Enhanced electrochemical performances of MoO2 nanoparticles composited with carbon nanotubes for lithium-ion battery anodes

Song Qiua, Guixia Lua, Jiurong Liu*a, Hailong Lyua, Chenxi Hua, Bo Lib, Xingru Yanc, Jiang Guoc and Zhanhu Guo*c
aKey Laboratory for Liquid–Solid Structural Evolution and Processing of Materials, Ministry of Education and School of Materials Science and Engineering, Shandong University, Jinan, Shandong 250061, People's Republic of China. E-mail: jrliu@sdu.edu.cn
bState Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
cIntegrated Composites Laboratory (ICL), Department of Chemical & Biomolecular Engineering, University of Tennessee, Knoxville, Tennessee 37996, United States. E-mail: zguo10@utk.edu

Received 25th August 2015 , Accepted 9th October 2015

First published on 9th October 2015


Abstract

The nanocomposites of carbon nanotubes (CNTs) with homogenously anchored molybdenum dioxide (MoO2) nanoparticles of 20–50 nm have been successfully synthesized by a hydrothermal method. Glucose has dual functions, i.e., reducing agent and surfactant to prohibit the anisotropic growth, resulting in the direct deposition of MoO2 nanoparticles on the CNT surfaces. As the anode of lithium-ion batteries (LIBs), the as-prepared MoO2/CNT nanocomposites deliver a higher reversible capacity of 640 mA h g−1 at a current density of 100 mA g−1 after 100 cycles, compared to only 246 and 259 mA h g−1 for MoO3 nanobelts/CNT composites and MoO2 nanoparticles, respectively. The superior electrochemical performances are attributed to the nanocomposite structure of MoO2 nanoparticles anchored on CNTs, which have efficiently enhanced the electrical conductivity and lithium ion diffusion, and maintained the integrity of electrode during the charge/discharge processes.


1. Introduction

Recently, many efforts have been devoted to exploring novel electrode materials for the application of ever-growing high-power density lithium ion batteries (LIBs). As potential anode materials, transition metal oxides including Fe3O4,1 Fe2O3,2 Co3O4,3 NiO,4 Mn3O4,5 and MoO3,6 have shown higher capacities than that of the commercially employed graphite anode (specific capacity of 372 mA h g−1). Among them, molybdenum oxides such as MoO3(ref. 6–8) and MoO2 (ref. 9–11) have attracted extensive attention due to their high theoretical capacities (1117 mA h g−1 for MoO3, and 838 mA h g−1 for MoO2), high electrochemical activity, suitable lithium reaction voltage, and affordable cost. However, most of the transition metal oxides serving as anode materials for LIBs usually suffer from large volume expansion/shrinkage and pulverization during the lithiation/delithiation processes, which will lead to the destruction of the anode structure and thereby rapidly degrade the electrochemical performances.9,12

To maintain the integrity of anode structure, the employment of nano-structured metal oxides is an efficient approach to relieve the stress–strain generated during the lithiation/delithiation processes. It also shortens the pathway for Li-ion diffusion and electron transport, and provides more contact area between anode material and electrolyte, leading to a higher energy density.6,13,14 For example, Lee et al.7 have reported that MoO3 nanocrystallites prepared by a chemical vapor deposition method exhibited a reversible capacity of 630 mA h g−1 after 150 cycles at the current rate of 1/2C, while MoO3 particles with the size of 5 μm displayed a rapid capacity loss with cycling. Core–shell MoO2 hierarchical microcapsules have been synthesized by a template-free solvothermal method.9 The MoO2 microcapsule anode delivered a reversible capacity of 623.8 mA h g−1 after 50 cycles at a current rate of 1C, while the anode made of commercial MoO2 powders showed a rapid capacity fading and only maintained the initial capacity of 35%, indicating that the nanostructured molybdenum oxides exhibited superior lithium storage capability.

The combination of metal oxides with various carbon materials, such as amorphous carbon,15–19 graphene,20,21 nanofibers,22,23 and nanotubes,24–26 is another strategy to improve the electric conductivity of anode and to accommodate the strain of volume change as a buffer, both are beneficial to enhance the electrochemical performances of LIBs. Among carbon materials, carbon nanotubes (CNTs) have attracted wide attentions as additives in constructing hybrid nanocomposites due to their superior electrical conductivity, high surface-to-volume ratio, ultrathin walls, and structural flexibility.27–32 For example, the composites of CNTs with MoO3 nanobelts showed a reversible capacity of 421 mA h g−1 at the current density of 200 mA g−1 after 100 cycles, and delivered 293 and 202 mA h g−1 at current densities of 2 and 4 A g−1, respectively.31 Inspired by the above researches, constructing the nanocomposites of MoO2 nanoparticles anchored on CNTs will be more beneficial to its LIBs application since CNTs can be intertwined to form unique conductive networks, which provide continuous pathways for electron transport, and MoO2 nanoparticles anchored on CNTs will supply high capacity. In addition, the CNTs with MoO2 nanoparticles deposited on surface can serve as the buffer to prevent the aggregation and pulverization of MoO2 nanoparticles during the lithiation/delithiation processes, and thereby maintain the integrity of electrode structure.

In this work, MoO2/CNTs nanocomposites have been synthesized by a facile hydrothermal approach with the assistance of glucose, which was employed as the reducing agent and the structure directing surfactant. The electrochemical measurements demonstrate that the as-synthesized MoO2/CNTs nanocomposite exhibits superior cycling and rate performances than the MoO2 nanoparticles and the composite of CNTs with MoO3 nanobelts.

2. Experimental

2.1. Materials

Sodium molybdate dihydrate (Na2MoO4·2H2O), D-glucose (C6H12O6), hydrochloric acid solution (HCl, 37 wt%), concentrated nitric acid (HNO3, 68 wt%) and CNTs were purchased from Sinopharm Chemical Reagent Co., Ltd. Carbon black, Li foil and Celgard 2300 were provided by Hefei Kejing Material Technology Co., Ltd, China. Polyvinylidene fluoride (PVDF), LiPF6 (dissolved in ethylene carbonate, dimethyl carbonate, and ethylene methyl carbonate with a volume ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1) were purchased from Shenzhen Biyuan Technology Co., Ltd, China. All the chemicals were analytical grade and were used as-received without further purification.

2.2. Preparation of MoO2/CNTs nanocomposite, MoO3 nanobelts/CNTs composite and MoO2 nanoparticles

Firstly, the purchased CNTs were treated by nitric acid at 140 °C for 6 h in a Teflon-lined stainless steel autoclave. After that, the acid-treated CNTs were centrifuged and washed with deionized water and ethanol several times until the pH value to about 7, and then dried under vacuum at 60 °C for 24 h. The MoO2/CNTs nanocomposites were synthesized as follows. Typically, 0.02 g CNTs were dispersed in 10 mL deionized water under ultrasonic treatment for 1 h, and then 0.2 mmol Na2MoO4·2H2O and 0.05 g glucose were added into the suspension, followed by the dropwise addition of 1 mL HCl (3 M) under magnetic stirring. Thirty minutes later, the mixture was transferred into a 30 mL Teflon-lined stainless steel autoclave, and subsequently sealed and heated at 180 °C for 12 h in an oven. The black precipitate was collected by centrifugation and washed with deionized water and ethanol several times, and then dried in vacuum overnight. The preparation procedures of the composites of CNTs with MoO3 nanobelts and MoO2 nanoparticles were the same as those of the MoO2/CNTs nanocomposites only without adding glucose or CNTs in the hydrothermal process, respectively.

2.3. Characterizations

The structure of products was determined by X-ray powder diffraction (XRD) on a Rigaku D/Max-RC X-ray diffractometer with Ni filtered Cu Kα radiation (λ = 0.1542 nm, V = 40 kV, I = 40 mA) in the range of 10–80° at a scanning rate of 4° min−1. A JSM-6700F field emission scanning electron microscope (FE-SEM) at an accelerating voltage of 20 kV and an electric current of 1.0 × 10−10 A; and A JEOL JEM-2100 high-resolution transmission electron microscopy (HR-TEM) with an accelerating voltage of 200 kV were utilized to examine the morphology and microstructure. X-ray photoelectron spectra (XPS) were recorded on a Kratos Analytical spectrometer, using Al Kα ( = 1486.6 eV) radiation as the excitation source, under a condition of anode voltage of 12 kV and an emission current of 10 mA. Thermal gravimetric analysis (TGA) measurement was carried out in an SDT Q600 (TA Instruments Ltd., New Castle, DE) thermal-microbalance apparatus at a heating rate of 10 °C min−1 in an air atmosphere from ambient temperature to 700 °C to evaluate carbon content in the products.

2.4. Electrochemical measurements

To prepare the working electrode, the homogenous slurry was formed by dispersing the as-prepared active material, carbon black, and polyvinylidene fluoride (PVDF) with a weight ratio of 8[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 in N-methyl-2-pyrrolidinone (NMP). The slurry was coated on a copper foil substrate, followed by drying in a vacuum oven at 120 °C for 12 h. The CR2025-type cells were assembled in a glovebox filled with argon atmosphere using Li foil as counter electrode, Celgard 2300 as separator, and 1 M LiPF6 (dissolved in ethylene carbonate, dimethyl carbonate, and ethylene methyl carbonate with a volume ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1) as electrolyte. As the active material, the typical mass loading of MoO2/CNTs, MoO3 nanobelts/CNTs and MoO2 on the working electrodes was 1.5–2.0 mg cm−2. The performance of the cells was evaluated galvanostatically in the voltage range from 0.02 to 3 V at various current densities on a LAND CT2001A battery test system. Cyclic voltammogram (CV) was obtained by an IviumStat electrochemistry workstation at a scan rate of 0.3 mV s−1 and the potential vs. Li/Li+ ranging from 0.01 to 3 V. The electrochemical impedance spectrum (EIS) was obtained on the same instrument with AC signal amplitude of 10 mV in the frequency range from 100 kHz to 0.01 Hz. The data were adopted to draw Nyquist plots using real part Z′ as X axis, and imaginary part Z′′ as Y axis.

3. Results and discussion

3.1. Characterizations of MoO2/CNTs nanocomposites, MoO3 nanobelts/CNTs composites and MoO2 nanoparticles

From the XRD pattern of MoO2/CNTs nanocomposite shown in Fig. 1a, all diffraction peaks can be assigned to monoclinic MoO2 (JCPDS 65-5787), indicating the reduction of Mo6+ to Mo4+ by glucose during the hydrothermal process. For the composites of CNTs and MoO3 nanobelts (Fig. 1b) fabricated without the addition of glucose, all the recorded peaks could be ascribed to those of orthorhombic MoO3 (JCPDS 05-0508), and no diffraction peaks of MoO2 or other molybdenum oxides were observed. The intensity of (020), (040) and (060) peaks is stronger than the standard diffraction data of MoO3 (JCPDS 05-0508), indicating the anisotropic growth of MoO3 nanobelts. Fig. 1c displays the XRD pattern of MoO2 nanoparticles prepared without the addition of CNTs. All diffraction peaks match well with the monoclinic MoO2 (JCPDS 65-5787), also confirming that Mo6+ has been reduced to Mo4+ by glucose in the hydrothermal reaction.
image file: c5ra17147d-f1.tif
Fig. 1 XRD patterns of (a) MoO2/CNTs nanocomposite, (b) MoO3 nanobelts/CNTs composite, and (c) MoO2 nanoparticles.

The FE-SEM image (Fig. 2a) of MoO2/CNTs nanocomposites shows that the surface of CNTs is smooth and no aggregated MoO2 particles are observed except for CNTs. However, from the high-magnification image (Fig. 2b), a lot of dispersed MoO2 nanoparticles are observed to be attached on the surface of CNTs. For the sample prepared without the addition of glucose, FE-SEM images (Fig. 2c and d) reveal that CNTs are mixed with MoO3 nanobelts with the width of ca. 500 nm and the length of ca. 10 μm to form composites. Thus, glucose is believed to play an important role as the surface modifier in controlling the morphology and size of molybdenum oxides. Glucose, a kind of surfactant, can physically absorb on the CNTs surface to link the subunits of MoO2 and prohibit the anisotropic growth, resulting in the direct deposition of MoO2 nanoparticles on CNTs surface. Fig. 2e and f show the SEM images of the sample fabricated without the addition of CNTs. The product is composed of nanoparticles with the size of 20–50 nm and a lot of nanoparticles are agglomerated together to form large clumps. No nanobelts are observed in the sample, further confirming that glucose prohibits the anisotropic growth of molybdenum oxides.


image file: c5ra17147d-f2.tif
Fig. 2 FE-SEM images of (a and b) MoO2/CNTs nanocomposites, (c and d) MoO3 nanobelts/CNTs composites, and (e and f) MoO2 nanoparticles.

Fig. 3 presents the HR-TEM images of MoO2/CNTs nanocomposite. A lot of MoO2 nanoparticles with the size of 20–50 nm are observed anchored on the surface of CNTs, Fig. 3a and b. The higher magnification image (Fig. 3c) reveals a fringe with a spacing of 0.34 nm, corresponding to the (002) crystalline plane of CNTs. The measured lattice distance of nanoparticle is around 0.34 nm, corresponding to the (011) plane of MoO2.


image file: c5ra17147d-f3.tif
Fig. 3 HR-TEM images of MoO2/CNTs nanocomposites at different magnification (a–c).

To evaluate the carbon content, TGA tests have been carried out in the fabricated MoO2/CNTs nanocomposites, MoO3 nanobelts/CNTs composites, and MoO2 nanoparticles. In the TGA and DSC curves of MoO2/CNTs nanocomposites (Fig. 4a), one exothermic peak accompanying with the weight loss of about 2.5% is observed in the temperature range of 300–400 °C, corresponding to the oxidation process of glucose residue.33 Another exothermic peak at 400–600 °C accompanying with an intensive 37.9% weight loss is attributed to the oxidation of CNTs to form volatile species such as CO and CO2.28 For the MoO3 nanobelts/CNTs composites (Fig. 4b), only one exothermic peak observed at 400–600 °C in the DSC curve accompanying with 38.5% weight loss corresponds to the oxidation of CNTs. In Fig. 4c, an exothermic peak, appeared in the DSC curve of MoO2 nanoparticles from 300 to 400 °C, indicates the oxidation of glucose residue. Meanwhile, a slight weight gain is observed in the TGA curve, suggesting the oxidation reaction from MoO2 to MoO3, which is also observed in previous works.34,35


image file: c5ra17147d-f4.tif
Fig. 4 TGA/DSC curves of (a) MoO2/CNTs nanocomposites, (b) MoO3 nanobelts/CNTs composites, and (c) MoO2 nanoparticles.

The XPS measurement was carried out to determine the chemical composition of MoO2/CNTs nanocomposites, Fig. 5. In the survey scan (Fig. 5a), the distinct peaks at 530.1, 415.5, 398.0, 284.6, and 233.2 eV are assigned to the O 1s, Mo 3p1/2, Mo 3p3/2, C 1s, and Mo 3d, respectively, indicating the presence of O, Mo and C elements in the nanocomposites. In the Mo 3d spectrum (Fig. 5b), two peaks at 233.3 and 230.2 eV correspond to Mo 3d3/2 and Mo 3d5/2, respectively, which are in agreement with the XPS results of MoO2 in the previous reports.36,37 The O 1s spectrum (Fig. 5c) is broad and deconvoluted into three peaks at 530.7, 532.3 and 533.6 eV, which are assigned to Mo–O–Mo, C–O–Mo and C[double bond, length as m-dash]O bonds, respectively.38 The C–O–Mo interfacial bonding introduced onto the CNTs surfaces has substantially improved the structural integrity and electrochemical cyclability of MoO2/CNTs nanocomposites. Similar phenomena have been demonstrated in other CNTs composites.39,40 The C 1s peak (Fig. 5d) can be deconvoluted into three peaks at 284.6, 285.7, and 286.6 eV, corresponding to the C–C, C–O and C[double bond, length as m-dash]O groups.41–43 On the base of above analysis, the proposed formation mechanism of the as-synthesized MoO2/CNTs nanocomposite is given in Fig. 6.


image file: c5ra17147d-f5.tif
Fig. 5 XPS spectra of MoO2/CNTs nanocomposites: (a) survey scan, (b) Mo 3d, (c) O 1s, and (d) C 1s.

image file: c5ra17147d-f6.tif
Fig. 6 Schematic diagrams illustrating the formation mechanism of MoO2/CNTs nanocomposites.

3.2. Electrochemical performances of MoO2/CNTs nanocomposite, MoO3 nanobelts/CNTs composite and MoO2 nanoparticles

To evaluate the applicability of MoO2/CNTs nanocomposites as LIBs anode material, the electrochemical performances were investigated in the voltage window of 0.01–3.0 V. For comparison, the MoO3 nanobelts/CNTs composites and MoO2 nanoparticles were also employed as anode materials. Fig. 7a–c display the cyclic voltammetry (CV) curves of the MoO2/CNTs, MoO3 nanobelts/CNTs, and MoO2 samples in the initial three cycles, respectively. For the MoO2/CNTs nanocomposites (Fig. 7a), there are six cathodic peaks appeared at ca. 0.06, 0.5, 1.1, 1.4, 2.2 and 2.7 V in the first discharge process. According to the CV result and previous works, two cathodic peaks at 1.4 V (shifted to 1.5 V from the second cycle) and 1.1 V (shifted to 1.2 V from the second cycle) are the characteristics for Li+ insertion reactions in MoO2 to form Lix1MoO2 and Lix2MoO2 (x2 > x1), respectively, accompanying with the phase transition.32,44–47 One weak peak at 0.5 V disappeared from the second discharge process is related to the irreversible reduction of electrolyte and the formation of solid electrolyte interphase (SEI) layer.48 The main cathodic peak at ca. 0.06 V (shifted to 0.01 V from the second cycle) can be ascribed to the Li conversion reaction, in which Mo metal and Li2O are produced.46,47 Compared with the CV curve of MoO3 nanobelts/CNTs (Fig. 7b), two weak peaks at 2.2 and 2.7 V (Fig. 7a) disappeared from the second discharge process originate from the Li+ insertion reaction of MoO3, suggesting that the slight oxidation of MoO2 nanoparticles still existed in the MoO2/CNTs nanocomposites due to the exposure to air although there is no diffraction peak of MoO3 observed in XRD pattern (Fig. 1a). In the subsequent charge process, a wide and split anodic peak in 1.4–1.8 V is attributed to the oxidation of Mo0 to Mo4+ and the decomposition of Li2O. After the second cycle, the CV curves tend to overlap very well, demonstrating that this electrode exhibits a gradually enhanced cycling stability for the insertion and extraction of lithium ions. Therefore, the reversible electrochemical reaction formula in lithium ion battery can be described as the lithium insertion/extraction in MoO2 along with the phase transition and subsequent conversion reaction:46,47
 
MoO2 (monoclinic) + x1Li+ + x1e ↔ Lix1MoO2 (orthorhombic) (1)
 
Lix1MoO2 (orthorhombic) + (x2x1)Li+ + (x2x1)e ↔ Lix2MoO2(monoclinic) (2)
 
Lix2MoO2 + (4 − x2)Li+ + (4 − x2)e ↔ Mo + 2Li2O (3)

image file: c5ra17147d-f7.tif
Fig. 7 Cyclic voltammogram curves of (a) MoO2/CNTs nanocomposites, (b) MoO3 nanobelts/CNTs composites, (c) MoO2 nanoparticles, and (d) Nyquist plots of the MoO2/CNTs, MoO3 nanobelts/CNTs and MoO2 samples. The insets in (b) and (d) are the highlighted CV curve of MoO3 nanobelts/CNTs composites and the equivalent circuit, respectively.

As to the MoO3 nanobelts/CNTs composites, there are four reduction peaks at ca. 0.07, 0.5, 2.1 and 2.6 V observed in the first cathodic process, and three oxidation peaks at ca. 0.07, 1.3 and 1.7 V appeared in the counterpart anodic process (Fig. 7b). The cathodic peaks at 0.5, 1.3 and 1.7 V disappeared from the second discharge process, suggesting the irreversible decomposition of the electrolyte, the formation of SEI layer, and the Li+ insertion in the Mo–O octahedron layers and Mo–O octahedron intralayers with the formation of LixMoO3.49–52 The intense reduction peak at 0.07 V (shifted to 0.01 V from the second cycle) corresponds to the electrochemical reduction reaction of LixMoO3 with Li to generate Mo0. From the second cycle, the CV spectra (inset of Fig. 7b) are similar to those of MoO2/CNTs (Fig. 7a) and MoO2 (Fig. 7c). A broad anodic peak in 1.2–1.8 V is attributed to the oxidation of Mo0 to Mo4+ and the decomposition of Li2O. The above experimental results suggest that, the reduced Mo0 is mainly re-oxidized to Mo4+ rather than Mo6+ in the first charge process. That is to say, from the second cycle, the mainly reversible electrochemical reaction in cell is the lithium insertion/extraction reactions in MoO2. Therefore, two cathodic peaks at ca. 1.5 and 1.2 V are the characteristics for Li+ insertion reactions in MoO2 to form LixMoO2 from the second cycle. For MoO2 sample, the CV profile (Fig. 7c) is similar to that of MoO2/CNTs. The two cathodic peaks at 1.3 and 1.0 V are attributed to the lithium insertion in MoO2 to form LixMoO2. The peak at 0.4 V disappeared from the second discharge process, corresponding to the irreversible reduction of electrolyte and the formation of SEI layer. The reduction peak at 0.04 V (shifted to 0.01 V from the second cycle) is ascribed to the reduction of lithiated LixMoO2 to metallic Mo. A weak and broad peak at ca. 2.2 V disappeared from the second discharge process indicates the presence of Mo6+ due to the slight oxidation of MoO2 nanoparticles exposed to air. In the following charge process, one broad and split anodic peak in 1.4–1.8 V is attributed to the oxidation of Mo0 to Mo4+ and the decomposition of Li2O.

To further investigate the electrochemical performances of MoO2/CNTs, MoO3 nanobelts/CNTs, and MoO2 samples, electrochemical impedance spectroscopy (EIS) measurements conducted on the cells at a current density of 100 mA g−1 after 100 cycling are shown in Fig. 7d. The Nyquist plot of each cell is comprised of arc in the high- and medium-frequency region, and an inclined line in the low frequency region.53,54 The diameter of the semicircle is in direct proportion to the impedance, which contains electrolyte resistance (Re), surface film resistance (Rsf) and charge transfer resistance (Rct).55–57 The inclined line is related to the lithium-ion diffusion inside the electrode materials corresponding to the Warburg impedance (Zw).58,59 The impedance spectra can be fitted based on a reasonable equivalent circuit (inset of Fig. 7d), in accordance with the Li-ion insertion/extraction mechanism in electrode.60,61 CPE in this model expresses the double layer capacitance. The (Re + Rsf + Rct) values for MoO2/CNTs, MoO3 nanobelts/CNTs, and MoO2 samples are ca. 39, 83 and 112 Ω, respectively. The measured results indicate that the addition of CNTs into molybdenum oxides enhanced the electrical conductivity and charge transfer. Comparing with MoO3 nanobelts/CNTs, the MoO2/CNTs nanocomposites exhibited lower impedance, which could be attributed to the nanostructure of MoO2 nanoparticles anchored on CNTs resulting in better electrical contact of CNTs network (Fig. 2). In the low frequency region, the slope of inclined line for MoO2/CNTs nanocomposites and MoO2 nanoparticles is larger than that of MoO3 nanobelts/CNTs composites, suggesting that the lithium ion diffusion ability of these two samples is superior to MoO3 nanobelts/CNTs composites. The reason could be ascribed to the size decrease of molybdenum oxides based on the SEM measurements (Fig. 2), which is beneficial to lithium ion diffusion. Therefore, the addition of CNTs in the MoO2/CNTs nanocomposites can enhance the electrical conductivity, and the nanosized MoO2 particles anchored on CNTs facilitate the lithium ion diffusion.

The galvanostatic discharge/charge curves of MoO2/CNTs, MoO3 nanobelts/CNTs and MoO2 in the potential range from 0.02 to 3.0 V vs. Li+/Li reference electrode at the current density of 100 mA g−1 are shown in Fig. 8. The MoO2/CNTs electrode exhibits initial discharge–charge capacities of 782 and 459 mA h g−1, respectively, with a coulombic efficiency (the ratio of charge capacity to discharge capacity) of 58.7% (Fig. 8a). The large initial irreversible loss is attributed to the irreversible reduction of MoO2 to Mo and other possible irreversible processes (electrolyte decomposition and the formation of SEI layer).47 The initial coulombic efficiency of MoO3 nanobelts/CNTs is 57.9% (Fig. 8b), with initial discharge–charge capacities of 694 and 402 mA h g−1, respectively. For the MoO2 electrode, the first discharge and charge capacities are 475 and 299 mA h g−1, respectively (Fig. 8c). The initial coulombic efficiency is 62.9%. The MoO2 electrode shows the highest initial coulombic efficiency. Furthermore, the curves for MoO2/CNTs and MoO2 electrode of the second and third discharge/charge cycles coincide better than those of MoO3 nanobelts/CNTs electrode. These results indicate the good reversibility of the electrochemical reaction of MoO2 and MoO2/CNTs.


image file: c5ra17147d-f8.tif
Fig. 8 Galvanostatic discharge/charge curves of the 1st, 2nd, and 3rd cycle for (a) MoO2/CNTs nanocomposites, (b) MoO3 nanobelts/CNTs composites and (c) MoO2 nanoparticles.

Fig. 9a shows the cycling performance of MoO2/CNTs nanocomposites at a constant current density of 100 mA g−1. The nanocomposite electrode delivers the reversible capacity of 640 mA h g−1 after 100 cycles. The reversible capacity of MoO2/CNTs electrode exhibits an increasing tendency with the cycling onward within the initial 50 cycles, even reaches 805 mA h g−1 after 50 cycles due to a gradual activation process of MoO2/CNTs electrode during the discharge/charge processes. Similar phenomena have been also found in other anode nanomaterials.5,30 As a comparison, the cycling performances of MoO3 nanobelts/CNTs and MoO2 nanoparticles at the current density of 100 mA g−1 are also shown in Fig. 9a. In the initial 20 cycles, MoO2 nanoparticles exhibit a similar capacity increase tendency as the MoO2/CNTs electrode, however, the capacity decreases successively with further cycling and is only 259 mA h g−1 after 100 cycles, suggesting that the addition of CNTs is beneficial to improve the cycling stability. For the MoO3 nanobelts/CNTs composites, the capacity shows intensive decrease to 246 mA h g−1 after 100 cycles probably due to the structural isolation between MoO3 nanobelts and CNTs (Fig. 2), and the pulverization of large-sized MoO3 nanobelts during the Li-ion insertion/extraction processes, which break down the integrity of electrode.


image file: c5ra17147d-f9.tif
Fig. 9 (a) Cycling performances of MoO2/CNTs, MoO3 nanobelts/CNTs, and MoO2 at the current density of 100 mA g−1, and (b) the rate capabilities of MoO2/CNTs and MoO3 nanobelts/CNTs nanocomposites.

The higher reversible capacity of MoO2/CNTs nanocomposites than those of MoO3 nanobelts/CNTs and MoO2 nanoparticles further demonstrates that the hybrid structure of MoO2 nanoparticles anchored on CNTs significantly improves the cycling performance and leads to higher retention capacity. The cycling performance of MoO2/CNTs nanocomposites is also much better than those of the reported MoO2/graphite composites (720 mA h g−1 after 30 cycles at the current density of 100 mA g−1)62 and MoO2/C hollow spheres (640 mA h g−1 after 80 cycles at the current density of 50 mA g−1)63 in long cycling performance and reversible capacity. Moreover, the coulombic efficiency quickly reaches to 90% in the second cycle, even remains more than 96% from fifth to 100th cycle. From the rate capability shown in Fig. 9b, the MoO2/CNTs nanocomposite electrode achieves reversible capacities of ca. 560, 572, 539 and 356 mA h g−1 at the current densities of 100, 200, 400 and 800 mA g−1, respectively. When the current density is recovered to 100 mA g−1 after the rate performance test, the discharge capacity reaches 692 mA h g−1 higher than 560 mA h g−1 acquired at the initial current density of 100 mA g−1, suggesting that the high current charge/discharge processes not only did little to break down the integrity of the electrodes, but also leaded to a gradual activation of the electrode materials, which have also been found in other anode nanomaterials.64–66 However, the MoO3 nanobelts/CNTs composites only deliver reversible capacities of ca. 305, 232, 179, and 129 mA h g−1 at the current densities of 100, 200, 400 and 800 mA g−1, respectively. The MoO3 nanobelts/CNTs composites show much lower rate capability at the same current density than MoO2/CNTs nanocomposites, mainly attributing to the structural isolation between MoO3 nanobelts and CNTs. On the base of the above results and discussion, the superior electrochemical performances of the MoO2/CNTs nanocomposites can be attributed to the following reasons. Firstly, the presence of intertwined CNTs improves the electrical conductivity of the nanocomposites and accommodates the volume change during the lithiation/delithiation processes. The nanostructure of MoO2 nanoparticles anchored on CNTs is beneficial to prevent the agglomeration of MoO2 nanoparticles and maintain the electrode integrity. In addition, the small size of MoO2 nanoparticles (ca. 20–50 nm) and the thin walls of CNTs facilitate the Li ions diffusion in the electrode and increase the contact surface between the active materials and electrolyte. Moreover, the good attachment between MoO2 nanoparticles and CNTs can effectively avoid the electrical isolation of MoO2/CNTs anode during the charge/discharge processes.

4. Conclusion

The nanocomposites of MoO2 nanoparticles anchored on carbon nanotubes (MoO2/CNTs) have been synthesized as anode material for Li-ion battery by a facile approach. Compared with the as-prepared MoO3 nanobelts/CNTs and MoO2 nanoparticles, the MoO2/CNTs nanocomposites exhibit superior cycling and rate performances. The CNTs network in the nanocomposites not only provides a continuous long-distance pathway for electron transfer, but also acts as a favorable buffer to the aggregation of the anchored MoO2 nanoparticles and the volume change, which maintains the integrity of electrode during the lithiation/delithiation processes, finally resulting in the enhanced electrochemical performances. The hybridization of MoO2 nanoparticles with CNTs is proved to be an efficient way to improve the electrochemical performances of LIBs anode.

Acknowledgements

This work was supported by the Fundamental Research Funds of Shandong University (2015JC016) and Natural Science Fund for Distinguished Young Scholars of Shandong (JQ201312). The authors also acknowledge the financial supports from the Shandong Provincial Natural Science Foundation, China (No. ZR2009FM035).

References

  1. D. Su, H. J. Ahn and G. Wang, J. Power Sources, 2013, 244, 742–746 CrossRef CAS PubMed.
  2. M. V. Reddy, T. Yu, C. H. Sow, Z. X. Shen, C. T. Lim, G. V. Subba Rao and B. V. R. Chowdari, Adv. Funct. Mater., 2007, 17, 2792–2799 CrossRef CAS PubMed.
  3. W. Y. Li, L. N. Xu and J. Chen, Adv. Funct. Mater., 2005, 15, 851–857 CrossRef CAS PubMed.
  4. D. Su, H. S. Kim, W. S. Kim and G. Wang, Chem.–Eur. J., 2012, 18, 8224–8229 CrossRef CAS PubMed.
  5. J. Gao, M. A. Lowe and H. D. Abruña, Chem. Mater., 2011, 23, 3223–3227 CrossRef CAS.
  6. L. A. Riley, S. H. Lee, L. Gedvilias and A. C. Dillon, J. Power Sources, 2010, 195, 588–592 CrossRef CAS PubMed.
  7. S. H. Lee, Y. H. Kim, R. Deshpande, P. A. Parilla, E. Whitney, D. T. Gillaspie, K. M. Jones, A. H. Mahan, S. Zhang and A. C. Dillon, Adv. Mater., 2008, 20, 3627–3632 CrossRef CAS PubMed.
  8. M. F. Hassan, Z. P. Guo, Z. Chen and H. K. Liu, J. Power Sources, 2010, 195, 2372–2376 CrossRef CAS PubMed.
  9. X. Zhao, M. Cao, B. Liu, Y. Tian and C. Hu, J. Mater. Chem., 2012, 22, 13334–13340 RSC.
  10. L. C. Yang, Q. S. Gao, Y. Tang, Y. P. Wu and R. Holze, J. Power Sources, 2008, 179, 357–360 CrossRef CAS PubMed.
  11. J. H. Ku, J. H. Ryu, S. H. Kim, O. H. Han and S. M. Oh, Adv. Funct. Mater., 2012, 22, 3658–3664 CrossRef CAS PubMed.
  12. K. Sakaushi, J. Thomas, S. Kaskel and J. Eckert, Chem. Mater., 2013, 25, 2557–2563 CrossRef CAS.
  13. B. Guo, X. Fang, B. Li, Y. Shi, C. Ouyang, Y. S. Hu, Z. Wang, G. D. Stucky and L. Chen, Chem. Mater., 2012, 24, 457–463 CrossRef CAS.
  14. Z. Yuan and L. Si, J. Mater. Chem. A, 2013, 1, 15247–15251 CAS.
  15. S. Qiu, X. Wang, G. Lu, J. Liu and C. He, Mater. Lett., 2014, 136, 289–291 CrossRef CAS PubMed.
  16. S. Guo, G. Lu, S. Qiu, J. Liu, X. Wang, C. He, H. Wei, X. Yan and Z. Guo, Nano Energy, 2014, 9, 41–49 CrossRef CAS PubMed.
  17. H. Qu, S. Wei and Z. Guo, J. Mater. Chem. A, 2013, 1, 11513–11528 CAS.
  18. C. Hu, S. Guo, G. Lu, Y. Fu, J. Liu, H. Wei, X. Yan, Y. Wang and Z. Guo, Electrochim. Acta, 2014, 148, 118–126 CrossRef CAS PubMed.
  19. H. Cao, X. Wang, H. Gu, J. Liu, L. Luan, W. Liu, Y. Wang and Z. Guo, RSC Adv., 2015, 5, 34566–34571 RSC.
  20. P. J. Lu, M. Lei and J. Liu, CrystEngComm, 2014, 16, 6745–6755 RSC.
  21. D. Qiu, G. Bu, B. Zhao, Z. Lin, L. Pu, L. Pan and Y. Shi, Mater. Lett., 2014, 119, 12–15 CrossRef CAS PubMed.
  22. W. Luo, X. Hu, Y. Sun and Y. Huang, Phys. Chem. Chem. Phys., 2011, 13, 16735–16740 RSC.
  23. H. Lv, S. Qiu, G. Lu, Y. Fu, X. Li, C. Hu and J. Liu, Electrochim. Acta, 2015, 151, 214–221 CrossRef CAS PubMed.
  24. C. Ban, Z. Wu, D. T. Gillaspie, L. Chen, Y. Yan, J. L. Blackburn and A. C. Dillon, Adv. Mater., 2010, 22, E145–E149 CrossRef CAS PubMed.
  25. X. Li, H. Gu, J. Liu, H. Wei, S. Qiu, Y. Fu, H. Lv, G. Lu, Y. Wang and Z. Guo, RSC Adv., 2015, 5, 7237–7244 RSC.
  26. S. Qiu, H. Gu, G. Lu, J. Liu, X. Li, Y. Fu, X. Yan, C. Hu and Z. Guo, RSC Adv., 2015, 5, 46509–46516 RSC.
  27. M. O. Guler, T. Cetinkaya, U. Tocoglu and H. Akbulut, Microelectron. Eng., 2014, 118, 54–60 CrossRef CAS PubMed.
  28. J. Wang, R. Ran, M. O. Tade and Z. Shao, J. Power Sources, 2014, 254, 18–28 CrossRef CAS PubMed.
  29. C. Xu, J. Sun and L. Gao, J. Power Sources, 2011, 196, 5138–5142 CrossRef CAS PubMed.
  30. Z. Wang, D. Luan, S. Madhavi, Y. Hu and X. W. D. Lou, Energy Environ. Sci., 2012, 5, 5252–5256 CAS.
  31. J. Ren, J. Yang, A. Abouimrane, D. Wang and K. Amine, J. Power Sources, 2011, 196, 8701–8705 CrossRef CAS PubMed.
  32. J. P. Jegal, H. K. Kim, J. S. Kim and K. B. Kim, J. Electroceram., 2013, 31, 218–223 CrossRef CAS.
  33. X. Wu, H. Niu, S. Fu, J. Song, C. Mao, S. Zhang, D. Zhang and C. Chen, J. Mater. Chem. A, 2014, 2, 6790–6795 CAS.
  34. Y. Sun, X. Hu, W. Luo and Y. Huang, ACS Nano, 2011, 5, 7100–7107 CrossRef CAS PubMed.
  35. K. H. Seng, G. D. Du, L. Li, Z. X. Chen, H. K. Liu and Z. P. Guo, J. Mater. Chem., 2012, 22, 16072–16077 RSC.
  36. J. G. Choi and L. T. Thompson, Appl. Surf. Sci., 1996, 93, 143–149 CrossRef CAS.
  37. X. Chen, Z. Zhang, X. Li, C. Shi and X. Li, Chem. Phys. Lett., 2006, 418, 105–108 CrossRef CAS PubMed.
  38. A. Bhaskar, M. Deepa and T. N. Rao, ACS Appl. Mater. Interfaces, 2013, 5, 2555–2566 CAS.
  39. C. Sun, K. Karki, Z. Jia, H. Liao, Y. Zhang, T. Li, Y. Qi, J. Cumings, G. W. Rubloff and Y. Wang, ACS Nano, 2013, 7, 2717–2724 CrossRef CAS PubMed.
  40. C. Sun, H. Zhu, M. Okada, K. Gaskell, Y. Inoue, L. Hu and Y. Wang, Nano Lett., 2015, 15, 703–708 CrossRef CAS PubMed.
  41. S. Kundu, Y. Wang, W. Xia and M. Muhler, J. Phys. Chem. C, 2008, 112, 16869–16878 CAS.
  42. R. H. Bradley, K. Cassity, R. Andrews, M. Meier, S. Osbeck, A. Andreu, C. Johnston and A. Crossley, Appl. Surf. Sci., 2012, 258, 4835–4843 CrossRef CAS PubMed.
  43. F. Sun, J. Gao, Y. Zhu, G. Chen, S. Wu and Y. Qin, Adsorption, 2013, 19, 959–966 CrossRef CAS.
  44. B. Liu, X. Zhao, Y. Tian, D. Zhao, C. Hu and M. Cao, Phys. Chem. Chem. Phys., 2013, 15, 8831–8837 RSC.
  45. L. Zhou, H. B. Wu, Z. Wang and X. W. D. Lou, ACS Appl. Mater. Interfaces, 2011, 3, 4853–4857 CAS.
  46. J. H. Ku, Y. S. Jung, K. T. Lee, C. H. Kim and S. M. Oh, J. Electrochem. Soc., 2009, 156, A688–A693 CrossRef CAS PubMed.
  47. Y. Liu, H. Zhang, P. Ouyang, W. Chen and Z. Li, Mater. Res. Bull., 2014, 50, 95–102 CrossRef CAS PubMed.
  48. Y. Sun, X. Hu, J. C. Yu, Q. Li, W. Luo, L. Yuan, W. Zhang and Y. Huang, Energy Environ. Sci., 2011, 4, 2870–2877 CAS.
  49. Y. Sun, J. Wang, B. Zhao, R. Cai, R. Ran and Z. Shao, J. Mater. Chem. A, 2013, 1, 4736–4746 CAS.
  50. W. Li, F. Cheng, Z. Tao and J. Chen, J. Phys. Chem. B, 2006, 110, 119–124 CrossRef CAS PubMed.
  51. T. Tsumura and M. Inagaki, Solid State Ionics, 1997, 104, 183–189 CrossRef CAS.
  52. J. Ni, G. Wang, J. Yang, D. Gao, J. Chen, L. Gao and Y. Li, J. Power Sources, 2014, 247, 90–94 CrossRef CAS PubMed.
  53. G. Lu, S. Qiu, J. Liu, X. Wang, C. He and Y. Bai, Electrochim. Acta, 2014, 117, 230–238 CrossRef CAS PubMed.
  54. J. Kang, Y. Ko, J. Park and D. Kim, Nanoscale Res. Lett., 2008, 3, 390–394 CrossRef CAS.
  55. Y. Huang, Z. Dong, D. Jia, Z. Guo and W. I. Cho, Electrochim. Acta, 2011, 56, 9233–9239 CrossRef CAS PubMed.
  56. Q. Zhang, Z. Shi, Y. Deng, J. Zheng, G. Liu and G. Chen, J. Power Sources, 2012, 197, 305–309 CrossRef CAS PubMed.
  57. Z. Wu, W. Ren, L. Xu, F. Li and H. Cheng, ACS Nano, 2011, 5, 5463–5471 CrossRef CAS PubMed.
  58. Y. Deng, Q. Zhang, Z. Shi, L. Han, F. Peng and G. Chen, Electrochim. Acta, 2012, 76, 495–503 CrossRef CAS PubMed.
  59. V. Freger, Electrochem. Commun., 2005, 7, 957–961 CrossRef CAS PubMed.
  60. W. Yue, S. Tao, J. Fu, Z. Gao and Y. Ren, Carbon, 2013, 65, 97–104 CrossRef CAS PubMed.
  61. M. V. Reddy, T. Yu, C. H. Sow, Z. X. Shen, C. T. Lim, G. V. Subba Rao and B. V. R. Chowdari, Adv. Funct. Mater., 2007, 17, 2792–2799 CrossRef CAS PubMed.
  62. Y. Xu, R. Yi, B. Yuan, X. Wu, M. Dunwell, Q. Lin, L. Fei, S. Deng, P. Andersen, D. Wang and H. Luo, J. Phys. Chem. Lett., 2012, 3, 309–314 CrossRef CAS PubMed.
  63. H. Gao, C. Liu, Y. Liu, Z. Liu and W. Dong, Mater. Chem. Phys., 2014, 147, 218–224 CrossRef CAS PubMed.
  64. G. Lu, S. Qiu, H. Lv, Y. Fu, J. Liu, X. Li and Y. Bai, Electrochim. Acta, 2014, 146, 249–256 CrossRef CAS PubMed.
  65. S. Grugeon, S. Laruelle, L. Dupont and J. M. Tarascon, Solid State Sci., 2003, 5, 895–904 CrossRef CAS.
  66. C. Gong, Y. Bai, Y. Qi, N. Lun and J. Feng, Electrochim. Acta, 2013, 90, 119–127 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.