: Pseudocapacitance vs. Intercalation Xuexia Song, Jicheng Li, Zhaohui Li, Qizhen Xiao, Gangtie Lei, Zhongliang Hu, Yanhuai Ding, Hirbod Maleki Kheimeh Sari, and Xifei Li ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b20494 • Publication Date (Web): 25 Feb 2019 Downloaded from http://pubs.acs.org on February 26, 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 33 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

Graphic abstract

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

Superior Sodium Storage of Carbon-coated NaV6O15 Nanotubes Cathode: Pseudocapacitance vs. Intercalation Xuexia Songa,c,d,1, Jicheng Lia,1, Zhaohui Lia*, Qizhen Xiaoa, Gangtie Leia, Zhongliang Hub*, Yanhuai Dinge, Hirbod Maleki Kheimeh Saric,d, Xifei Lic,d* a

Key Laboratory of Environmentally Friendly Chemistry and Applications of Ministry of

Education, College of Chemistry, Xiangtan University, Hunan 411105, P.R. China b

College of Metallurgy and Material Engineering, Hunan University of Technology, Hunan 412007, P.R. China c

Institute of Advanced Electrochemical Energy & School of Materials Science and

Engineering, Xi'an University of Technology, Xi'an, Shaanxi 710048, P.R. China d

Shaanxi International Joint Research Centre of Surface Technology for Energy Storage Materials, Xi'an, Shaanxi 710048, China e

Institute of Fundamental Mechanics and Materials Engineering, Xiangtan University, Hunan 411105, P.R. China

ABSTRACT To realize the effect of Na+ pseudocapacitance on sodium storage of cathode materials, clew­like carbon­coated sodium vanadium bronze (NaV6O15) nanotubes (Na­VBNT@C) were synthesized via a facile combined sol­gel/hydrothermal method. The resultant Na­VBNT@C delivers high reversible capacities of 209 and 105 mAh g­1 at the rates of 0.1C and 10C, respectively. Notably, at the higher rate of 5C (1250 mA g­1), it can retain 94% of the initial capacity after 3000 cycles. It was found that the outstanding rate performance and long­term cycling life of Na­VBNT@C is primarily due to the Na+ pseudocapacitance. Our study reveals that the design of Na+ pseudocapacitance is beneficial for harvesting superior performance of NaV6O15 cathode material in sodium­ion batteries. KEYWORDS: vanadium bronze; sodium­ion batteries; nanotube; pseudocapacitance; hydrothermal

1. Introduction 1 ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33 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

Among the reported transition metal oxides (TMOs), layered V2O5 cathode material can deliver a theoretical capacity of 441 mAh g­1 within the potential range of 1.5­4.0 V (vs. Li+/Li) in lithium­ion batteries (LIBs).1­4 However, the unstable structure, weak conductivity and low ionic diffusion coefficient of V2O5 cause a sharp decrease in capacity upon cycling. In our previous study, the Mg­doped V2O5 nanoparticles embedded in carbon matrix presented an excellent electrochemical performance for LIBs5 because the doped Mg2+ ions pillared the layered structure to form a three­dimensional (3D) structure.6­8 Nonetheless, the layered structure pre­intercalated by smaller cations such as Mg2+, Fe2+, Mn2+, Al3+ and Sn4+ does not favor de­/intercalation of the large Na+ ions. Notably, chemical pre­intercalation of Na+ would significantly improve the sodium­storage performance of the layered vanadium oxides. When 1/3 moles of Na+ ions are pre­intercalated per V2O5 unit, a sodium vanadium bronze NaV6O15 (i.e. β­Na0.33V2O5) with two­dimensional (2D) V4O12 layers connected by VO5 and edge­sharing oxygen atoms can be achieved. The pre­intercalated Na+ ions between the V4O12 layers act as “pillars” to stabilize the structure upon insertion/extraction of Li+/Na+ ions. Although NaV6O15 has revealed high performance when used as the cathode in LIBs,9,10 its cyclability and rate capability are still unsatisfactory as the cathode in SIBs.11­18 Thus, the carbon was coated on the surface of the NaV6O15 nanomaterials in order to further improve its electrical conductivity. Due to the synergistic effect of NaV6O15 and carbon coating, the NaV6O15@C composite materials with novel nanostructure would exhibit higher specific capacity and intensive cycle stability during cycling.The surface coating of electrode particles with a highly conductive substance (such as carbon) can increases the conductivity and facilitate the electron transportation in the electrode. Moreover, carbon acts as a bridge to connect the particles, as well as the protective shell layer to alleviate the volume change of electrodes, prevent the vanadium dissolution, and alleviate the aggregation of the particles during the charge and discharge processes, ensuring better cycleability during long­term charge/discharge.5,6,19 It is well known that nanomaterials possess shorter diffusion pathway and more intercalation sites than bulk materials. In particular, the electrode materials present a pseudocapacitive storage behavior during sodiation/desodiation processes when they have a structure consisted of nanoparticles integrated with carbon matrix.20 The resultant 2 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

pseudocapacitance is relative to a faradaic redox reaction that happens on/near the surface of the electrode materials; In consequence, the charge storage mechanisms are different from those in traditional batteries because they are not confined by the ion diffusion into the bulk material.21 In addition, pseudocapacitance can be divided into two types: intrinsic pseudocapacitance (such as RuO2, MnO2, Nb2O5, MoO3 and polyaniline) and extrinsic one (such as LiCoO2, TiO2 and V2O5). Compared with the diffusion­controlled process, intercalation pseudocapacitance shows obvious advantages for batteries because it can occur within the bulk of electrode materials, providing batteries with faster charging/discharging property by reducing the solid state diffusion distance of sodium ions within the electrodes. Hence, it can lead to high­power and faster charging/discharging rates together with long­term cycling stability, and accordingly delay the degradation of electrode material.21­24 For instance, Dunn and co­workers25 research showed great prospect towards high­rate electrodes for LIBs propelled by an intercalation pseudocapacitive mechanism. Chen26, inspired by this, proved that intercalation pseudocapacitance predominates the charge storage progress in G­TiO2 for SIB anodes, which shows excellent rate performance and long­term stability. In brief, the pseudocapacitance­dominated contribution guarantees much faster and more stable Na storage performance.23 Unfortunately, all previously reported intercalation pseudocapacitances are focused on the capacitors,27­30 and less attention has been drawn to the pseudocapacitance contribution in batteries. In our work, carbon­coated NaV6O15 nanotubes with open ends are used as the cathode material for SIBs. The nanotubes with high crystallinity are intertwined with quasi spheres, and the carbon layer is homogeneously coated on the surface (Scheme 1). Owing to the opened terminates, the electrolyte could penetrate into the Na­doped nanotubes, rendering Na+ ions able to intercalate into the inner surface of nanotubes rapidly; while the coated carbon layer facilitates electrons passing through the outer surface of the nanotubes. As a result, the intercalation­based electrochemical reaction can occur rapidly. Moreover, excessive Na+ ions in nanotube and electrons in carbon layer simultaneously might lead to double layer capacitor at the interface, devoting additional capacity for the composite materials. More importantly, we demonstrate ultrafast, highly reversible, and cycle­stable sodium storage, boosted by a pseudocapacitance contribution from the carbon­coated NaV6O15 nanotubes. 3 ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33 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

Indeed, developing a highly pseudocapacitive contribution for high­performance Na­ion batteries may be particularly significant for development of the next­generation electrochemical energy storage devices.

2.

Experimental

2.1 Synthesis of the carbon-coated Na-VBNTs Vanadium oxide (VOx) nanotubes were prepared according to the reported method.31­35,41 Firstly, 1.82 g (0.01 mol)of V2O5 powder (AR., Xilong Chemical Reagent Company, China) was added slowly to 50 ml of H2O2 solution (30%, AR., Sinopharm Chemical Reagent Co. Ltd ), and stirred for 2h to form a sol. Subsequently, 7.42 g (0.04 mol) of dodecylamine (C12H25NH2, A.R., Sigma­Aldrich) was added to the sol, and agitated for 5h. After that, the sol was transferred to a Teflon­lined stainless steel autoclave, and heated at 200 o

C for five days. The precipitates were obtained after centrifugation, and then washed with

distilled water and ethanol for three times. About 1.26 g of amine­intercalated VOx powders was obtained after the precipitates were dried at 80 oC overnight under vacuum. For comparison, a part of VOx powders was separated and calcined at 450 oC at a ramping rate of 2 oC min­1 in air for 3h to obtain the V2O5 nanotubes, designated as VONT. The calcination temperature was determined from the thermo­gravimetric analysis of the precursors (Figure S1a, supporting information). Secondly, 2.13 g (0.015 mol) of Na2SO4 and 0.91 g (0.005 mol) of the obtained VOx powder were dispersed in 50 ml of distilled water/ethanol (1:4 in volume) solution, and stirred for 24h. The mixture was then filtered, and the residue was washed with water and ethanol. When the washed residue was dried at 80 oC under vacuum for 24h, the Na­doped precursors were obtained, and then heated from room temperature to 450 oC at a heating rate of 2 oC min­1, and held in air for 3h. After being cooled in muffle furnace, the sodium vanadium bronze nanotubes were acquired, designated as the Na­VBNT. Finally, 0.50 g of Na­VBNT powder was dispersed in a solution of 0.15 g sucrose in 10 ml deionized water and stirred for 1h, then heated at 80 oC until the water evaporated almost. After that, the dried powders were annealed at 400 oC for 0.5h in a tube furnace under argon atmosphere to get the carbon­coated sodium vanadium bronze nanotubes, designated as the 4 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

Na­VBNT@C. The mass percent of carbon was determined to be about 7.2% in the composite material through the TG analysis (Figure S1b) when the product was fired in air at a ramping rate of 10 oC min­1 from 30 to 550 oC.

2.2 Materials characterization Thermo­gravimetric analysis (TGA) was measured using the TA­SDTQ 600 thermal analysis system (Texas Instruments, Inc., New Castle, DE) at a ramping rate of 10 oC min­1. Scanning electron microscopy (SEM, JEOL JSM­6510A) and transmission electron microscopy (TEM, JEOL JEM­2010) were applied to observe the morphology of the sample. X­ray diffraction (XRD) patterns were performed by a powder diffractometer (Bruker D8 Advanced) with a Ni filter at a sweep rate of 2° min­1 in the 2θ range of 5°­65° (Cu Kα, (λ = 0.154 nm) source). The nitrogen adsorption­desorption isotherm was characterized using TriStar II 3020 instrument (Micromeritics Instrument Corp.), and the specific surface area was characterized by the Brunauer­Emmett­Teller (BET) method. Na and V contents in the samples were conducted by inductively coupled plasma atomic emission spectrometry (ICP­AES, Profile spectrometer, Leeman Laboratories). Raman spectrum was collected using a RenishawInVia Raman Microscope. The valence state of the V element was analyzed by X­ray photoelectron spectroscopy (XPS, Thermo Escalab 250XI).

2.3 Electrochemical characterization Na­VBNT@C electrodes were prepared through mixing the as­prepared Na­VBNT@C powder, acetylene black (Super P) and PVDF (Knyar 2700) binder with the mass ratio of 80:10:10 at first. N­methyl­2­pyrrolidone (NMP) was then added to the mixture and agitated to form slurry. The slurry was cast on aluminum foil and dried at 120 oC under vacuum overnight. Finally, the dried plate was punched into the discs with a diameter of 12 mm to creat the Na­VBNT@C electrodes. The VONT and Na­VBNT electrodes were fabricated similar to the Na­VBNT@C electrodes. About 3.0 mg of the electroactive materials were loaded on each disc electrode. The tap densities were measured to 1.140±0.022, 1.161±0.018 and 1.025±0.020 g cm­3 for the VONT, Na­VBNT and Na­VBNT@C samples, respectively. 5 ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33 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

Cyclic voltammetry was conducted using the CHI660E electrochemical work station at various sweep rates over the voltage window of 4.0­1.5 V (vs. Na+/Na) using a two­electrode system, in which the cathode acts as the working electrode and sodium plate as the counter and the reference electrodes. The cathode and Na electrode were separated by glass microfiber filter (GF/A, Whatman), immersed in the electrolyte 1 M NaClO4 in ethyl carbonate (EC)/dimethyl carbonate (DMC) (1:1 in weight), and sealed in CR2025­type button cell. Electrochemical impedance spectroscopy (EIS) measurement was carried out by the same electrochemical workstation. Galvanostatic charge­discharge measurements were performed on the CT­3008W Neware battery tester using the CR2025 coin­type cells.

3.

Results and discussion

3.1 Morphology of the sample SEM images of the VONT, Na­VBNT and Na­VBNT@C samples are shown in Figure S2. It is found that the Na­VBNT sample reveals the morphology of clew­like quasi spheres with a diameter range of 2­5 μm (Figure S2c), which is similar to that of VONT sample (Figure S2a,b). These spheres are constructed with one­dimensional (1D) nanomaterials (Figure S2d) that are nanotubes in fact, as shown in Figure 1f. Na­VBNT@C particles (Figure S2f) are larger than Na­VBNT particles (Figure S2e), and some ultrafine particles are observed on their surfaces (Figure S2f). In addition, the porous structure of Na­VBNT@C (Figure S2f) is different from that of Na­VBNT (Figure S2d). The reason is that the carbon component generated from sucrose coated the nanotubes and filled the pores and made the clews integrate into larger particles as well. Evidently, the Na­VBNT nanotubes display a length range of 0.3~2 μm (Figure 1a) and an outer diameter range of 15~30 nm (Figure 1b). Open ends of the nanotubes marked with yellow arrows in Figure 1b, are beneficial for electrolyte penetration and ions pass through the nanotubes. From the high­resolution TEM image (HRTEM, Figure 1c), the inner diameter and wall thickness of the nanotube are observed to be about 5 nm and 8 nm, respectively. The distance between two adjacent nanoscrolls is about 0.45 nm, which is lower than that of the reported Na­VOx nanotubes without calcination (0.92 nm).41 This might be due to decomposition of the intercalated amine template at high temperature. One can see that the 6 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

nanoscrolls exhibit some defects compared to those of VOx nanotubes without calcination.36­40 In addition, Na­VBNT displays a polycrystalline structure, which can be verified by the selected area electron diffraction (SAED) pattern (Figure 1d). Moreover, Na­VBNT@C morphology appears to be consisted of adherent clews­like spheres (Figure 1e), with open­ended nanotubes (Figure 1e,f marked with yellow arrows). It is observable that Na­VBNT@C still retains the nanotube morphology (Figure 1g) as well as the polycrystalline structure (Figure 1h) after repeated annealing at 400 oC under the inert atmosphere. A coating layer around 1 nm in thickness is noticeable on surface of the nanotube (Figure 1g), which is determined to be amorphous (Figure 1g, marked with white areas). This coating layer is confirmed to be carbon component that generated from the pyrolysis of sucrose by comparing the Energy­dispersive X­ray spectroscopy (EDS) spectra of Na­VBNT (Figure S3a) and Na­VBNT@C (Figure S3b), which come from the selected area in Figure 1a and 1e, respectively. The co­existence of Na, V and O elements in the sample indicates that the product may be a kind of vanadium bonze doped with sodium (Figure S3a). According to the ICP­AES analysis, the precise molar ratio of Na/V is determined to be 0.167 and 0.159 in Na­VBNT and Na­VBNT@C, respectively, both close to that in the chemical formula of NaV6O15(0.163).

3.2 Porous structure of the samples To understand the porous structures of the resultant samples, N2 adsorption/desorption tests were conducted. As shown in Figure S5, all three isotherms accord with the type­IV behavior with a H3­hysteresis loop, which illustrated a mesoporous structure. The BET specific surface areas are 11.9, 12.5 and 15.7 m2 g­1 for VONT, Na­VBNT and Na­VBNT@C, respectively. Different from other samples, Na­VBNT@C presents a little steep profile over the p/p0 values ranged from 0 to 0.1, indicating a microporous structure (Figure S5e). The pore size distribution (PSD) plots of three samples are calculated based on the Barrett­Joyner­Halenda (BJH) procedure, and are shown in their respective N2 adsorption/desorption isothermal figures. It is found that three peaks centered at 3.3, 19.0 and 40.0 nm (Figure S5d) for Na­VBNT are similar to those of VONT (3.2, 18.8 and 40.4 nm, Figure S5b), implying that the porous structure is retained after repeated firings at high temperatures. In the case of Na­VBNT@C, a number of micropores (centered at 1.3 nm) 7 ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33 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

co­exist, and the mesopores become smaller (17.2 and 34.3 nm) (Figure S5f). It is worth noting that the inner diameter of nanotube hardly changes for this carbon­coated sample, suggesting the carbon layer is completely deposited onto the surface of nanotubes. Among the pores, the smaller pores originate from the penetrating nanotube (Figure 1c and g) while the larger pores stem from the cavities surrounded by the intertwined nanotubes (Figure S2d and f). The latter could serve as a reservoir to store the electrolyte within bulk particles, and the former could provide the channels for ion transportation.19 Such a unique architecture, schematically depicted in Figure 3i, is beneficial to accelerating the electrochemical reaction due to fast transportation of both Na+ ions and electrons. Besides, side reactions like the decomposition of the electrolyte components on the surface of electroactive particles would be suppressed to some extent since the BET surface area of the micrometer quasi­spheres is lower than that of nanomaterials. Moreover, the micron­size spheres are much more easily casted on the current collector compared with the nanotubes. Because nanoparticles with huge specific surface energies are thermodynamic unstable, they extensively aggregate into irregular particles to low their surface energy. On the other hand, the nanotubes were difficultly dispersed homogeneously in the casting slurry even if they could retain the nanoscale morphology. 3.3 Crystal structure of the samples Figure 2a exhibits the XRD pattern of the Na­VBNT sample. All characteristic peaks could be indexed to the monoclinic phase sodium vanadium bronze (NaV6O15). According to the retrieved refinement by GSAS II software,42 the fitted XRD pattern (Rp=6.58%, χ2=3.2) agrees well with that of the standard card (JCPDS no. 77­0146).13 The refined lattice parameters of Na­VBNT are a=10.077 Å, b=3.601 Å, c=15.458 Å and β=109.38o. This suggests the as­prepared sample has a relatively pure phase with a space group of A2/m (12). Indeed, no impurities were detected in the sample and its high degree of crystallinity well­matched the SAED pattern (Figure 1d). Additionally, its crystal structure model (Figure 2b) indicates that the sample exhibits a 3D­tunnel structure along the [010] direction, which is constructed of VO5 pyramids and VO6 octahedra with three kinds of vanadium sites, labeled as V1, V2 and V3. This results in zigzag double chains composed of edge­sharing (V1)O6 and corner­sharing (V2)O6 distorted octahedra as same as edge­sharing (V3)O5 squared­base 8 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

pyramids.44,45 Na+ ions are aligned in the 3D tunnels indeed. Figure 2c displays the XRD patterns of the the three samples. It is found that the diffraction peaks of VONT match those of the standard card (JCPDS no. 41­1426). Na­VBNT@C exhibits the similar characteristic peaks as those of Na­VBNT except for a less intensity. Owing to amorphous nature and little amount, the diffraction peaks that belong to carbon are not detected. Raman spectra of the three samples are displayed in Figure 2d. One can see that twelve vibration peaks locate at 126, 151, 221, 271, 323, 388, 440, 516, 553, 693, 993 and 1011 cm­1 for Na­VBNT. According to those peaks of β­Na0.33V2O543,46,47 and β­Ca0.33V2O5,48 the higher­frequency modes can be ascribed to the stretching vibrations of the different V­O bonds in (V1)O6, (V2)O6 and (V3)O5 polyhedra. The band at 1011 cm­1 represents V2­O6 stretching vibration while the one at 993 cm­1 can be ascribed to V1­O4 stretching vibration. The Raman bands at 693 and 440 cm­1 correspond to the asymmetric V1­O2­V1 stretching vibration and V1­O5­V3 bending vibration, respectively. The peak at 553 cm­1 is attributed to V1­O3 stretching vibration while the one at 516 cm­1 stems from V3­O7 stretching vibration. The Raman bands at the lower frequencies below 400 cm­1 might be relative to those ladder modes of the V atoms and Na­O bonds.49,50 The Raman spectrum is in agreement with the molecular structure of NaV6O15 presented in Figure 2b. After coating with carbon, two additional vibration peaks located at 1353 and 1590 cm­1 appear in the Raman spectrum of the Na­VBNT@C sample. of which, the former is related to the presence of disordered carbon structures (D­band) whereas the latter is bound up with the vibration of sp2­hybridized carbon in the graphite crystallites (G­band).51 The intensity ratio of D­band to G­band (ID/IG) is about 1.1, indicative of amorphous carbon. VONT displays ten vibration peaks at 98, 142, 193, 281, 298, 405, 478, 524, 696 and 993 cm­1, similar to the reported Raman spectra.5,6 Figure 2e and 2f show the high resolution XPS spectra of V2p signal of Na­VBNT and Na­VBNT@C, respectively. As it appears in Figure 2e, two peaks centered at 517.6 and 525.0 eV belong to the binding energy of V2p3/2 and V2p1/2, respectively. After the XPS spectra are fitted using XPS peak 4.1 software, these two peaks are split correspondingly into two pairs of peaks related to V4+ and V5+ valences. Clearly, as for Na­VBNT, the peaks at 517.4 eV and 524.8 eV are ascribed to V5+2p3/2 and V5+2p1/2, respectively,52 while those at 516.4 eV and 523.5 eV with fewer intensity can be ascribed to V4+2p3/2 and V4+2p1/2. 9 ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33 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

Na­VBNT@C displays similar peaks to those of Na­VBNT: 517.5 eV (V5+2p3/2)/516.9 (V4+2p3/2) and 525.1 eV (V5+2p1/2)/523.8 eV (V4+2p1/2). The fitted areas from the V5+ and V4+ components can be utilized to estimate the content of V4+ in the compound. The atomic ratio of V4+/V5+ is calculated to be 0.202 for Na­VBNT and 0.231 for Na­VBNT@C, matching the ratio (0.200) of the given chemical formula NaV4+1V5+5O15. The amount of V4+ component in Na­VBNT@C is more than that of the Na­VBNT because the former performed a pyrolysis of sucrose in Ar atmosphere. The peaks related to nitrogen element are not observed in the XPS survey spectrum of Na­VBNT (Figure S4), indicating the organic amine template has been almost removed from the crystals. 3.4 Battery performance of the samples Figure 3 displays the galvanostatic charge­discharge curves of VONT (a), Na­VBNT (b) and Na­VBNT@C (c) at 0.1C rate within the potential range of 1.5­4.0 V for different cycles and the corresponding cycling performances (Figure 3e). It is found that the initial discharge profiles for Na­VBNT and Na­VBNT@C are different from that of VONT, which displays a lower voltage plateau than those subsequent ones. This indicates that the Na­doped electrodes possess higher polarization than that of the VONT electrode during the first discharge process. The initial discharge capacities are 152, 132 and 168 mAh g­1 for the VONT, Na­VBNT and Na­VBNT@C electrodes, respectively. With continuous cycling, the discharge capacity of the VONT electrode significantly degrades(Figure 3e); whereas those of the Na­VBNT and Na­VBNT@C electrodes increase and reach the maximum values of 188 (after ten cycles, Figure 3e) and 209 (after five cycles, Figure 3e) mAh g­1, respectively. It is noticeable that the VONT electrode has a poor cycling performance compared to the Na­doped samples. It is observed that the Na­VBNT@C electrode delivers a higher capacity than the Na­VBNT electrode, even the Na­VBNT@C electrode offered the maximum discharge capacity of 209, 198, 185, 160, 141, 130, 125, and 105 mAh g­1 at the rates of 0.1, 0.2, 0.5, 1, 2, 4, 5, and 10C (Figure 3d), respectively, although the former possesses a less amount of NaV6O15 and a larger amount of V4+ component in the formula than the latter. To explain this phenomenon, a “job sharing” mechanism might be applied.53 As is depicted in Figure 3i, the NaV6O15 nanotubes are coated with nanocarbon layers, leading to a “Janus” interface in the composite material. Na+ ions intercalate into the inner surface of nanotubes while the 10 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

electrons pass through the nanocarbon layer (outer surface of the nanotubes). Owing to the short diffusion distance of Na+ into the nanotube (l Na­VBNT (0.12 S cm­1) > VONT (0.006 S cm­1), which are determined by the four­probe method. On the other hand, one can see, in Figure S6a, the charge transfer resistance of the Na­VBNT@C is lower than those of the other samples. It is due that carbon coating can improve the electrical conductivity of materials.It is noted that the CV curves appear as a rectangle over the potential ranges of 1.5­2.0 V and 3.5­4.0 V, similar to those of with pseudocapacitive behaviors.63 This result confirms the linear charge and discharge plots during the latter stages. It has been widely that the electrode materials based on intercalation reaction would exhibit capacitive behavior in a large part if they are produced in form of nanoscale particles.64 Owing to higher surface area and shorter diffusion pathway, nanomaterials capable of cationic intercalating would undergo a serious redox reaction at the surface or in the near­surface regions. This leads to a faradic capacitance, so­called “intercalation pseudocapacitance”,65 which contributes to the whole delivered capacity of the electrode delivered.66­69 The contribution of this pseudocapacitance can be characterized by analyzing the CV data on the basis of the relationship between measured current (i) and sweep rate (v) (Eq. (1)).

i  av b (1) where a and b are adjustable constants. When the value of b approaches 0.5, the current is contributed by ion diffusion; when the value of b is close to 1.0, the current is dominated by capacitive behavior.20 Figure 4b displays the CV profiles at various scan rates for the Na­VBNT@C. From these curves, the plot of log(i) versus log(v) can be drawn (Figure 4d­f). The slope of the fitting line equals to the value of b. The plots of log(i) versus log(v) for VONT and Na­VBNT are shown in Figure 4d and 4e, respectively. It is found that the b values for all samples are larger than 0.5 in both anodic and cathodic directions, implying that diffusion and capacitive controls occur during Na+ de­/intercalation process. As a result, the current can be divided into two parts: capacitive current and ion­diffusion current (Eq. (2)). 13 ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33 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

i (V )  k1v  k2 v1/2

(2)

In the equation, k1v and k2v1/2 come from the capacitive contribution and ion­diffusion contribution, respectively. The constants of k1 and k2 can be obtained from the i(V)/v1/2 versus v1/2 plot at a given potential (Figure S8). Thus, we can quantify the fraction of the two types of currents at specific potentials. Figure 4c presents the contribution of the intercalation pseudocapacitance estimated from the CV curve at 0.5 mV s­1 based on Eq (2) for Na­VBNT@C. It is found that the capacitive contribution (Figure 6c, violet area) forms a large constitution of the stored charges, agreeing well with those calculated b values (Figure 4i). Regarding the analysis of ex­situ XRD patterns (Figure 5a), the crystallographic phase in Na­VBNT@C hardly changes in the Na­VBNT@C during sodiation/desodiation process. Accordingly, the capacitive behavior for the Na­VBNT could be regarded as intercalation­type pseudocapacitance. Figure 4g, 4h and 4i display the total capacities for the VONT, Na­VBNT and Na­VBNT@ electrodes at various sweep rates, respectively. It is found that VONT possesses the highest percentage of diffusion contribution to total capacity among the three samples whereas the Na­VBNT@C reveals the highest percentage of capacitive contribution. In addition, each sample presents an increased percentage in capacitive contribution with increasing the sweep rate. The results suggest that the pseudocapacitance­dominated storage accounts for a large proportion of the total capacity, particularly for more conductive materials at higher sweep rates, which is in agreement with the reports.69­71 Therefore, the pseudocapacitive behavior plays an important role in storing charge for Na­VBNT because the rate of faradic redox reaction at the surface of nanotubes is considerably faster than that of ionic diffusion into the electrode. 3.6 Structural analysis of Na-VBNT@C at different charge states Figure 5a presents the ex­situ XRD patterns of the Na­VBNT@C electrode at different charge­discharge states. The detailed lattice parameters form the XRD patterns are summarized in Table 1. Notably, during the discharge (Na­intercalation) process, the lattice parameter along the c-axis continuously rises up while those along the a- and b-axes hardly change, resulting in expansion of unit cells. Indeed, during the charge (Na­deintercalation) process, the lattice parameter along the c-axis reduces, leading to contraction of unit cells. At 14 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

the fully charged state, the core nanotubes presented similar lattice parameters to those of the fresh one (Table 1), indicating that Na+ ions can be reversibly inserted into and extracted from the obtained nanotubes. It is also observable that the diffraction peaks related to the (11­1), (104), (21­3) and (31­7) planes moved to lower diffraction angles during discharge course, but shifted towards higher diffraction angles during charge process (Figure 5b), which is consistent with the structural change tendency of lattice parameters. This phenomenon of lattice expansion and shrinkage could be called a “lattice breathing”,7,72,73 which allows the core nanotubes to keep their crystal structure during de­/intercalation of Na+ ions, confirming the Na­intercalation pseudocapacitive behaviors for the samples. Interestingly, at the fully charged state after 100 cycles, the core nanotubes owned a structure close to that at the 2.3 V discharged state, implying that a small amount of Na+ ions might be trapped in the crystal. These entrapped Na+ ions would reduce the delivered capacity of the core nanotubes because they occupy the intercalating sites in the crystal, but they could serve as “pillars” to stable the structure. The pillar effect from the intercalated Na+ ions can be conspicuously observed from the HRTEM images of the core nanotubes sample before cycling (Figure S9a) and after 100 cycles (Figure S9b). We can see that the core nanotubes still present distinct crystalline structure, although some disorders appear inside the crystals after cycling. As a consequence, Na­VBNT@C possesses a superior cycling performance compared to the other vanadium­based nanomaterials (Table S1, supporting information.). Figure 6 shows SEM images of the surface of Na­VBNT@C electrode after 1000 cycles when run at 5C rate. One can see that although a few of cracks appear at the surface of the electrode after cycling (Figure 6a), some sphere­like particles still maintain the morphology of nanotube­intertwined clews (Figure 6b), similar to the original one (Figure S2e). This result implies that the Na­VBNT@C quasi­spheres could accommodate the volume change of NaV6O15 crystallites during the cycling process. It is reasonably concluded that Na­intercalation pseudocapacitance could endow the Na­VBNT@C electrode with outstanding rate capability and long­term cyclability because the faradaic redox reactions on/near the surface of nanotubes cannot destroy the crystal structure and hierarchical architecture. 15 ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33 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

4. Conclusions Clew­like sodium vanadium bronze/carbon (Na­VBNT@C) composites composed of carbon­coated NaV6O15 nanotubes have been successfully prepared by a hydrothermal method using an organic amine template. The Na­VBNT@C sample showed quasi­spherical morphology with a hierarchical architecture, rendering the electroactive Na­VBNT to own outstanding rate capability and relatively long life span when used as the cathode for sodium­ion batteries. Na­VBNT@C could deliver an initial capacity of 168 mAh g­1, and reach the maximum capacity of 209 mAh g­1 at 0.1C rate. By the analysis of CV measurements, an intercalation pseudocapacitance could be employed to explain its charge­storage mechanism. Owing to this mechanism, it revealed an excellent rate capability, and delivered a high capacity up to 105 mAh g­1 at 10C rate (2500 mA g­1). Benefitting from the hierarchical architecture, the sample could retain 94% of the initial capacity after 3000 cycles when operated at 5C rate. The results suggest that one­dimensional carbon­coated nanotubes with open ends can facilitate not only Na+ ions diffusion but also electrons transportation. Moreover, the intertwined clews could alleviate the volume change of crystallites during the cycling process. This work demonstrates that the hierarchical carbon­coated NaV6O15 nanotube­based quasi­spheres would be developed to a promising candidate as cathode material for sodium­ion batteries. The sluggish Na­diffusion kinetics and large­volume change of crystallite, which considerably limit the wide application of insertion­type cathode materials in SIBs, would be improved significantly by introducing intercalation pseudocapacitive behaviour to cathode materials through the rational designing of particles morphology.

ASSOCIATED CONTENT Supporting information Thermo­gravimetric (TG) and differential TG (DTG) curves of the Na­doped precursors and the Na­VBNT@C in air atmosphere at a ramping rate of 10 oC min­1, SEM images of the VONT sample. SEM images of the Na­VBNT and Na­VBNT@C samples, EDX spectra of the Na­VBNT and Na­VBNT@C samples, XPS survey spectrum of the Na­VBNT sample, Nitrogen adsorption­desorption isotherms of the VONT, Na­VBNT and Na­VBNT@C 16 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

samples, and the corresponding pore size distribution plots, Nyquist curves and the corresponding Zre versus ω­1/2 plots of the VONT, Na­VBNT and Na­VBNT@C samples, Cyclic voltammograms of for the first three cycles at a sweep rate of 0.1 mV s­1 over the potential range of 1.5­4.0 V (vs. Na+/Na), Galvanostatic charge­discharge curves at 0.1C rate for various cycles, cyclic performance at the 5C rate, and Rate capability of the coating­carbon electrode within the 1.5­4.0 V range. The coating­carbon electrodes were prepared by mixing the carbon that generated from the pyrolysis of sucrose under Ar atmosphere with acetylene black (Super P) and PVDF (Knyar 2700) binder (mass ratio of 80:10:10) in N­methyl­2­pyrrolidone (NMP) to form slurry, cast on aluminum foil and dried at 120 oC under vacuum overnight, and punched into the discs with a diameter of 12 mm, The i(V)/v1/2 versus v1/2 plots of the Na­VBNT@C sample at various potentials, HRTEM images of the Na­VBNT@C sample before cycling and after 100 cycles, Comparison of the electrochemical properties of the Na­VBNT@C and other vanadium­based nanomaterials for Na­ion batteries over the voltage range of 1.5­4.0 V.

ACKNOWLEDGEMENTS The authors greatly appreciate the financial support from National Natural Science Foundation of China (grant numbers 21376069, 21576075, 51672189), Hunan Provincial Natural Science Foundation of China (grant numbers 2018JJ2386, 2018JJ2393), National Key Research and Development Program of China (2018YFB0105900), and Xi'an Science and Technology Project of China (201805037YD15CG21(20)).

REFERENCES (1) Ni, J.; Fu, S.; Yuan, Y.; Ma, L.; Jiang, Y.; Li, L.; Lu, J. Boosting Sodium Storage in TiO2

Nanotube Arrays through Surface Phosphorylation. Adv. Mater. 2018, 30, 1704337. (2) Yao, J.; Li, Y.; Massé, R. C.; Uchaker, E.; Cao, G. Revitalized Interest in Vanadium Pentoxide as Cathode Material for Lithium­Ion Batteries and Beyond, Energy Storage Mater. 2018, 11, 205­259.

17 ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33 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

(3) McNulty, D.; Buckley, D. N.; O'Dwyer, C. Synthesis and Electrochemical Properties of Vanadium Oxide Materials and Structures as Li­Ion Battery Positive Electrodes. J. Power Sources

2014, 267, 831­873. (4) Yuan, Y.; Hong, L. Micro­ and Nano­Structured Vanadium Pentoxide (V2O5) for Electrodes of Lithium­Ion Batteries. Adv. Energy Mater. 2017, 7, 1602545. (5) Xiao, F.; Song, X.; Li, Z.; Zhang, H.; Zhang, L.; Lei, G.; Xiao, Q.; Hu, Z.; Ding, Y. Embedding of Mg­doped V2O5 Nanoparticles in a Carbon Matrix to Improve Their Electrochemical Properties for High­Energy Rechargeable Lithium Batteries. J. Mater. Chem. A 2017, 5, 17432­17441. (6) Peng, C.; Xiao, F.; Yang, J.; Li, Z.; Lei, G.; Xiao, Q.; Ding, Y.; Hu, Z. Carbon­Encapsulated Mn­Doped V2O5 Nanorods with Long Span Life for High­Power Rechargeable Lithium Batteries.

Electrochim. Acta 2016, 192, 216­226. (7) Clites, M.; Pomerantseva, E. Bilayered Vanadium Oxides by Chemical Pre­Intercalation of Alkali and Alkali­Earth Ions as Battery Electrodes. Energy Storage Mater. 2018, 11, 30­37. (8) Song, H.; Liu, C.; Zhang, C.; Cao, G. Self­doped V4+­V2O5 Nanoflake for 2 Li­Ion Intercalation with Enhanced Rate and Cycling Performance. Nano Energy 2016, 22, 1­10. (9) Tang, H.; Peng, Z.; Wu, L.; Xiong, F.; Pei, C.; An, Q. Vanadium­Based Cathode Materials for Rechargeable Multivalent Batteries: Challenges and Opportunities. Electrochem. Energy Rev. 2018, 1, 169­199. (10) Baddour­Hadjean, R.; Bach, S.; Emery, N.; Pereira­Ramos, J. P. The Peculiar Structural Behaviour of β­Na0.33V2O5 upon Electrochemical Lithium Insertion. J. Mater. Chem. 2011, 21, 11296­11305. (11) He, H.; Zeng, X.; Wang, H.; Chen, N.; Sun, D.; Tang, Y.; Huang, X.; Pan, Y. NaV6O15 Nanoflakes with Good Cycling Stability as a Cathode for Sodium Ion Battery. J. Electrochem.

Soc. 2015, 162, A39­A43. (12) Safrany Renard, M.; Emery, N.; Baddour­Hadjean, R.; Pereira­Ramos, J.­P. γ'­V2O5: A New High Voltage Cathode Material for Sodium­Ion Battery. Electrochim. Acta 2017, 252, 4­11. (13) Jiang, D.; Wang, H.; Li, G.; Lan, X.; Abib, M. H.; Zhang, Z.; Jiang, Y. Self­Combustion Synthesis and Ion Diffusion Performance of NaV6O15 Nanoplates as Cathode Materials for Sodium­Ion Batteries. J. Electrochem. Soc. 2015, 162, A697­A703.

18 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

Page 20 of 33

(14) Hu, F.; Jiang, W.; Dong, Y.; Lai, X.; Xiao, L.; Wu, X. Synthesis and Electrochemical Performance of NaV6O15 Microflowers for Lithium and Sodium Ion Batteries. RSC Adv. 2017, 7, 29481­29488. (15) Liu, H.; Zhou, H.; Chen, L.; Tang, Z.; Yang, W. Electrochemical Insertion/Deinsertion of Sodium on NaV6O15 Nanorods as Cathode Material of Rechargeable Sodium­Based Batteries. J.

Power Sources 2011, 196, 814­819. (16) Cai, Y.; Zhou, J.; Fang, G.; Cai, G.; Pan, A.; Liang, S. Na0.282V2O5: A High­Performance Cathode Material for Rechargeable Lithium Batteries and Sodium Batteries. J. Power Sources. 2016, 328, 241­249. (17) He, H.; Jin, G.; Wang, H.; Huang, X.; Chen, Z.; Sun, D.; Tang, Y. Annealed NaV3O8 Nanowires with Good Cycling Stability as a Novel Cathode for Na­Ion Batteries. J. Mater. Chem.

A. 2014, 2, 3563­3570. (18) Kang, H.; Liu, Y.; Shang, M.; Lu, T.; Wang, Y.; Jiao, L. NaV3O8 Nanosheet@Polypyrrole Core­Shell Composites with Good Electrochemical Performance as Cathodes for Na­Ion Batteries. Nanoscale 2015, 7, 9261­9267. (19) Zhang, X.­F.; Wang, K.­X.; Wei, X.; Chen, J.­S. Carbon­Coated V2O5 Nanocrystals as High Performance Cathode Material for Lithium Ion Batteries. Chem. Mater. 2011, 23, 5290­5292. (20) Chao, D.; Zhu, C.; Yang, P.; Xia, X.; Liu, J.; Wang, J.; Fan, X.; Savilov, S. V.; Lin, J.; Fan, H. J. Array of Nanosheets Render Ultrafast and High­Capacity Na­Ion Storage by Tunable Pseudocapacitance. Nat. Commun. 2016, 7, 12122. (21) Rauda, I. E.; Augustyn, V.; Saldarriaga­Lopez, L. C.; Chen, X.; Schelhas, L. T.; Rubloff, G. W.; Dunn, B.; Tolbert, S. H., Nanostructured Pseudocapacitors Based on Atomic Layer Deposition of V2O5 onto Conductive Nanocrystal­based Mesoporous ITO Scaffolds. Adv. Funct. Mater. 2014, 24, 6717­6728. (22) Liu, J.; Wang, J.; Xu, C.; Jiang, H.; Li, C.; Zhang, L.; Lin, J.; Shen, Z. X. Advanced Energy Storage Devices: Basic Principles, Analytical Methods, and Rational Materials Design, Adv. Sci. 2017, 5, 1700322. (23) K.

Ma,

H.

Jiang,

Y.

Hu,

C.

Li,

2D

Nanospace

Confined

Synthesis

of

Pseudocapacitance­Dominated MoS2­in­Ti3C2 Superstructure for Ultrafast and Stable Li/Na­Ion Batteries, Adv. Funct. Mater. 2018, 28, 1804306. 19 ACS Paragon Plus Environment

Page 21 of 33 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

(24) Xia, C., Z. Lin, Y. Zhou, C. Zhao, H. Liang, P. Rozier, Z. Wang, and H. N. Alshareef, Large Intercalation Pseudocapacitance in 2D VO2 (B): Breaking through the Kinetic Barrier, Adv Mater. 2018, 1803594. (25) Augustyn, V.; Come, J.; Lowe, M. A.; Kim, J. W.; Taberna, P. L.; Tolbert, S. H.; Abruna, H. D.; Simon, P.; Dunn, B., High­Rate Electrochemical Energy Storage through Li+ Intercalation Pseudocapacitance. Nat Mater 2013, 12 , 518­522. (26) Chen, C.; Wen, Y.; Hu, X.; Ji, X.; Yan, M.; Mai, L.; Hu, P.; Shan, B.; Huang, Y., Na+ Intercalation Pseudocapacitance in Graphene­Coupled Titanium Oxide Enabling Ultra­Fast Sodium Storage and Long­Term Cycling. Nat Commun 2015, 6, 6929. (27) Wang, H.; Zhang, Y.; Ang, H.; Zhang, Y.; Tan, H. T.; Zhang, Y.; Guo, Y.; Franklin, J. B.; Wu, X. L.; Srinivasan, M.; Fan, H. J.; Yan, Q. A High­Energy Lithium­Ion Capacitor by Integration of a 3D Interconnected Titanium Carbide Nanoparticle Chain Anode with a Pyridine­Derived Porous Nitrogen­Doped Carbon Cathode. Adv. Funct. Mater. 2016, 26, 3082­3093. (28) Shen, L.; Lv, H.; Chen, S.; Kopold, P.; van Aken, P. A.; Wu, X.; Maier, J.; Yu, Y. Peapod­like Li3VO4 /N­Doped Carbon Nanowires with Pseudocapacitive Properties as Advanced Materials for High­Energy Lithium­Ion Capacitors. Adv Mater 2017, 29,1700142. (29) Kim, H.­S.; Cook, J. B.; Lin, H.; Ko, Jesse S.; Tolbert, Sarah H.; Ozolins, V.; Dunn, B. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3­x. Nat. Mater., 2016, 16, 454. (30) Jiang, Y.; Li, Y.; Zhou, P.; Lan, Z.; Lu, Y.; Wu, C.; Yan, M. Ultrafast, Highly Reversible, and Cycle­Stable Lithium Storage Boosted by Pseudocapacitance in Sn­Based Alloying Anodes. Adv Mater 2017, 29,1606. (31) Liu, X.; Zeng, J.; Yang, H.; Zhou, K.; Pan, D. V2O5­Based Nanomaterials:Synthesis and Their Applications. RSC Adv. 2018, 8, 4014­4031. (32) Cui, C.­J.; Wu, G.­M.; Shen, J.; Zhou, B.; Zhang, Z.­H.; Yang, H.­Y.; She, S.­Y. Synthesis and Electrochemical Performance of Lithium Vanadium Oxide Nanotubes as Cathodes for Rechargeable Lithium­Ion Batteries. Electrochim. Acta 2010, 55, 2536­2541. (33) Dobley, A.; Ngala, K.; Yang, S.; Zavalij, P. Y.; Whittingham, M. S.

Manganese Vanadium

Oxide Nanotubes:  Synthesis, Characterization, and Electrochemistry. Chem. Mater. 2001, 13, 4382­4386. 20 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

(34) Zhou, X.; Wu, G.; Gao, G.; Wang, J.; Yang, H.; Wu, J.; Shen, J.; Zhou, B.; Zhang, Z. Electrochemical Performance Improvement of Vanadium Oxide Nanotubes as Cathode Materials for Lithium Ion Batteries through Ferric Ion Exchange Technique. J. Phys. Chem. C 2012, 116, 21685­21692. (35) Zhou, X.; Chen, X.; He, T.; Bi, Q.; Sun, L.; Liu, Z. Fabrication of Polypyrrole/Vanadium Oxide Nanotube Composite with Enhanced Electrochemical Performance as Cathode in Rechargeable Batteries. Appl. Surf. Sci. 2017, 405, 146­151. (36) Zakharova, G. S.; Liu, Y.; Enyashin, A. N.; Yang, X.; Zhou, J.; Jin, W.; Chen, W. Metal Cations Doped Vanadium Oxide Nanotubes: Synthesis, Electronic Structure, and Gas Sensing Properties. Sens. Actuators B: Chem. 2018, 256, 1021­1029. (37) Sun, D.; Kwon C., W.; Baure, G.; Richman, E.; Maclean, J.; Dunn, B.; Tolbert S., H. The Relationship Between Nanoscale Structure and Electrochemical Properties of Vanadium Oxide Nanorolls. Adv. Funct. Mater. 2010, 14, 1197­1204. (38) Mcnulty, D.; Buckley, D. N.; O'Dwyer, C. V2O3 Polycrystalline Nanorod Cathode Materials for Li­Ion Batteries with Long Cycle Life and High Capacity Retention. ChemElectroChem 2017, 4, 2037­2044. (39) Mcnulty, D.; Buckley, D. N.; O'Dwyer, C. Optimizing the Structure and Yield of Vanadium Oxide Nanotubes by Periodic 2D Layer Scrolling. RSC Adv. 2016, 6, 40932­40944. (40) Kim, H.; Kim, D. Y.; Kim, Y.; Lee, S.­S.; Park, K. Na Insertion Mechanisms in Vanadium Oxide Nanotubes for Na­Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 1477­1485. (41) Nordlinder, S.; Lindgren, J.; Gustafsson, T.; Edström, K. The Structure and Electrochemical Performance of Na+­ ,   K+­ ,  and Ca2+­Vanadium Oxide Nanotubes. J. Electrochem. Soc. 2003, 150, E280­E284. (42) Toby, B. H.; Dreele, R. B. V. GSAS­II: The Genesis of a Modern Open­Source All Purpose Crystallography Software Package, J. Appl. Crystallogr. 2013, 46, 544­549. (43) He, P.; Zhang, G.; Liao, X.; Yan, M.; Xu, X.; An, Q.; Liu, J.; Mai, L. Sodium Ion Stabilized Vanadium Oxide Nanowire Cathode for High­Performance Zinc­Ion Batteries. Adv. Energy Mater. 2018, 8, 1702463.

21 ACS Paragon Plus Environment

Page 22 of 33

Page 23 of 33 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

(44) Nagaraju, G.; Chandrappa, G. T. Solution Phase Synthesis of Na0.28V2O5 Nanobelts into Nanorings and the Electrochemical Performance in Li Battery. Mater. Res. Bull 2012, 47, 3216­3223. (45) Nagaraju, G.; Sarkar, S.; Dupont, J.; Sampath, S. Na0.33V2O5·1.5H2O Nanorings/Nanorods and Na0.33V2O5·1.5H2O/RGO Composite Fabricated by a Facile One Pot Synthesis and Its Lithium Storage Behavior. Solid State Ionics 2012, 227, 30­38. (46) Niu, C.; Liu, X.; Meng, J.; Xu, L.; Yan, M.; Wang, X.; Zhang, G.; Liu, Z.; Xu, X.; Mai, L. Three Dimensional V2O5/NaV6O15 Hierarchical Heterostructures: Controlled Synthesis and Synergistic Effect Investigated by In situ X­ray Diffraction. Nano Energy 2016, 27, 147­156. (47) Baddour­Hadjean, R.; Boudaoud, A.; Bach, S.; Emery, N.; Pereira­Ramos, J.­P. A Comparative Insight of Potassium Vanadates as Positive Electrode Materials for Li Batteries: Influence of the Long­Range and Local Structure. Inorg. Chem. 2014, 53, 1764­1772. (48) Popović, Z. V.; Gajić, R.; Konstantinović, M. J.; Provoost, R.; Moshchalkov, V. V.; Vasil’ev, A. N.; Isobe, M.; Ueda, Y. Infrared and Raman Spectra of LiV2O5 Single Crystals, Phys. Rev. B 2000, 61, 70740­11459. (49) Smirnov, M. B.; Roginskii, E. M.; Kazimirov, V. Y.; Smirnov, K. S.; Baddour­Hadjean, R.; Pereira­Ramos, J. P.; Zhandun, V. S. Spectroscopic and Computational Study of Structural Changes in γ­LiV2O5 Cathodic Material Induced by Lithium Intercalation. J. Phys. Chem. C 2015,119, 20801­20809. (50) Niemann, R. G.; Kontos, A. G.; Palles, D.; Kamitsos, E. I.; Kaltzoglou, A.; Brivio, F.; Falaras, P.; Cameron, P. J. Raman Scattering Study of β­Sr0.33V2O5 in Charge Disordered and Ordered Phase, J. Physics: Condens. Mat. 2006, 18, 7779­7787. (51) Sun, K.; Yu, S.; Hu, Z.; Li, Z.; Lei, G.; Xiao, Q.; Ding, Y. Oxygen­Containing Hierarchically Porous Carbon Materials Derived From Wild Jujube Pit for High­Performance Supercapacitor. Electrochim. Acta 2017, 231, 417­428. (52) Sarkar, A.; Sarkar, S.; Sarkar, T.; Kumar, P.; Bharadwaj, M. D.; Mitra, S. Rechargeable Sodium­Ion Battery: High­Capacity Ammonium Vanadate Cathode with Enhanced Stability at High Rate. ACS Appl. Mater. Interfaces 2015, 7, 17044­17053.

22 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

Page 24 of 33

(53) Fu, L.; Tang, K.; Oh, H.; Manickam, K.; Bräuniger, T.; Chandran, C. V.; Menzel, A.; Hirscher, M.; Samuelis, D.; Maier, J. "Job­sharing" storage of hydrogen in Ru/Li2O nanocomposites. Nano Lett. 2015, 15, 4170­4175. (54) Yuan, S.; Zhao, Y.; Wang, Q. Layered Na2V6O16 Nanobelts as Promising Cathode and Symmetric Electrode for Na­Ion Batteries with High Capacity. J. Alloys Compd. 2016, 688, 55­60. (55) Raju, V.; Rains, J.; Gates, C.; Luo, W.; Wang, X.; Stickle, W. F.; Stucky, G. D.; Ji, X. Superior cathode of sodium­ion batteries: Orthorhombic V2O5 Nanoparticles Generated in Nanoporous Carbon by Ambient Hydrolysis Deposition. Nano Lett. 2012, 14, 4119­4124. (56) Uchaker, E.; Zheng, Y. Z.; Li, S.; Candelaria, S. L.; Hu, S.; Cao, G. Z.

Better than

Crystalline: Amorphous Vanadium Oxide for Sodium­Ion Batteries. J. Mater. Chem. A 2014, 2, 18208­18214. (57) Li, S.; Li, X.; Li, Y.; Yan, B.; Song, X.; Fan, L.; Shan, H.; Li, D. Design of V2O5·xH2O Cathode for Highly Enhancing Sodium Storage. J. Alloys Compd. 2017, 722, 278­286. (58) Li, S.; Li, X.; Li, Y.; Yan, B.; Song, X.; Li, D. Superior Sodium Storage of Vanadium Pentoxide Cathode with Controllable Interlamellar Spacing. Electrochim. Acta 2017, 244, 77­85 . (59) Wang, H.; Bi, X.; Bai, Y.; Wu, C.; Gu, S.; Chen, S.; Wu, F.; Amine, K.; Lu, J. Open­Structured V2O5·nH2O Nanoflakes as Highly Reversible Cathode Material for Monovalent and Multivalent Intercalation Batteries. Adv. Energy Mater. 2017, 7, 1602720. (60) Li, H.­Y.; Yang, C.­H.; Tseng, C.­M.; Lee, S.­W.; Yang, C.­C.; Wu, T.­Y.; Chang, J.­K. Chang, Electrochemically Grown Nanocrystalline V2O5 as High­Performance Cathode for Sodium­Ion Batteries. J. Power Sources 2015, 285, 418­424. (61) Su, D.; Wang, G. Single­Crystalline Bilayered V2O5 Nanobelts for High­Capacity Sodium­Ion Batteries. ACS Nano 2013, 7, 11218­11226. (62) Su, D. W.; Dou, S. X.; Wang, G. X. Hierarchical Orthorhombic V2O5 Hollow Nanospheres as High Performance Cathode Materials for Sodium­Ion Batteries, J. Mater. Chem. A 2014, 2, 11185­11194. (63) Gogotsi, Y.; Penner, R. M. Energy Storage in Nanomaterials­Capacitive, Pseudocapacitive, or Battery­Like? ACS Nano 2018, 12, 2081­2083. (64) Simon, P.; Gogotsi, Y.; Dunn, B. Where Do Batteries End and Supercapacitors Begin? Science 2014, 343, 1210­1211. 23 ACS Paragon Plus Environment

Page 25 of 33 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

(65) Lesel, B. K.; Cook, J. B.; Yan, Y.; Lin, T. C.; Tolbert, S. H. Using Nanoscale Domain Size to Control Charge Storage Kinetics in Pseudocapacitive Nanoporous LiMn2O4 powders. ACS Energy Lett. 2017, 2, 2293­2298. (66) Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive Oxide Materials for High­rate Electrochemical Energy Storage, Energy Environ. Sci. 2014, 7, 1597­1614. (67) Wei, Q.; Liu, J.; Feng, W.; Sheng, J.; Tian, X.; He, L.; An, Q.; Mai, L. Hydrated Vanadium Pentoxide with Superior Sodium Storage Capacity, J. Mater. Chem. A 2015, 3, 8070­8075. (68) Wang, H.; Zhu, C.; Chao, D.; Yan, Q.; Fan, H. J. Nonaqueous Hybrid Lithium­Ion and Sodium­Ion Capacitors. Adv. Mater. 2017, 29, 1702093. (69) Chen, Z.; Augustyn, V.; Jia, X.; Xiao, Q.; Dunn, B.; Lu, Y. High­Performance Sodium­Ion Pseudocapacitors Based on Hierarchically Porous Nanowire Composites. ACS Nano 2012, 6, 4319­4327. (70) Deng, B.; Lei, T.; Zhu, W.; Xiao, L.; Liu, J. In­Plane Assembled Orthorhombic Nb2O5 Nanorod Films with High­Rate Li+ Intercalation for High­Performance Flexible Li­Ion Capacitors. Adv. Funct. Mater. 2017,28,1704330. (71) Xie, L.; Yang, Z.; Sun, J.; Zhou, H.; Chi, X.; Chen, H.; Li, A. X.; Yao, Y.; Chen, S. Bi2Se3/C Nanocomposite as a New Sodium­Ion Battery Anode Material. Nano-Micro Lett. 2018,10, 50. (72) Wei, Q.; Jiang, Z.; Tan, S.; Li, Q.; Huang, L.; Yan, M.; Zhou, L.; An, Q.; Mai, L. Lattice Breathing Inhibited Layered Vanadium Oxide Ultrathin Nanobelts for Enhanced Sodium Storage. ACS Appl. Mater. Interfaces 2015, 7, 18211­18217. (73) Yu, S.; Peng, C.; Li, Z.; Zhang, L.; Xiao, Q.; Lei, G.; Ding, Y. K­Doped Li­Rich Molybdenum­Based Oxide with Omproved Electrochemical Properties for Lithium­Ion Batteries, Arabian J. Sci. Eng. 2017, 42, 4291­4298.

24 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

Scheme 1. Schematic illustration of the formation process of the Na­VBNT@C sample.

25 ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33 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. TEM (a), HRTEM images (b, c) and SAED pattern (d) of the Na­VBNT sample. TEM (e), HRTEM images (f, g) and SAED patterns (h) of the Na­VBNT@C sample.

26 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 2. XRD pattern and Retrieved refinement result (a), crystal structure model (b) of the Na­VBNT. XRD patterns (c) and Raman spectra (d) of the VONT, Na­VBNT, Na­VBNT@C samples. High­resolution V2p XPS spectra of the Na­VBNT (e) and Na­VBNT@C (f) samples.

27 ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33 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 3. Galvanostatic charge­discharge curves of the VONT (a), Na­VBNT (b) and Na­VBNT@C (c) electrodes at 0.1C rate (25 mA g­1) for various cycles. The voltages profiles at the maximum capacity when cycled at different rate from 0.1C to 10C(d). Cycling performances of the VONT, Na­VBNT and Na­VBNT@C electrodes at 0.1C rate (25 mA g­1) (e). Rate capabilities of the VONT, Na­VBNT and Na­VBNT@C samples (f,g). Cycling performances of the Na­VBNT and Na­VBNT@C electrodes at 5C rate (h). Schematic illustration of the morphology of the Na­VBNT@C sample (i).All charge­discharge tests were carried out at room temperature over the voltage range of 1.5­4.0 V using the assembled button­type cells.

28 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. Cyclic voltammograms of the VONT, Na­VBNT and Na­VBNT@C electrodes at a sweep rate of 0.1 mV s­1 (a), and the Na­VBNT@C electrode at different sweep rates ranged from 0.1 to 0.5 mV s­1 (b). Capacitive contribution of the Na­VBNT@C electrode estimated from the CV curve at the sweep rate of 0.5 mV s­1 (c). Log (i)~log(v) plots of the VONT (d), Na­VBNT (e) and Na­VBNT@C (f) electrodes. Estimated values of capacitive contribution of the VONT (g), Na­VBNT (h) and Na­VBNT@C (i) electrodes at various sweep rates.

29 ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33 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 5. Ex­situ XRD patterns of the Na­VBNT@C electrode at the different charge states in the first cycle and after 100 cycles (a), and the corresponding enlarged XRD patterns belonging to the (11­1), (104), (21­3) and (31­7) planes (b).

30 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 6. Surface SEM images with the magnifications of 200 (a) and 10,000 (b) for the Na­VBNT@C electrode after cycled at 5C rate for 1000 times.

31 ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33 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 1. Lattice parameters refined from XRD patterns of the Na­VBNT@C electrode during the first cycle and after 100 cycles. Electrode state

Cell parameter a (Å)

b (Å)

c (Å)

β (o)

V (Å3)

ρ(g cm-3)

Fresh

10.102(2)

3.624(7)

15.456(7)

109.49(3)

533.57

3.5391

Discharged to 2.3 V

10.078(3)

3.617(4)

15.487(2)

110.77(8)

527.81

3.5777

Discharged to 1.5 V

10.084(4)

3.619(2)

15.758(4)

110.98(6)

536.97

3.5166

Charged to 2.6 V

10.021(3)

3.657(6)

15.571(1)

110.22(6)

533.04

3.5492

Charged to 3.1 V

10.098(8)

3.614(0)

15.507(8)

109.90(7)

532.26

3.5478

Charged to 4.0 V

10.124(3)

3.617(6)

15.458(1)

109.79(0)

531.66

3.5411

10.051(3)

3.622(2)

15.502(3)

110.50(4)

528.34

3.5741

After 100 cycles at fully charged

32 ACS Paragon Plus Environment