Next Article in Journal
The Roles of Nanomaterials in Conventional and Emerging Technologies for Heavy Metal Removal: A State-of-the-Art Review
Previous Article in Journal
Control of Nitrogen Inhomogeneities in Type-I and Type-II GaAsSbN Superlattices for Solar Cell Devices
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Hydrothermal Growth and Li+ Storage Performance of 1D VOx Nanobelts with Variable Vanadium Valence

1
Department of Physics, School of Physics and Astronomy, Yunnan University, Kunming 650504, China
2
Yunnan Key Laboratory of Micro/Nano-Materials and Technology, Yunnan University, Kunming 650504, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(4), 624; https://doi.org/10.3390/nano9040624
Submission received: 29 March 2019 / Revised: 12 April 2019 / Accepted: 15 April 2019 / Published: 17 April 2019

Abstract

:
One-dimensional (1D) vanadium oxide nanobelts (VOx NBs) with variable V valence, which include V3O7·H2O NBs, VO2 (B) NBs and V2O5 NBs, were prepared by a simple hydrothermal treatment under a controllable reductive environment and a following calcination process. Electrochemical measurements showed that all these VOx NBs can be adopted as promising cathode active materials for lithium ion batteries (LIBs). The Li+ storage mechanism, charge transfer property at the solid/electrolyte interface and Li+ diffusion characteristics for these as-synthesized 1D VOx NBs were systematically analyzed and compared with each other. The results indicated that V2O5 NBs could deliver a relatively higher specific discharge capacity (213.3 mAh/g after 50 cycles at 100 mA/g) and median discharge voltage (~2.68–2.71 V vs. Li/Li+) during their working potential range when compared to other VOx NBs. This is mainly due to the high V valence state and good crystallinity of V2O5 NBs, which are beneficial to the large Li+ insertion capacity and long-term cyclic stability. In addition, the as-prepared VO2 (B) NBs had only one predominant discharge plateau at the working potential window so that it can easily output a stable voltage and power in practical LIB applications. This work can provide useful references for the selection and easy synthesis of nanoscaled 1D vanadium-based cathode materials.

Graphical Abstract

1. Introduction

The efficient storage and release of electric energy is an important topic in the field of energy storage research. At present, two kinds of energy storing devices, that is, lithium ion batteries (LIBs) and supercapacitors (SCs), have received great attention because of their respective features in high energy density and power density. As one of the typical energy storage units, LIBs are widely applied in various consumer electronics and electric tools since they have the advantages of large energy density, low self-discharge rate, no memory effect, etc. Electrode materials are directly responsible for the electrochemical performance of LIBs. It is crucial for the development of cathode materials with large capacities, high working voltage and high rate capability. Traditional cathode materials for LIBs, such as LiCoO2, LiMn2O4, LiNixCoyMnzO2 and LiFePO4, which only possess a specific capacity of ~120–180 mAh/g, could not meet the requirement of high-energy density for next generation LIBs [1]. Thus, it is necessary to explore new-type cathodes with large capacity. Among the diverse transition metal oxides, vanadium oxide (VOx) could act as a kind of promising cathode candidate with relatively high operating voltage and capacity during the Li+ intercalation/de-intercalation process, which was noted very early by researchers and battery manufacturers [2,3].
There are a wide variety of configurations for VOx including V2O5, V3O7, VO2, V6O13, etc. To improve the electrochemical performance of the VOx matrix (mainly the conductivity and structural stability upon the repeated charge/discharge), the commonly adopted methods are the design of nanostructure and further modifications, like doping, as well as combination with conductive frameworks [3,4,5]. The reported nanostructures involve 0-Dimensional (0D) (nanosphere and nanoparticle), 1D (nanobelt, nanofiber, nanorod and nanowire), 2D (nano-film, nanoplate and nanosheet) and 3D hierarchical (nanoflower, nanourchin and mesoporous structure) VOx, which were produced through various methods, such as template synthesis, self-assembly, hydrothermal process, sol-gel, spray pyrolysis, electro-spinning and chemical vapor deposition (CVD) [4,5,6,7,8,9]. The electrochemical Li+ storage performance of these VOx nanostructures is significantly superior to those of their bulk counterparts and commercial cathodes. During a variety of VOx nanostructures, one-dimensional (1D) VOx has the advantages of convenient synthesis, structural simplicity and easy to scale-up production, which makes VOx prospective as alternative cathode materials for LIBs. In recent years, the solvothermal method was mainly applied to prepare 1D V3O7, VO2 and V2O5 nanobelts, whose specific capacity and Li+ storage properties were investigated under different working voltage windows [10,11,12,13,14,15]. However, it is rarely reported that a series of 1D nanostructured VOx, which possess variable components and crystal structures, were acquired in a similar synthetic environment just by altering reaction reagents and combining them with a simple post-treatment.
During our study, we propose an easy route to prepare 1D VOx nanostructures with different crystalline textures and vanadium valence states based on hydrothermal treatment. For one thing, 1D V3O7·H2O and VO2 (B) nanobelts (NBs) with relatively low V valence were obtained through the previous solation of V2O5 powder and introduction of reducing ingredients (ethanol and CNTs) under hydrothermal condition. For another, V2O5 NBs with high V valence were further achieved by simple post-sintering to VOx NBs in air under the circumstance of maintaining the 1D nanobelt shape. The microscopic morphologies, crystal lattices and V-O bonding of these 1D VOx NBs were characterized and analyzed. When these VOx NBs were used as active cathodes for LIBs, their Li+ storage properties were evaluated and compared systematically. According to this research, we get a simple method to obtain 1D VOx NBs with variable V valence on a large scale. In addition, the micro-architecture and electrochemical Li+ insertion/extraction performance of these VOx NBs were discussed and perceived.

2. Experimental Section

2.1. Sample Preparations

All reaction reagents were analytically pure grade and used as received.
Firstly, 1.82 g of V2O5 (Shanghai Kefeng industrial Co., Ltd., Shanghai, China) powder (referring to ESI, Figure S1) was dispersed into 75 ml of de-ionized water (self-produced by EPED-10TH apparatus from Nanjing Yipu Yida technology development Co. Ltd., Nanjing, China) with strong magnetic stirring, which formed a yellow suspension. Then, 5 ml of H2O2 (30%, mass concentration, Tianjin Fengchuan chemical reagent co. Ltd., Tianjin, China) was poured to the above V2O5 suspension under continuous agitation for 5 hours, generating translucent and orange V2O5 sol. After that, ~80 ml of V2O5 sol was further mixed with 0.2 ml of absolute ethanol (Tianjin Fengchuan) and transferred into a 100 ml of Teflon-lined hydrothermal autoclave (100 ml, Shanghai Lingtuo instrument and equipment Co., Ltd., Shanghai, China). The stainless steel reactor was sealed, kept at 180 °C for 2 days in an electric heating oven and cooled to room temperature naturally. The resultant black precipitate was rinsed, filtrated and dried at 80 °C. The sample obtained was labeled as V3O7·H2O nanobelts (NBs).
In order to study the effect of ethanol content on the final product, 2 ml of absolute ethanol was added to previously described V2O5 sol (80ml) and followed by hydrothermal treatment with identical operating procedures. The product was denoted as V3O7·H2O/VO2 (B)* NBs.
For further comparison, 2 ml of ethanol and 0.2 g of multi-walled carbon nanotubes (MWCNTs, Shenzhen Nanotech Port Co., Ltd., Shenzhen, China) were simultaneously introduced into the above mentioned V2O5 (80 ml) sol under ultrasonic oscillation, forming black turbid liquid that was also subjected to the same hydrothermal process. The MWCNTs used here were previously treated with the mixed concentrated acid (H2SO4:HNO3 = 3:1, volume ratio, Chengdu Kelong Chemical Reagent Factory, Chengdu, China) in order to create the surface functional groups and enhance their dispersibility in water. The collected powder has a dark black-blue color, which is labeled as VO2 (B) NBs/MWCNTs. The filling degree for autoclave during all hydrothermal reactions was controlled at 80%.
V2O5 NBs were prepared by directly sintering the previously acquired V3O7·H2O NBs in a tube furnace (OTF-1200X, Hefei Kejing-MTI technology Co., Ltd., Hefei, China) at 400 °C for 3 hours under air condition with the ramping rate of 5 °C/min.

2.2. Characterizations

A scanning electron microscope (SEM, QUANTA 200, FEI Company, Hillsboro, OR, USA) was employed to observe the surface morphologies of the as-prepared samples. The crystal structures of the products were determined by X-Ray Diffractometer (XRD, TTRIII-18KW, Rigaku, Tokyo, Japan) at the 2θ angle range of 10–90° (recording interval, 0.02°) with Cu Kα radiation (λ = 1.5406 Å). Transmission electron microscope (TEM, JEM-2100, JEOL, Akishima, Tokyo, Japan) equipped with the module for Selected area electron diffraction (SAED) was applied to examine the microstructures of samples. Material composition and atomic bonding for the products obtained were analyzed through Fourier transform infrared (FTIR) spectroscope with a Bruker-TENSOR27 (Bruker Optics, Karlsruhe, Germany) spectrometer at wave number region of 400–4000 cm−1 using KBr pellets.

2.3. Electrochemical Tests

The cathode slurry was composed of active material (V3O7·H2O NBs, V3O7·H2O/VO2 (B)* NBs, VO2 (B) NBs/MWCNTs or V2O5 NBs), conductive acetylene black (Shenzhen Kejing Zhida-MTI technology Co., Ltd., Shenzhen, China) and polyvinylidene fluoride (PVDF, Shenzhen Kejing), with a mass ratio of 7:2:1, using N-methyl-2-pyrrolidone (NMP, Shenzhen Kejing) as an organic solvent. The slurry was uniformly coated onto the Al foil (Shenzhen Kejing) by a doctor blade and dried at 120 °C over 8 h in a vacuum to wipe off the NMP completely. Then, the formative cathode plate was cut into circular discs of 12 mm in diameter for electrochemical performance evaluations and the average active material loaded for all electrodes was ~1.4 mg/cm2. Coin cells (CR-2025, Shenzhen Kejing) were assembled by using the cathode disc as a working electrode, a porous polyolefin tri-layer membrane (Celgard 2325, Shenzhen Kejing) as a separator and Li metal as reference/counter electrode. A 1 M LiFP6 dissolved in ethylene carbonate/dimethyl carbonate/ethyl methylcarbonate (EC/DMC/EMC, 1/1/1, volume ratio, Shenzhen Kejing) was adopted as an electrolyte. The battery assembly was performed in a glove box with inert Ar atmosphere where both H2O and O2 gas were below 1 ppm.
Cyclic voltammograms (CV) were recorded by an electrochemical workstation (CHI660, Shanghai Chenhua metrologic instruments Co., Ltd., Shanghai, China) at a scanning speed of 0.2 mV s−1 within the voltage region of 1.5–4 V or 2–4 V. Galvanostatic charge/discharge (GCD) tests were conducted using LAND (CT2001A, Wuhan Landian electronic Co., Ltd., Wuhan, China) cell-testing system between 1.5–4 V or 2–4 V under specific current densities. Electrochemical impedance spectroscopy (EIS) was also collected by CHI660 electrochemical workstation at the frequency region from 100 k to 0.01 Hz under the state of charge (SOC) of ~3 V with an alternating current (AC) signal of 5 mV. Nyquist plots were fitted by Z-view software.

3. Result and Discussion

Figure 1 shows the synthetic process for V3O7·H2O NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs, whose microscopic molecular crystal structures are illustrated in the insets. The V3O7·H2O NBs with V4+ and V5+ coexisting was prepared through hydrothermal reaction using V2O5 sol as a precursor and ethanol as a weak reducing agent. When increasing the amount of ethanol and simultaneously introducing MWCNTs as an auxiliary reducing ingredient, VO2 (B) NBs with predominant V4+ were obtained at the same condition. After direct sintering in the air at 400 °C for 3 h, V3O7·H2O NBs were transformed into V2O5 NBs with complete V5+.
XRD patterns of all nanostructured VOx samples prepared under different conditions are presented in Figure 2. Through the crystal structure analysis, the pattern of Figure 2a could be ascribed to the orthorhombic phase of V3O7·H2O (PDF# 28-1433) with characteristic peaks of (020), (120), (130), etc. Two sets of diffraction peaks can be deconstructed from Figure 2b, which belongs to the orthorhombic V3O7·H2O (major component) and monoclinic VO2 (B) (minor component with asterisk mark, PDF# 31-1438) with peaks of (200), (110), (−311), etc., respectively. This indicates that the ethanol, as a weak reducing agent, would lead to the conversion of V5+ to V4+ and meanwhile be oxidized to aldehyde itself. When maintaining a relatively high ethanol content, like the sample V3O7·H2O/VO2 (B)* NBs, and bringing in MWCNTs, the monoclinic VO2 (B) was primarily acquired as shown in Figure 2c. This result shows that MWCNTs would play the role of inductive reducibility during hydrothermal treatment, which may owe to the oxygen-contained functional groups on the surface of MWCNTs generated during the mixed acid treatment and hydrothermal reaction. There is a broad peak envelope at 2θ ≈ 26° in Figure 2c, which was caused by the (002) plane of amorphous MWCNTs during the sample VO2 (B) NBs/MWCNTs [16,17]. For sample V2O5 NBs, whose XRD pattern in Figure 2d confirms its orthorhombic V2O5 phase (PDF# 41-1426), with a typical crystal plane of (200), (001), (101), etc., illustrating the transformation of V4+ to V5+ when V3O7·H2O was subject to sintering in air.
SEM images of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) are shown in Figure 3. It can be seen that all of them display one-dimensional nanobelted morphology with a width of 100–300 nm and a length of over 10 microns. Due to the role of energy modulation under hydrothermal conditions, the growth kinetics of the vanadium oxide precursor is faster in a specific direction during the process of self-assembly crystallization and, conversely, the growth in other orientations are inhibited, resulting in a one-dimensional VOx nano-belted morphology. In Figure 3c, VO2 (B) NBs were evenly mingled with MWCNTs. In addition, SEM images of VO2 (B) NBs/MWCNTs (Figure 3c) and V2O5 NBs (Figure 3d) also demonstrate that the adding of MWCNTs or post-sintering will not change the 1D nanostructure of VOx samples, but result in the further decrease or increase of V valence as recognized by the XRD detections.
Figure 4 illustrates TEM and HRTEM (High-Resolution TEM) images of V3O7·H2O NBs (a,b), VO2 (B) NBs/MWCNTs (c,d) and V2O5 NBs (e,f). We can see from the TEM images (a,c,e) that the diameter of these samples distributes approximately between 100–300 nm, which is in accordance with previous SEM observations. Figure 4b gives the lattice fringes of (101) and (200) planes with distances of 0.338 nm and 0.466 nm for V3O7·H2O NBs [18,19]. The inset shows the corresponding SAED pattern. The HRTEM image of VO2 (B), whose (−201) crystal plane is marked with the space of ~0.5 nm, is presented in Figure 4d. Figure 4f demonstrates the (110) crystal plane for V2O5 NBs with a distance of 0.341 nm, as well as its SAED pattern in the upper right corner [20,21]. Thus, it could be learned that the results of TEM characterizations are consistent with those values from the calculations of the associated XRD peak positions based on the Scherrer formula.
FTIR spectra of the as-prepared four samples are shown in Figure 5. The characteristic absorption peaks near 1000 cm−1 (976, 980, 1007, 1016, 1020 and 1022 cm−1) can be attributed to stretching vibrations of terminal oxygen bonds (V=O). It should be noted that the appearance of relatively low wave numbers (976, 980 and 1007 cm−1) for V=O vibrations is caused by the V4+ in V3O7·H2O NBs, V3O7·H2O/VO2 (B)* NBs and VO2 (B) NBs/MWCNTs because the bond length of V4+=O is longer than that of V5+=O, leading to the decrease of frequency [22]. The peaks at 802 cm−1 and below 600 cm−1 (449, 480, 546, 557, 567 and 577 cm−1) emerge because of the vibration modes of the doubly coordinated oxygen (V-O-V) bonds in V2O5 and the triply coordinated oxygen (V3–O) bonds within VOx, respectively [23]. The slight shifting for these V–O coordinated vibration peaks among different samples may be connected with the microscopic stress change in the crystal lattice. Besides, the absorption peaks at 1628–1633 cm−1 and 3412–3444 cm−1 could be ascribed to the bending and stretching vibrations of H–O bonds from the absorbed/bound water in samples, respectively [24,25]. The wave-number at ~2350 cm−1 is due to the contamination of CO2 molecules in the measuring atmosphere [26].
The initial three CV curves of the four samples between 1.5–4 V are depicted in Figure 6.
For V3O7·H2O NBs, four pairs of stable cathodic/anodic peaks appear at ~3.30/3.37, 3.09/3.21, 2.56/2.92 and 1.80/2.11 V, in which two pairs of redox peaks at ~3.30/3.37 and 3.09/3.21 V between 3.0–3.5 V correspond to the reaction:
V3O7·H2O + Li+ + e ↔ LiV3O7·H2O.
The other pair of peaks at ~2.56/2.92 V between 2.3–3.0 V can be ascribable to the reaction:
LiV3O7·H2O + Li+ + e ↔ Li2V3O7·H2O.
Another pair of peaks at ~1.80/2.11 V in the region of 1.5–2.3 V is related with the reaction:
Li2V3O7·H2O + xLi+ + xe ↔ Li2+xV3O7·H2O (x < 1).
Both Equations (1) and (2) involve the conversion of V5+↔V4+, whereas Equation (3) refers to the partial V4+↔V3+ [27,28]. The CV curves of VO2 (B) NBs/MWCNTs are shown in Figure 6c. One pair of predominant redox peaks emerges at ~2.48/2.78 V, which is the typical trait for the monoclinic VO2 (B) phase during the repeated Li+ insertion/extraction, corresponding to the reaction [29]:
VO2 (B) + xLi+ + xe ↔ LixVO2 (B) (x ≈ 0.5)
According to the analysis of a previous XRD, we know that V3O7·H2O/VO2 (B)* NBs contain a small quantity of the VO2 (B) phase. Thus, the overall outline of the CV curves for V3O7·H2O/VO2 (B)* NBs in Figure 6b is similar to that of the V3O7·H2O NBs, but there is an upward anodic shoulder peak appearing at ~2.76 V and a slightly lower shift for the downward cathodic peak from 2.56 to 2.52 V. Further, the relative peak area between 2.3–3.0 V for V3O7·H2O/VO2 (B)* NBs is larger compared to the V3O7·H2O NBs. These changes are concerned with the redox process of V4+↔V3+ in the VO2 (B) phase contained. The solid and dotted lines in Figure 6d present the CV curves of the V2O5 NBs at the voltage sweep range of 1.5–4 V and 2–4 V, respectively. During the initial Li+ intercalation between 1.5–4 V, the cathodic peaks at ~3.33, 3.08 and 2.14 V correspond to the phase transitions of α-LixV2O5 to ε-LixV2O5, ε-LixV2O5 to δ-LixV2O5 (x < 1) and δ-LixV2O5 to γ-LixV2O5 (1 < x < 2), respectively. When the working voltage lowers below 2 V, the irreversible phase transition happens from γ-LixV2O5 to ω-LixV2O5(x > 2) [30,31]. Therefore, the anodic peaks associated with the phase transitions above do not arise in the subsequent Li+ extraction process. There is only a wide anodic peak at ~2.71 during the following 2nd–3rd cycles, illustrating the amorphous features and irreversible structural change of V2O5 NBs operating in this voltage region (1.5–4 V). When V2O5 NBs work at 2–4 V, three pairs of primary redox peaks appear at ~3.33/3.45, 3.08/3.35 and 2.14/2.56 V, indicating the reversible phase transition between α-LixV2O5 ↔ γ-LixV2O5 (x < 2) [31,32]. Thus, the subsequent GCD evaluation for V2O5 NBs was performed between 2–4 V.
GCD profiles of four samples are plotted in Figure 7 at the 1st, 2nd, 5th, 20th and 50th charge/discharge cycles under 100 mA/g. Because VOx NBs non-embedded by Li+ are at a full charge state in the beginning, there is only the 1st discharge process for all samples. With regard to the V3O7·H2O NBs in Figure 7a, four voltage plateaus emerged in sequence at ~3.3, 3.1, 2.6 and 1.8 V during the discharge (Li+ intercalation), which are related with the four downward cathodic peaks in its CV curve. Conversely, the voltage plateaus turn up at ~2.1, 2.8, 3.2 and 3.4 V sequentially during the charge (Li+ de-intercalation), corresponding to its upward anodic peaks in the CV. For the VO2 (B) NBs/MWCNTs in Figure 7c, there was an obvious discharge and charge plateau at 2.5 V and 2.6 V, respectively, which are concerned with the pair of dominant redox peaks in its CV. The GCD profile of the V3O7·H2O/VO2 (B)* NBs in Figure 7b is similar to that of V3O7·H2O NBs. However, the discharge plateau at ~2.5 V was elongated visibly. Further, the charge plateau at ~2.8 V was also lengthened and split into two close plateaus. These variations in GCD for the V3O7·H2O/VO2 (B)* NBs are consistent with the changes in its CV curves. As for the V2O5 NBs in Figure 7, when working between 2–4 V, the stable GCD profile was formed through the first few discharge/charge activations, whose reversible plateaus correspond to the prime redox peaks in its CV (dotted line). From the standpoint of median discharge voltage (MDV), V3O7·H2O NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs show the MDV values of ~2.61–2.70 V, ~2.50 V and ~2.68–2.71 V during their operating voltage ranges, respectively.
The cycling performances of the four samples at 100 mA/g are successively displayed in Figure 8a–d. V3O7·H2O NBs in Figure 8a delivered a highest specific capacity of 219.8 mAh/g at the first discharge and kept at 181.9 mAh/g after 50 cycles. As shown in Figure 8b, V3O7·H2O/VO2 (B)* NBs presented the highest discharge capacity of 222.2 mAh/g for the first cycle and remained at 176.2 mAh/g on the 50th discharge. The VO2 (B) NBs/MWCNTs in Figure 8c gave the largest value of 178.8 mAh/g at the first discharge and maintained at 148.9 mAh/g for the 50th cycle. Although the specific capacity of VO2 (B) NBs is unsatisfactory, its charge/discharge plateau is very flat. During practical usage, VO2 (B) NBs will output stable voltage and power. The highest capacity of 240.4 mAh/g was released at the third discharge for V2O5 NBs after initial two activation cycles in Figure 8d. When the cycling number went to the 50th cycle, this value remained at 213.3 mAh/g. We can see from the above cycling diagrams that the capacities for all samples exhibit periodic fluctuation with cyclic numbers, which are due to the large diurnal temperature variation between 10–21 °C in the local lab. This result shows that temperature has a non-negligible effect on the Li+ storage property for all VOx samples. It is noteworthy that V2O5 NBs possesses higher capacity and cycling stability owing to their complete V5+ and better crystallinity. Thus, the rate capability of V2O5 NBs was further evaluated, as illustrated in Figure 8e. The typical CD curves at the last cycle under different current densities of 50, 100, 200, 400 and back to 50 mA/g are shown in Figure 8f, whose corresponding discharge capacities were 237.1, 217.1, 199.9, 184.9 and 221.8 mAh/g, respectively.
Nyquist plots and Warburg coefficients (Aw) linear fittings of four samples at the SOC of ~3 V are illustrated in Figure 9a,b, respectively. Nyquist plots are composed of a compressed semicircle arc at a high-frequency region and an oblique straight line at low-frequency region, in which the X-axis intercept of semicircle on the extreme left denotes the equivalent series resistance (Re), the span of the semicircle is associated with the charge transfer resistance (Rct) at the electrolyte/electrode interface and the sloping line is connected with the Li+ diffusion in the solid electrode [33]. Based on the equivalent circuit (in which CPE is the constant-phase element, Wo represents the Warburg impedance) in the inset of Figure 9a, we can get the fitted values of Re and Rct for four samples, as listed in Table 1. All of their Re values are small and distribute at 4–6 Ω. The Rct of VO2 (B) NBs/MWCNTs is the lowest, which could be due to incorporation of conducting MWCNTs as well as the existence of predominant V4+ in the sample [34,35]. The lower Rct is beneficial to the migration and transport of Li+ at the electrolyte/electrode interface.
To discuss the diffusion of Li+ during the active VOx NBs, we need to access the Li+ diffusion coefficient (DLi+) of the samples, which can be expressed as [36]:
D Li + = R 2 T 2 2 n 4 F 4 A 2 C 2 A w 2
where, R (ideal gas constant), T (absolute temperature), n (electron transfer number per mole), F (Faraday’s constant), A (geometric area of electrode) and C (molar concentration of Li+) are fixed experimental parameters under certain sample and test condition. Aw is the Warburg coefficient that is related to the Li+ diffusion process. Because DLi+ ∝ Aw−2, we can use the value of Aw to quantitatively evaluate the Li+ diffusion. According to the equation Z′ = Re + Rct + Awω−1/2 (6) at the low-frequency region, a linear relationship between Z′ and ω−1/2 can be observed for all samples, as shown in Figure 9b. The slope of a straight line is the value of Aw as listed in Table 1, which would be acquired by linear fitting. The results show that V3O7·H2O/VO2 (B)* NBs and V2O5 NBs exhibit a relatively lower Aw at the given SOC, signifying the preferable diffusion behavior.

4. Conclusions

In this work, three kinds of 1D VOx NBs, referring to V3O7·H2O NBs, VO2 (B) NBs and V2O5 NBs, which have different crystalline structure and vanadium valence, were synthesized under hydrothermal conditions merely by adjusting and adding the reductive ingredients (ethanol and MWCNTs), as well as conducting post-sintering treatment. XRD and SEM demonstrated the crystal structures and morphologies of acquired VOx NBs (showing a typical width of 100–300 nm and over 10 microns in length). TEM and FTIR revealed their micro-architectures and components. The electrochemical Li+ storage properties of these VOx NBs were systematically analyzed and compared with each other as cathode materials for LIBs. Among them, V2O5 NBs exhibited a relatively larger capacity, higher median discharge voltage and more stable cycling performance, which can be attributed to its higher V valence state and better crystallization. From the perspective of actual use, although VO2 (B) NBs showed an unsatisfactory specific capacity (148.9 mAh/g after 50 cycles at 100 mA/g), it displayed just one flat discharge platform at the testing potential window, which benefits the stable output of working voltage and power in practical battery applications. These representative 1D VOx NBs derived from a simple hydrothermal reaction could be used as promising cathode candidates for future LIBs according to actual demands.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/9/4/624/s1, Figure S1. SEM images of commercial V2O5 powder at 1000 (a) and 5000 (b) times magnification, respectively.

Author Contributions

Y.J. and X.Z. carried out the experiments and wrote the manuscript together. X.Z. conceived the experimental procedure. X.C. and J.W. provided useful help in electrochemical tests. L.G. and M.S. gave necessary assistance for the characterization of samples. Y.R. and Z.L. put forward valuable advices for the improvement of article.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 61664009, 51771169), Young/Middle-aged Backbone Teacher Cultivating Scheme of Yunnan University (Grant No. C176220200) and High-end Scientific & Technological Talents Introduction Project of Yunnan Province (Grant No. 2013HA019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manthiram, A. An Outlook on Lithium Ion Battery Technology. ACS Cent. Sci. 2017, 3, 1063–1069. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Y.; Cao, G. Synthesis and Enhanced Intercalation Properties of Nanostructured Vanadium Oxides. Cheminform 2006, 37, 2787–2804. [Google Scholar]
  3. McNulty, D.; Buckley, D.N.; O’Dwyer, C. Synthesis and electrochemical properties of vanadium oxide materials and structures as Li-ion battery positive electrodes. J. Power Sources 2014, 267, 831–873. [Google Scholar] [CrossRef] [Green Version]
  4. Liu, X.; Zeng, J.; Yang, H.; Zhou, K.; Pan, D. V2O5-Based nanomaterials: Synthesis and their applications. RSC Adv. 2018, 8, 4014–4031. [Google Scholar] [CrossRef]
  5. Liu, Y.; Wang, Y.; Zhang, Y.; Liang, S.; Pan, A. Controllable Preparation of V2O5/Graphene Nanocomposites as Cathode Materials for Lithium-Ion Batteries. Nanoscale Res. Lett. 2016, 11, 549. [Google Scholar] [CrossRef]
  6. Armer, C.F.; Yeoh, J.S.; Li, X.; Lowe, A. Electrospun vanadium-based oxides as electrode materials. J. Power Sources 2018, 395, 414–429. [Google Scholar] [CrossRef]
  7. Armstrong, M.J.; Burke, D.M.; Gabriel, T.; O’Regan, C.; O’Dwyer, C.; Petkov, N.; Holmes, J.D. Carbon nanocage supported synthesis of V2O5 nanorods and V2O5/TiO2 nanocomposites for Li-ion batteries. J. Mater. Chem. A 2013, 1, 12568. [Google Scholar] [CrossRef]
  8. Liu, M.; Su, B.; Tang, Y.; Jiang, X.; Yu, A. Recent Advances in Nanostructured Vanadium Oxides and Composites for Energy Conversion. Adv. Energy Mater. 2017, 7, 1700885. [Google Scholar] [CrossRef] [Green Version]
  9. Petnikota, S.; Chua, R.; Zhou, Y.; Edison, E.; Srinivasan, M. Amorphous Vanadium Oxide Thin Films as Stable Performing Cathodes of Lithium and Sodium-Ion Batteries. Nanoscale Res. Lett. 2018, 13, 363. [Google Scholar] [CrossRef]
  10. Li, S.; Liu, G.; Liu, J.; Lu, Y.; Yang, Q.; Yang, L.Y.; Yang, H.R.; Liu, S.; Lei, M.; Han, M. Carbon fiber cloth@VO2 (B): Excellent binder-free flexible electrodes with ultrahigh mass-loading. J. Mater. Chem. A 2016, 4, 6426–6432. [Google Scholar] [CrossRef]
  11. Li, Z.; Sun, H.; Xu, J.; Zhu, Q.; Chen, W.; Zakharova, G.S. The synthesis, characterization and electrochemical properties of V3O7·H2O/CNT Nanocomposite. Solid State Ion. 2014, 262, 30–34. [Google Scholar] [CrossRef]
  12. Niu, C.; Li, J.; Jin, H.; Shi, H.; Zhu, Y.; Wang, W.; Cao, M. Self-template processed hierarchical V2O5 nanobelts as cathode for high performance lithium ion battery. Electrochim. Acta 2015, 182, 621–628. [Google Scholar] [CrossRef]
  13. Rui, X.; Sim, D.; Xu, C.; Liu, W.; Tan, H.; Wong, K.; Hng, H.H.; Lim, T.M.; Yan, Q. One-pot synthesis of carbon-coated VO2(B) nanobelts for high-rate lithium storage. RSC Adv. 2012, 2, 1174–1180. [Google Scholar] [CrossRef]
  14. Yu, Y.; Li, J.; Wang, X.; Chang, B.; Wang, J.; Ahmad, M.; Sun, H. Oxygen vacancies enhance lithium storage performance in ultralong vanadium pentoxide nanobelt cathodes. J. Colloid Interface Sci. 2018, 539, 118–125. [Google Scholar] [CrossRef]
  15. Zhang, C.; Song, H.; Zhang, C.; Liu, C.; Liu, Y.; Cao, G. Interface Reduction Synthesis of H2V3O8 Nanobelts–Graphene for High-Rate Li-Ion Batteries. J. Phys. Chem. C 2015, 119, 11391–11399. [Google Scholar] [CrossRef]
  16. Baro, M.; Nayak, P.; Baby, T.T.; Ramaprabhu, S. Green approach for the large-scale synthesis of metal/metal oxidenanoparticle decorated multiwalled carbon nanotubes. J. Mater. Chem. A 2013, 1, 482–486. [Google Scholar] [CrossRef]
  17. Wei, S.; Wang, X.; Zhang, R.; Hu, H.; Shen, Y.; Liu, J. Preparation and performance of spherical FeF2.5·0.5H2O nanoparticles wrapped by MWCNTs as cathode material of lithium ion batteries. RSC Adv. 2016, 6, 97759–97769. [Google Scholar] [CrossRef]
  18. Li, H.; Zhai, T.; He, P.; Wang, Y.; Hosono, E.; Zhou, H. Single-crystal H2V3O8 nanowires: A competitive anode with large capacity for aqueous lithium-ion batteries. J. Mater. Chem. 2011, 21, 1780–1787. [Google Scholar] [CrossRef]
  19. Wang, D.; Wei, Q.; Sheng, J.; Hu, P.; Yan, M.; Sun, R.; Xu, X.; An, Q.; Mai, L. Flexible additive free H2V3O8 nanowire membrane as cathode for sodium ion batteries. Phys. Chem. Chem. Phys. 2016, 18, 12074–12079. [Google Scholar] [CrossRef]
  20. An, Q.; Zhang, P.; Wei, Q.; He, L.; Xiong, F.; Sheng, J.; Wang, Q.; Mai, L. Top-down fabrication of three-dimensional porous V2O5 hierarchical microplates with tunable porosity for improved lithium battery performance. J. Mater. Chem. A 2014, 2, 3297–3302. [Google Scholar] [CrossRef]
  21. Li, W.D.; Xu, C.Y.; Pan, X.L.; Huang, Y.D.; Zhen, L. High capacity and enhanced structural reversibility of β-LixV2O5 nanorods as the lithium battery cathode. J. Mater. Chem. A 2013, 1, 5361. [Google Scholar] [CrossRef]
  22. Ahirrao, D.J.; Mohanapriya, K.; Jha, N. V2O5 nanowires-graphene composite as an outstanding electrode material for high electrochemical performance and long-cycle-life supercapacitor. Mater. Res. Bull. 2018, 108, 73–82. [Google Scholar] [CrossRef]
  23. Laorodphan, N.; Pooddee, P.; Kidkhunthod, P.; Kunthadee, P.; Tapala, W.; Puntharod, R. Boron and pentavalent vanadium local environments in binary vanadium borate glasses. J. Non-Cryst. Solids 2016, 453, 118–124. [Google Scholar] [CrossRef]
  24. Hassan, N.; Riaz, J.; Qureshi, M.T.; Razaq, A.; Rahim, M.; Toufiq, A.M.; Shakoor, A. Vanadium oxide (V2O3) for energy storage applications through hydrothermal route. J. Mater. Sci. Mater. Electron. 2018, 29, 16021–16026. [Google Scholar] [CrossRef]
  25. Roppolo, M.; Jacobs, C.B.; Upreti, S.; Chernova, N.A.; Whittingham, M.S. Synthesis and characterization of layered and scrolled amine-templated vanadium oxides. J. Mater. Sci. 2008, 43, 4742–4748. [Google Scholar] [CrossRef]
  26. Bordelet, G.; Yamaguchi, I.; Manabe, T.; Tsuchiya, T. Low temperature vanadium oxide thin film sintering by thermal and excimer-laser-assisted Metal-Organic Deposition (MOD). Ceram. Int. 2018, 44, S26–S29. [Google Scholar] [CrossRef]
  27. An, Q.; Sheng, J.; Xu, X.; Wei, Q.; Zhu, Y.; Han, C.; Niu, C.; Mai, L. Ultralong H2V3O8 nanowire bundles as a promising cathode for lithium batteries. New J. Chem. 2014, 38, 2075–2080. [Google Scholar] [CrossRef]
  28. Liu, P.; Bian, K.; Zhu, K.; Xu, Y.; Gao, Y.; Luo, H.; Lu, L.; Wang, J.; Liu, J.; Tai, G. Ultrathin Nanoribbons of in Situ Carbon-Coated V3O7·H2O for High-Energy and Long-Life Li-Ion Batteries: Synthesis, Electrochemical Performance, and Charge-Discharge Behavior. ACS Appl. Mater Interfaces 2017, 9, 17002–17012. [Google Scholar] [CrossRef]
  29. Mai, L.; Wei, Q.; An, Q.; Tian, X.; Zhao, Y.; Xu, X.; Xu, L.; Chang, L.; Zhang, Q. Nanoscroll buffered hybrid nanostructural VO2 (B) cathodes for high-rate and long-life lithium storage. Adv. Mater. 2013, 25, 2969–2973. [Google Scholar] [CrossRef]
  30. Shao, L.; Wu, K.; Lin, X.; Shui, M.; Ma, R.; Wang, D.; Long, N.; Ren, Y.; Shu, J. Sol–gel preparation of V2O5 sheets and their lithium storage behaviors studied by electrochemical and in-situ X-ray diffraction techniques. Ceram. Int. 2014, 40, 6115–6125. [Google Scholar] [CrossRef]
  31. Sakunthala, A.; Reddy, M.V.; Selvasekarapandian, S.; Chowdari, B.V.R.; Selvin, P.C. Energy storage studies of bare and doped vanadium pentoxide, (V1.95M0.05)O5, M = Nb, Ta, for lithium ion batteries. Energy Environ. Sci. 2011, 4, 1712. [Google Scholar] [CrossRef]
  32. Xu, Y.; Dunwell, M.; Fei, L.; Fu, E.; Lin, Q.; Patterson, B.; Yuan, B.; Deng, S.; Andersen, P.; Luo, H.; et al. Two-dimensional V2O5 sheet network as electrode for lithium-ion batteries. ACS Appl. Mater. Interfaces 2014, 6, 20408–20413. [Google Scholar] [CrossRef]
  33. Tang, K.; Yu, X.; Sun, J.; Li, H.; Huang, X. Kinetic analysis on LiFePO4 thin films by CV, GITT, and EIS. Electrochim. Acta 2011, 56, 4869–4875. [Google Scholar] [CrossRef]
  34. Jiang, H.; Wei, Z.; Cai, X.; Lai, L.; Ma, J.; Huang, W. A cathode for Li-ion batteries made of vanadium oxide on vertically aligned carbon nanotube arrays/graphene foam. Chem. Eng. J. 2019, 359, 1668–1676. [Google Scholar] [CrossRef]
  35. Liu, D.; Liu, Y.; Garcia, B.B.; Zhang, Q.; Pan, A.; Jeong, Y.H.; Cao, G. V2O5 xerogel electrodes with much enhanced lithium-ion intercalation properties with N2 annealing. J. Mater. Chem. 2009, 19, 8789. [Google Scholar] [CrossRef]
  36. Song, H.; Luo, M.; Wang, A. High Rate and Stable Li-Ion Insertion in Oxygen-Deficient LiV3O8 Nanosheets as a Cathode Material for Lithium-Ion Battery. ACS Appl. Mater. Interfaces 2017, 9, 2875–2882. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram for preparation of V3O7·H2O NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs as well as their corresponding molecular crystal structures.
Figure 1. Schematic diagram for preparation of V3O7·H2O NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs as well as their corresponding molecular crystal structures.
Nanomaterials 09 00624 g001
Figure 2. XRD patterns of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d).
Figure 2. XRD patterns of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d).
Nanomaterials 09 00624 g002
Figure 3. SEM images of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d).
Figure 3. SEM images of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d).
Nanomaterials 09 00624 g003
Figure 4. TEM and HRTEM images of V3O7·H2O NBs (a,b), VO2 (B) NBs/MWCNTs (c,d) and V2O5 NBs (e,f); the insets in (b) and (f) show SAED patterns.
Figure 4. TEM and HRTEM images of V3O7·H2O NBs (a,b), VO2 (B) NBs/MWCNTs (c,d) and V2O5 NBs (e,f); the insets in (b) and (f) show SAED patterns.
Nanomaterials 09 00624 g004
Figure 5. FTIR spectra of V3O7·H2O NBs, V3O7·H2O/VO2 (B)* NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs between 400–4000 cm−1.
Figure 5. FTIR spectra of V3O7·H2O NBs, V3O7·H2O/VO2 (B)* NBs, VO2 (B) NBs/MWCNTs and V2O5 NBs between 400–4000 cm−1.
Nanomaterials 09 00624 g005
Figure 6. CV curves of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) with a scanning speed of 0.2 mV s−1.
Figure 6. CV curves of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) with a scanning speed of 0.2 mV s−1.
Nanomaterials 09 00624 g006
Figure 7. Galvanostatic charge/discharge profiles of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) under different cycles at the current density of 100 mA g−1.
Figure 7. Galvanostatic charge/discharge profiles of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) under different cycles at the current density of 100 mA g−1.
Nanomaterials 09 00624 g007
Figure 8. Cycling performances and coulombic efficiencies of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) at 100 mA g−1; rate capability (e) and typical galvanostatic charge/discharge (GCD) profiles (f) of V2O5 NBs under varying current densities.
Figure 8. Cycling performances and coulombic efficiencies of V3O7·H2O NBs (a), V3O7·H2O/VO2 (B)* NBs (b), VO2 (B) NBs/MWCNTs (c) and V2O5 NBs (d) at 100 mA g−1; rate capability (e) and typical galvanostatic charge/discharge (GCD) profiles (f) of V2O5 NBs under varying current densities.
Nanomaterials 09 00624 g008
Figure 9. Nyquist plots (a) and dependences (b) of real resistance (Z′) on reciprocal of square root of angular frequency (ω) for all samples; the inset in (a) illustrating the equivalent circuit.
Figure 9. Nyquist plots (a) and dependences (b) of real resistance (Z′) on reciprocal of square root of angular frequency (ω) for all samples; the inset in (a) illustrating the equivalent circuit.
Nanomaterials 09 00624 g009
Table 1. Fitted values of Re, Rct and Aw for all prepared samples.
Table 1. Fitted values of Re, Rct and Aw for all prepared samples.
SamplesRe (Ω)Rct (Ω)Aw (Ω·s−1/2)
V3O7·H2O NBs5.1210.569.0
V3O7·H2O/VO2(B)* NBs4.6276.933.1
VO2(B) NBs/MWCNTs5.1199.842.9
V2O5 NBs5.7230.134.4

Share and Cite

MDPI and ACS Style

Jiang, Y.; Zhou, X.; Chen, X.; Wen, J.; Guan, L.; Shi, M.; Ren, Y.; Liu, Z. Controlled Hydrothermal Growth and Li+ Storage Performance of 1D VOx Nanobelts with Variable Vanadium Valence. Nanomaterials 2019, 9, 624. https://doi.org/10.3390/nano9040624

AMA Style

Jiang Y, Zhou X, Chen X, Wen J, Guan L, Shi M, Ren Y, Liu Z. Controlled Hydrothermal Growth and Li+ Storage Performance of 1D VOx Nanobelts with Variable Vanadium Valence. Nanomaterials. 2019; 9(4):624. https://doi.org/10.3390/nano9040624

Chicago/Turabian Style

Jiang, Yuhan, Xiaowei Zhou, Xu Chen, Jia Wen, Linlin Guan, Mingxia Shi, Yang Ren, and Zhu Liu. 2019. "Controlled Hydrothermal Growth and Li+ Storage Performance of 1D VOx Nanobelts with Variable Vanadium Valence" Nanomaterials 9, no. 4: 624. https://doi.org/10.3390/nano9040624

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop