High-performance lithium storage of Ti3+-doped anatase TiO2@C composite spheres

Xiaoying Chen, Li Liu*, Lingguang Yi, Guoxiong Guo, Min Li, Jianjun Xie, Yan Ouyang and Xianyou Wang
Key Laboratory of Environmentally Friendly Chemistry and Applications of Ministry of Education, School of Chemistry, Xiangtan University, Hunan, Xiangtan 411105, China. E-mail: liulili1203@126.com; Fax: +86-731-58292477; Tel: +86-731-58292206

Received 3rd September 2016 , Accepted 8th October 2016

First published on 10th October 2016


Abstract

Ti3+-Doped anatase TiO2@C composite spheres (TD-TiO2@C) are synthesized for the first time by the solvothermal method, followed by calcination using as-prepared anatase/TiO2-B hybrid TiO2 spheres and glycerol solution in alcohol as reactants. TD-TiO2@C spheres exhibit excellent electrochemical performance as anode materials for lithium ion batteries (LIBs). The electrodes not only exhibit a superior capacity of 244.80 mA h g−1 at 1C (0.168 A g−1) after 100 cycles, but also show an eminent rate capability. TD-TiO2@C spheres retain discharge capacities of 171.78, 143.45, 119.17, 105.82 mA h g−1 after 500 cycles at 5C, 10C, 20C, 30C, holding a capacity retention as high as 91.0%, 84.2%, 82.1%, 82.0% when compared with discharge capacities at the 10th cycle, respectively. The improved electrochemical properties are mainly due to the synergistic effects of carbon coating and Ti3+ doping, which can enhance significantly the inherent electronic conductivity of TiO2. Therefore, TD-TiO2@C can be an attractive candidate for anode materials in LIBs, also with great promise.


1. Introduction

Lithium-ion batteries with superior electrochemical properties (e.g., large energy density, good cycle performance and eco-friendliness) compared to conventional batteries, have attracted great attention and have been considered to have great value in theoretical studies and commercial applications.1,2 Although LIBs are widely used for cell phones, notebook computers, electronic electric vehicles and even hybrid electric vehicles, their anode materials’ electrochemical properties and safety need to be improved further.2,3 So, LIBs have continuously attracted lots of research in the design of satisfactory electrodes to meet the requirements of high energy/power density, long lifespan and safety.

TiO2 with different polymorphs (rutile, anatase, brookite, etc.) has been studied consumingly in recent years for its outstanding properties like minimal toxicity, low cost, nonpolluting nature, high discharge potential and most importantly, inherent chemical stability, which is derived from TiO2’s negligible volume expansion during the Li-ion insertion/extraction processes.4–7 Nevertheless, the widespread practical application of TiO2 as anodes in LIBs is seriously impeded by its inferior rate performance resulting from both extremely poor electronic conductivity (10−13 S cm−1) from the huge band gap (3.0 eV for anatase and 3.2 eV for rutile) and sluggish lithium ion diffusion (10−10 to 10−17 cm2 s−1).8–11 Many researches have been devoted to overcome these intrinsic drawbacks.

In recent years, more efforts have been devoted to enhance significantly the inherent electronic conductivity through metal or non-metal doping, such as B,12 N,13 Zn14 and Sn15 doping, which can offer more open channels in a particular direction for Li+ diffusion owing to the slight modification in the TiO2 lattice. Recently, it has been found that introducing Ti3+ or oxygen vacancies into TiO2, usually displaying a black or dark color, can markedly narrow the band gap to about 1.54 eV and thus enhance the inherent electronic conductivity significantly, which has drawn much interest, especially in photocatalysis.16–18 Since S. Myung et al. reported the ultra high cycling rates of black anatase titania via a solution evaporation process, the application of Ti3+-doped TiO2 for lithium storage has been studied very recently.19 It was revealed that the presence of Ti3+ narrowed the inherent high band gap energy to a semiconductor level, which was 1.8 eV, resulting in a very high electrical conductivity of 8 × 10−2 S cm−1.19 However, in order to produce Ti3+ or oxygen deficiency into bulk TiO2, harsh conditions, tedious experimental steps, and hazardous operation are usually required, which limits the practical utilization. In most reports, heat-treating TiO2 for several hours in a reducing atmosphere, such as H2, Ar/H2 or CO, was usually used to prepare Ti3+ doped TiO2.17,18,20–23 Although this method is effective, the need for expensive reducing atmospheres limits the large scale production of black TiO2. Other novel solution processes were also successfully carried out to prepare black Ti3+ doped TiO2. For example, S. Myung et al.,19 and Y. Ren et al.24 reported a solution evaporation process or a solvothermal process to prepare black anatase TiO2 using an amount of HF solution and TiCl4 solution as reactants. However, the requirements of corrosive HF and noxious TiCl4 are unsafe.

As we know, interconnecting electrode materials with electronic conductive additives (such as carbon) is a highly efficient way to enhance the electron transport of a material. Combining with CNTs,25 carbon hollow spheres,26 porous carbons,27 etc. to form TiO2/C composites is an effective method to improve the electrochemical properties and has been extensively investigated. Nevertheless, Ti3+ doping combined with carbon coating to improve the electrochemical properties of TiO2 hasn’t yet been reported.

In this work, Ti3+-doped anatase TiO2@C composite spheres (TD-TiO2@C) have been firstly synthesized by the solvothermal method followed by calcination using as-prepared anatase/TiO2-B hybrid TiO2 spheres and glycerol solution in alcohol as reactants. Comparing with the conventional doping method, this special method not only introduces Ti3+ into the bulk TiO2, but also produces a uniform carbon network. It decreases the band gap and enhances the inherent electronic conductivity of anatase TiO2 significantly. Here, the TD-TiO2@C spheres show excellent electrochemical performance as anode materials for LIBs.

2. Experimental

2.1. Synthesis of anatase–TiO2-B spheres

The anatase–TiO2-B spheres were prepared via a solvothermal process. This method is similar to our group’s and Yan’s work.28 The final products were white anatase/TiO2-B hybrid TiO2 (named as W-TiO2).

2.2. Synthesis of TD-TiO2@C spheres

1 g anatase–TiO2-B precursor was added into a mixture of 15 mL glycerol and 45 mL alcohol with stirring for 2 h. Afterwards, the mixture was transferred into a Teflon-lined stainless steel autoclave (80 mL), which was subsequently heated at 180 °C for 24 h. The precipitate was rinsed with distilled water and alcohol, then dried in air at 60 °C, and a creamy-white powder was obtained. The creamy-white precursor was annealed in a tubular furnace at 470 °C for 4 h in Ar to obtain black Ti3+-doped anatase TiO2@C composite spheres (TD-TiO2@C).

2.3. Structure and morphology characterization

X-Ray photoelectron spectroscopy (XPS ESCALAB 250 Xi Thermo Fisher Scientific, America) was performed to examine the composition of the surface materials. The structures of the as-synthesized samples were characterized by X-ray diffraction (XRD). XRD data were obtained by using a Rigaku D/MAX-2500 powder diffractometer with graphite monochromatic Cu Kα radiation (λ = 0.15418 nm). It operated at a scan rate of 5° min−1 in the 2θ range of 5–90°. Scanning electron microscope (SEM) images of the samples were collected using a JEOL JSM-6610 scanning electron microscope. Besides, high-resolution transmission electron microscopy (HRTEM) and selected-area electron diffraction were carried out using a JEOL JEM-2100F transmission electron microscope at an acceleration voltage of 200 kV. Raman spectra were conducted on a Horiba–Jobin–Yvon–Labor Raman HR-800 spectrometer with an argon ion laser of 523 nm. The specific surface areas (SBET) of the powders were evaluated following the multipoint Brunauer–Emmett–Teller (BET) equation and Barrett–Joyner–Halenda (BJH) method. The electronic conductivity of the powders was evaluated with an RTS-8 four point probe tester.

2.4. Electrochemical characterization

The W-TiO2 and TD-TiO2@C working electrodes for lithium cells were fabricated from the as-synthesized samples, carbon black, and polyvinylidene fluoride (PVDF) binder with a weight ratio of 80[thin space (1/6-em)]:[thin space (1/6-em)]10[thin space (1/6-em)]:[thin space (1/6-em)]10 in N-methyl pyrrolidinone, which were then pasted onto copper foil followed by drying under vacuum at 110 °C for 12 h. The average mass loading of the active material was about 1.6 mg cm−2. The testing cells were assembled with the working electrode thus fabricated, metallic lithium anode, Celgard 2300 film separator and LiPF6, (at a concentration of 1 M) in ethylene carbonate (EC)/dimethyl carbonate (DMC) electrolyte (v/v = 1[thin space (1/6-em)]:[thin space (1/6-em)]1). The testing cells were assembled in an argon-filled glove box, where the water and oxygen concentrations were kept less than 1 ppm.

All the cells were allowed to age for one night before testing. The charge–discharge cycle tests of cells (168 mA g−1, assumed to be at a 1C rate) were run at different current densities between 1.4 and 2.4 V. Cyclic voltammetry (CV) tests and EIS experiments were performed on a Zahner Zennium electrochemical workstation. CV tests were carried out at various cycles with a scan rate of 0.1 mV s−1 with the potential interval 1.4–2.4 V (vs. Li+/Li). The ac perturbation signal was ±5 mV and the frequency range was from 10 mHz to 100 kHz. All the tests were performed at room temperature.

3. Results and discussion

3.1. Characterization of TD-TiO2@C

The crystal structures of W-TiO2 and TD-TiO2@C were analyzed through XRD as shown in Fig. 1. The visual photographs of these samples are shown in the inset, where the TD-TiO2@C showed a dark black color. Most of diffraction peaks of W-TiO2 can be indexed well, based on an anatase phase corresponding to JCPDS no. 21-1272. Besides, a minor peak at 29.34° is obviously observed, which can be identified as the monoclinic TiO2(B) phase (JCPDS no. 35-0088). So, W-TiO2 can be indicated as anatase/TiO2-B hybrid TiO2. The hydrothermal treatment of anatase TiO2 in NaOH solution would facilitate its phase transformation to TiO2-B.28–30 Based on the nitrogen adsorption–desorption isotherms (Fig. S1), the anatase/TiO2-B hybrid TiO2 (W-TiO2) shows a much larger BET surface area than anatase TiO2 prepared without NaOH solution treatment, which were 130.7321 m2 g−1 and 94.7824 m2 g−1, respectively. Besides, it has also been proved that anatase/TiO2-B hybrid TiO2 shows a much better chemical reactivity than pure anatase TiO2,30 which maybe is beneficial for the formation of Ti3+ during the following processes. After undergoing solvothermal treatment with glycerol solution in alcohol, the white anatase/TiO2-B hybrid TiO2 transformed to a creamy-white powder, which can be identified as titanium glycerolate (TiGly) according its XRD pattern shown in Fig. S2. All the diffraction peaks of TD-TiO2@C can be indexed well based on anatase TiO2 (JCPDS no. 21-1272). The peak of TiO2-B disappeared after solvothermal treatment with glycerol and the following calcination, which may be due to the instability of TiO2-B at high temperature. TiO2-B transfers to anatase TiO2 close to 500 °C.31 It is noteworthy that the (101) diffraction peak of DT-TiO2@C shifts to lower angles and the width of (101) plane diffraction peak becomes broader as compared with W-TiO2.
image file: c6ra22105j-f1.tif
Fig. 1 XRD patterns of W-TiO2 and TD-TiO2@C (the inset is a visual photographs of the two samples).

Fig. 2 displays the X-ray photoelectron spectra (XPS) of W-TiO2 and TD-TiO2@C. The survey scan of XPS (Fig. 2a) shows a typical TiO2 spectrum and again confirms the absence of impurities. The Ti 2p core-level XPS spectra were measured in order to obtain detailed information on the chemical state of the titanium ion in the W-TiO2 and TD-TiO2@C. As shown in Fig. 2b, there are two strong peaks at 458.6 and 464.4 eV in the spectra of W-TiO2, corresponding to the typical 2p3/2 and 2p1/2 core levels of Ti4+, respectively. In contrast, as displayed in Fig. 2c, two major peaks at 458 and 463.5 eV were observed in the spectra of TD-TiO2@C, assigned to the 2p3/2 and 2p1/2 core levels of Ti4+, whereas the two shoulder peaks at 456.4 and 461.8 eV are attributed to the 2p3/2 and 2p1/2 core levels of Ti3+, respectively, revealing the existence of Ti3+ on TiO2, which is in good accordance with previous reports on Ti3+ doped TiO2.32 It is obvious that the typical Ti4+ 2p3/2 and Ti4+ 2p1/2 peaks of TD-TiO2@C shift to low binding energy compared with W-TiO2, demonstrating the existence of Ti3+ on TiO2 again.32,33 Since the radius of Ti4+ (0.606 Å) is almost the same as Ti3+ (0.670 Å), the Ti3+ ions can enter into the crystal structure of the anatase titanium dioxide and locate at the interstices or occupy some of the lattice sites of TiO2.34,35 For anatase TiO2, Ti3+ doping would induce the XRD diffraction peaks’ intensities to decrease and the width of (101) plane diffraction peak would become broader, which is in good agreement with the XRD pattern of TD-TiO2@C shown in Fig. 1.35


image file: c6ra22105j-f2.tif
Fig. 2 (a) The XPS survey scans of W-TiO2 and TD-TiO2@C. (b) Ti 2p XPS spectrum of W-TiO2. (c) Ti 2p XPS spectrum of TD-TiO2@C. (d) Raman spectra of W-TiO2 and TD-TiO2@C (the inset is the Raman spectra of W-TiO2 and TD-TiO2@C between a smaller range of Raman shift).

In addition, Raman spectra (Fig. 2d) also confirmed the presence of oxygen vacancies due to the existence of Ti3+. It is well known that the pure anatase TiO2 displays 6 typical peaks around 144, 197, 639, 399, 519 and 513 cm−1, corresponding to the E(1)g, E(2)g, E(3)g, B(1)1g, B(2)2g and A1g modes.36,37 As shown in Fig. 2d, the W-TiO2 and the fresh TD-TiO2@C exhibited a principal peak around 145 and 153 cm−1, respectively. According to the Raman spectra, TD-TiO2@C has a positive shift of the principal peak, which is characteristic of oxygen vacancies.38 These results clearly demonstrated that the oxygen vacancies were produced due to the replacement of Ti4+ with Ti3+ in the TiO2 lattice. This is in a good agreement with XPS spectra.

Besides, although no characteristic peak of carbon can be observed in the XRD pattern of TD-TiO2@C (Fig. 1), its Raman spectrum proves clearly the existence of carbon. The two peaks at around 1355 cm−1 and 1597 cm−1 (the inset of Fig. 2d), indicates that the carbon is partially graphitized with some defects and disorders. The G-band signifies the in-plane vibration of sp2 carbon atoms, while the D-band originates from the carbon defects.39,40 The calculated peak intensity ratio (ID/IG) is a useful index for comparing the degree of crystallinity of various carbon materials, i.e., a smaller value of the ID/IG ratio reflects a higher degree of ordering in the carbon material, which is related to the higher electronic conductivity.41 As shown in Fig. 2d, a small value of the ID/IG ratio of TD-TiO2@C is close to 0.81, suggesting that in the TD-TiO2@C sample has ordered graphitic carbon, and the amount of ordered graphitic carbon in the TD-TiO2@C sample is more than that of the disordered carbon. These results demonstrate that the TD-TiO2@C sample will have good conductivity. Certainly, the G-band and D-band are nonexistent in the Raman spectrum of W-TiO2. The conductivities of TD-TiO2@C and W-TiO2@C determined using the four point probe tester at 5 MPa (the sheet thickness is 1.5 mm, radius is 10 mm), are 8.918 × 10−6 S cm−1 and 3.234 × 10−7 S cm−1, respectively. It is obvious that TD-TiO2@C shows much higher electronic conductivity than of W-TiO2@C because of the tight carbon coating. Besides, it is interesting to note that W-TiO2@C shows a relatively higher electronic conductivity than some related reports in previous work,8–11 which can be attributed to the special bicrystalline structure.

As shown in Fig. 3a and b, the SEM image of W-TiO2 reveals the distribution of inhomogeneous and irregular spheres with a relatively smooth surface and diameters ranging from 200 to 500 nm. After glycerol treatment, the TD-TiO2@C (Fig. 3c and d) get smaller and quite regular, with wrinkled skin. The surface of the particles is changed from smooth to rough, which is clearly visible and shows the adoption of random orientations.


image file: c6ra22105j-f3.tif
Fig. 3 SEM images of W-TiO2 (a and b) and TD-TiO2@C (c and d).

In order to further investigate the morphology and structure of the final products, TEM and HRTEM images of W-TiO2 spheres and TD-TiO2@C spheres are shown in Fig. 4. In the TEM image, it can be found that both samples are composed of tiny nanoparticles. This special structure can increase the diffusion path and provide highly efficient solid-state diffusion of Li+. Compared with W-TiO2, with diameters ranging from 15 to 20 nm, the TD-TiO2@C spheres composed with nanosized particles show better regularity and smaller size with diameters ranging from 10 to 12 nm. This is in good agreement with the results of nitrogen adsorption–desorption isotherms. From Fig. S1, the specific surface area of TD-TiO2@C is 182.6725 m2 g−1, which is much larger than that of W-TiO2 (130.7321 m2 g−1). HRTEM analysis is employed to determine the crystal facets. Fig. 4d and h clearly show the lattice fringes with d-spacing values of 0.34–0.35 nm corresponding to the (101) planes of TiO2, while an amorphous carbon layer is observed over the TD-TiO2@C surface, exactly as attested by the Raman spectra. The homogeneously dispersed carbon not only sheathes the single TiO2 particles but also combines all individual TiO2 particles in a stable union.


image file: c6ra22105j-f4.tif
Fig. 4 TEM images of W-TiO2 (a–c) and TD-TiO2@C (e–g). HRTEM images of W-TiO2 (d) and TD-TiO2@C (h).

Based on the above analysis, the synthetic mechanism of TD-TiO2@C is shown in Fig. 5. Here, the solvothermal process converts the TiO2 and glycerol to titanium glycerolate (TiGly). In the process of calcination, glycerol was employed to reduce Ti4+ into Ti3+ and to avoid the oxidation of Ti3+,42,43 and was regarded as the source of the carbon. The formation of the carbon is due to the high temperature carbonization of the oligomer.


image file: c6ra22105j-f5.tif
Fig. 5 Schematic illustration for the formation of the TD-TiO2@C spheres’ microstructure.

3.2. Electrochemical analysis of TD-TiO2@C

Fig. 6 shows the charge/discharge performance of W-TiO2 and TD-TiO2@C. Fig. 6a shows the cycle performance and charge/discharge profiles (insets) of W-TiO2. W-TiO2 shows initial discharge capacities of 175.37 and 127.71 mA h g−1 at 1C and 0.5C, respectively. As shown in Fig. 6b, the cycling performances of TD-TiO2@C at the rates of 0.5C and 1C are depicted. TD-TiO2@C shows initial discharge capacities of 371.9, 362.2 mA h g−1 at 0.5C and 1C, respectively. It is easily observed that after modification by carbon and Ti3+, TD-TiO2@C exhibits a much higher capacity, which is 2–3 times higher than that of W-TiO2. TD-TiO2@C shows a relative large irreversible capacity loss on the first cycle. The coulombic efficiency of TD-TiO2@C on the first cycle is 78.2% at 0.5C, which is lower than that of W-TiO2. This may be due to the following reasons. First, TD-TiO2@C has a larger specific surface area than W-TiO2, which contributes to the improvement of the capacity of electrode, but could be detrimental for the coulombic efficiency. The formation of heterojunction interfaces involving electronic conductive additives can induce carrier trapping centers and hamper the reversible lithiation/de-lithiation process.44,45 Second, Fransson et al.46 reported that the carbon shows a tremendous uptake of Li+, resulting that the lithium ions intercalated in the material cannot deintercalate promptly. So, the amorphous carbon layer of TD-TiO2@C could also induce the uptake phenomenon, therefore the coulombic efficiency is low in the first cycle.
image file: c6ra22105j-f6.tif
Fig. 6 (a) Charge (hollow) and discharge (solid) capacity versus cycle number and charge/discharge profiles (insets) for (a) W-TiO2, and (b) TD-TiO2@C at current densities of 0.5C and 1C (1C = 168 mA g−1) in the range of 1.4–2.4 V in Li half-cells. (c) The rate capabilities of W-TiO2 and TD-TiO2@C. (d) Charge–discharge capacities of TD-TiO2@C over 500 cycles at 5C, 10C, 20C, 30C in Li half-cells.

To compare with the rate performance of TD-TiO2@C and W-TiO2, these two samples were cycled at various rates (from 0.5C to 30C) after cycling for 100 cycles at 0.5C, as shown in Fig. 6c. W-TiO2 delivers a discharge capacity of 138.8 mA h g−1 (0.5C), 129.9 mA h g−1 (1C), 110.4 mA h g−1 (2C), 96.9 mA h g−1 (5C), 76.3 mA h g−1 (10C), 57.6 mA h g−1 (20C) and 40.7 mA h g−1 (30C). TD-TiO2@C shows 265.1 mA h g−1 (0.5C), 219.3 mA h g−1 (1C), 193.9 mA h g−1 (2C), 178.8 mA h g−1 (5C), 147.7 mA h g−1 (10C), 129.1 mA h g−1 (20C) and 109.5 mA h g−1 (30C), which is far better than the electrochemical performance of W-TiO2. The results reveal that Ti3+ self-doping and carbon modification can enhance the electrochemical performance of TiO2 remarkably, which can be attributed to the enhanced electrical conductivity. The introduction of Ti3+ species enables fast electron transfer.19 According to the report of Wang et al.,47 the presence of abundant defects makes the c axis of anatase shrink significantly, whereas the a and b axes elongated slightly, this widens the valence band top edge. So, the defects of TiO2 switch TiO2 conductivity from n-type to p-type with a higher charge mobility. Carbon combines all individual TiO2 nano-particles in stable spherical unions, which form a good network of electrically conductive paths among the TiO2 particles. So, the TiO2 active material can get electrons from all directions and be fully utilized for lithium ion insertion and extraction reactions, which can improve lithium storage properties with enhanced rate capability. Besides, a larger capacity could be obtained at high rates if the particle diameter of the electrode material can be made smaller. It has been reported that during the charge/discharge process, the effective diffusion length for Li+ ion is 24–240 nm at a rate of 1C, and 3.2–32 nm at 60C.48 The average diameter of the well-dispersed TD-TiO2@C was 200–500 nm, over 240 nm, but the the whole TD-TiO2@C consisted of ultrafine nanoparticles of about 10–12 nm; this special structure can increase the diffusion path and provide highly efficient solid-state diffusion of Li+, and could also shorten remarkably the Li+ diffusion length, especially at high rates. Besides, the spaces between the nanoparticles were beneficial for Li+ diffusion.

To evaluate the cyclability of TD-TiO2@C at high rate, the cells were tested at a high rate. TD-TiO2@C electrode was sequentially discharged–charged at 5C to 30C over 500 cycles (as shown in Fig. 6d). It is found that the coulombic efficiency increases upon cycling; it reaches nearly 100% after 10 cycles. TD-TiO2@C shows discharge capacities of 188.54, 170.2, 145.19, and 121.9 mA h g−1 at the 10th cycle at 5C, 10C, 20C, and 30C, respectively. After 500 cycles, the retained discharge capacities are 171.78, 143.45, 119.17, and 100.02 mA h g−1, holding a capacity retention as high as 91%, 84.2%, 82.1%, and 82% compared with the capacities at the 10th cycle, respectively. These results dramatically show that TD-TiO2@C can be widely employed in research and development for the improvement of LIBs because of its excellent electrochemical performance.

Fig. 7 shows cyclic voltammetry (CV) profiles of W-TiO2 and TD-TiO2@C spheres between 1.4 and 2.4 V at a scanning rate of 0.1 mV s−1. As shown in Fig. 7a, W-TiO2 shows a pair of oxidation/reduction peaks (A) at 2.07/1.64 V, corresponding to the extraction and insertion of lithium ions in anatase TiO2; the other reduction and oxidation potentials of the remaining two pairs of weak peaks, denoted as S1 and S2, are 1.50 V/1.54 V and 1.50 V/1.60 V vs. Li+/Li, respectively. These other two pairs (S1, S2) reflect the characteristic pseudo-capacitive behavior of lithium storage in TiO2-B.49,50 The pseudo-capacitive behavior derives from the unique sites and energetics of lithium absorption and diffusion in the TiO2-B structure, which have been studied by theoretical methods.51,52 As shown in Fig. 7b, the voltammogram of TD-TiO2@C, there is a pair of clearly defined reduction/oxidation peaks, observed at 1.70 V (reduction peak) and 2.00 V (oxidation peak), which can be attributed to the discharge and charge processes of the Li+ intercalation/deintercalation of the anatase framework. As shown in the Figure, after ten cycles, the consistency of the TD-TiO2@C voltammetry curves was better than W-TiO2; that is say, the stability of TD-TiO2@C is better than W-TiO2. According to the second cycles, the value of the reduction and oxidation potentials’ difference for TD-TiO2@C is 0.32 V, which is much smaller than that of W-TiO2 (0.4 V). This obviously demonstrates that TD-TiO2@C has better reversibility.


image file: c6ra22105j-f7.tif
Fig. 7 Cyclic voltammograms of (a) W-TiO2 and (b) TD-TiO2@C at a scan rate of 0.1 mV s−1 between 1.4 and 2.4 V (vs. Li+/Li).

In order to compare and gain insight into the further changes of the electrode microstructure during cycling in the TD-TiO2@C and W-TiO2 battery, electrochemical impedance spectroscopy (EIS) was conducted; these were tested at different cycles at the rate of 5C and charged to the same voltage of 2.4 V. The shapes of the Nyquist plots for each cycle are similar. It is evident from Fig. 8 that there is a sloping line in the low frequency region, also one semicircle was observed in the high frequency region for all of the Nyquist plots. Nyquist plots are fitted using the simplified equivalent circuit model (Fig. 8c). Table 1 shows the fitted impedance data. The fitting patterns show that fitting data are in good agreement with experimental data, as shown in Fig. 8a and b. The equivalent circuit model includes Rs at the small intercept of the Z′ axis, showing the internal resistance of the electrode and electrolyte in the battery, a constant phase element (CPE) associated with the surface resistance of the electrode, while the semicircle in the high frequency region corresponding to Rct represents the lithium charge transfer resistance on the interface of the electrode electrolyte. The sloping line portion is assigned to the Warburg impedance (W1), which is attributed to the diffusion of lithium ion into the bulk of the electrode materials. It is explicit from Table 1 that the Rs and Rct of TD-TiO2@C are much smaller than those of W-TiO2, indicating that the TD-TiO2@C has higher conductivity and faster lithium ion diffusion than W-TiO2. The Rct of W-TiO2 is 56.88 Ω after one cycle, and this value increases to 122.40 Ω after 500 cycles. However, the Rct of the TD-TiO2@C is 59.90 Ω after one cycle, and this value only increases to 90.30 Ω after 500 cycles. As everyone knows, a lower growth rate of impedance during cycling represents lower polarization, which indicates good cycling behavior. The result is attributed to the higher electronic conductivity resulting from the presence of Ti3+ and the unique structures of TD-TiO2@C, enabling much faster charge transfer and transport at the electrode/electrolyte interface. These results are in good agreement with the excellent electrochemical performance of TD-TiO2@C nanocomposites.


image file: c6ra22105j-f8.tif
Fig. 8 Three-dimensional Nyquist plots measured for (a) W-TiO2 and (b) TD-TiO2@C after cycling for different cycles at 5C in Li half-cells. (c) The equivalent circuit model.
Table 1 Rs and Rct values of W-TiO2 and TD-TiO2@C nanocomposites after different cycles
Rs Rct
1st 100th 200th 300th 400th 500th 1st 100th 200th 300th 400th 500th
TD-TiO2@C 1.15 1.52 1.95 2.16 2.17 2.31 59.9 60.88 66.91 70.07 84.25 90.3
W-TiO2 2.66 2.79 2.83 2.85 3.1 3.33 56.88 58.05 69.07 102.7 108.5 122.4


4. Conclusion

Ti3+-Doped anatase TiO2@C composite spheres (TD-TiO2@C) presenting a dark color were successfully prepared by the solvothermal method followed by calcination. TD-TiO2@C demonstrates remarkable electrochemical performance as the anode material for LIBs. The electrode exhibits a capacitance of 244.8 mA h g−1 at 1C (0.168 A g−1) after 100 cycles, which is far superior to the white TiO2, and shows a good rate capability. TD-TiO2@C retains the discharge capacities of 171.78, 143.45, 119.17 and 105.82 mA h g−1 after 500 cycles at 5C, 10C, 20C and 30C, holding a capacity retention of as high as 91.0%, 84.2%, 82.1% and 82.0% when compared with discharge capacities at the 10th cycle, respectively. The considerably enhanced durable high-rate performance is attributed to the improved electronic conductivity resulting from the introduction of Ti3+ and carbon into the TD-TiO2@C composites. It is anticipated that the Ti3+-doped anatase TiO2@C composite spheres can be employed as promising candidates for anode materials for fast and durable LIBs.

Acknowledgements

This study was financially supported by the National Natural Science Foundation of China (Grant No. 51672234), the Research Foundation of Education Bureau of Hunan Province (Grant No. 15B229), the Research Foundation for Hunan Youth Outstanding People from Hunan Provincial Science and Technology Department (2015RS4030), the Hunan Provincial Natural Science Foundation of China (Grant No. 14JJ6017), and the Program for Innovative Research Cultivation Team in University of the Ministry of Education of China (1337304).

References

  1. F. Wu, X. Li, Z. Wang and H. Guo, Nanoscale, 2013, 5, 6936 RSC.
  2. Q. Zhou, L. Liu, J. Tan, Z. Yan, Z. Huang and X. Wang, J. Power Sources, 2015, 283, 243 CrossRef CAS.
  3. J. Chen, Y. Tan, C. Li, Y. L. Cheah, D. Luan, S. Madhavi, F. Y. C. Boey, L. A. Archer and X. Lou, J. Am. Chem. Soc., 2010, 132, 6124 CrossRef CAS PubMed.
  4. Z. Yan, L. Liu, H. Shu, X. Yang, H. Wang, J. Tan, Q. Zhou, Z. Huang and X. Wang, J. Power Sources, 2015, 274, 8 CrossRef CAS.
  5. L. Liu, M. Zhou, L. Yi, H. Guo, J. Tan, H. Shu, X. Yang, Z. Yang and X. Wang, J. Mater. Chem., 2012, 22, 17539 RSC.
  6. M. C. Tsai, J. Chang, H. S. Sheu, H. T. Chiu and C. Y. Lee, Chem. Mater., 2009, 21, 499 CrossRef CAS.
  7. Q. Zhou, L. Liu, H. Guo, R. Xu, J. Tan, Z. Yan, Z. Huang, H. Shu, X. Yang and X. Wang, Electrochim. Acta, 2015, 151, 502 CrossRef CAS.
  8. G. Li, Z. Lian, X. Li, Y. Xu, W. Wang, D. Zhang, F. Tian and H. Li, J. Mater. Chem. A, 2015, 3, 3748 CAS.
  9. I. Moriguchi, R. Hidaka, H. Yamada, T. Kudo, H. Murakami and N. Nakashima, Adv. Mater., 2006, 18, 69 CrossRef CAS.
  10. I. Abayev, A. Zaban, F. Fabregat-Santiag and J. Bisquert, Phys. Status Solidi A, 2003, 196, R4 CrossRef CAS.
  11. S. Moitzheim, C. S. Nimisha, S. Deng, D. J. Cott, C. Detavernier and P. M. Vereecken, Nanotechnology, 2014, 25, 504008 CrossRef CAS PubMed.
  12. H. Tian, F. Xin, X. Tan and W. Han, J. Mater. Chem. A, 2014, 2, 10599 CAS.
  13. Y. Yang, X. Ji, M. Jing, H. Hou, Y. Zhu, L. Fang, X. Yang, Q. Chen and C. E. Banks, J. Mater. Chem. A, 2015, 3, 5648 CAS.
  14. Z. Ali, S. N. Cha, J. I. Sohn, I. Shakir, C. Yan, J. M. Kim and D. J. Kang, J. Mater. Chem., 2012, 22, 17625 RSC.
  15. M. Lübke, I. Johnson, N. M. Makwana, D. Brett, P. Shearing, Z. Liu and J. A. Darr, J. Power Sources, 2015, 294, 94 CrossRef.
  16. X. Chen, L. Liu and F. Huang, Chem. Soc. Rev., 2015, 44, 1861 RSC.
  17. X. Chen, L. Liu, Y. Y. Peter and S. S. Mao, Science, 2011, 331, 746 CrossRef CAS PubMed.
  18. A. Naldoni, M. Allieta, S. Santangelo, M. Marelli, F. Fabbri, S. Cappelli, C. L. Bianchi, R. Psaro and V. Dal Santo, J. Am. Chem. Soc., 2012, 134, 7600 CrossRef CAS PubMed.
  19. S. T. Myung, M. Kikuchi, C. S. Yoon, H. Yashiro, S. Kim, Y. Sun and B. Scrosati, Energy Environ. Sci., 2013, 6, 2609 CAS.
  20. Y. Yan, B. Hao, D. Wang, G. Chen, E. Markweg, A. Albrecht and P. Schaaf, J. Mater. Chem. A, 2013, 1, 14507 CAS.
  21. J. Eom, S. Lim, S. Lee, W. Ryu and H. Kwon, J. Mater. Chem. A, 2015, 3, 11183 CAS.
  22. F. Zuo, L. Wang, T. Wu, Z. Zhang, D. Borchardt and P. Feng, J. Am. Chem. Soc., 2010, 132, 11856 CrossRef CAS PubMed.
  23. D. Liu, Y. Zhang, P. Xiao, B. Garcia, Q. Zhang, X. Zhou, Y. Jeong and G. Cao, Electrochim. Acta, 2009, 54, 6816 CrossRef CAS.
  24. Y. Ren, J. Li and J. Yu, Electrochim. Acta, 2014, 138, 41 CrossRef CAS.
  25. R. Qing, L. Liu, H. Kim and W. Sigmund, Electrochim. Acta, 2015, 180, 295 CrossRef CAS.
  26. F. Yang, Z. Zhang, Y. Han, K. Du, Y. Lai and J. Li, Electrochim. Acta, 2015, 178, 871 CrossRef CAS.
  27. J. Lee, Y. Chen, Y. Zhu and B. Vogt, ACS Appl. Mater. Interfaces, 2014, 6, 21011 CAS.
  28. Z. Yan, L. Liu, J. Tan, Q. Zhou, Z. Huang, D. Xia, H. Shu, X. Yang and X. Wang, J. Power Sources, 2014, 269, 37 CrossRef CAS.
  29. T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino and K. Niihara, Langmuir, 1998, 14, 3160 CrossRef CAS.
  30. H. Izawa, S. Kikkawa and M. Koizumi, J. Solid State Chem., 1985, 60, 264 CrossRef CAS.
  31. X. Zhu, X. Yang, C. Lv, S. Guo, J. Li, Z. Zheng, H. Zhu and D. Yang, ACS Appl. Mater. Interfaces, 2016, 8, 18815 CAS.
  32. J. Chen, W. Song, H. Hou, Y. Zhang., M. Jing, X. Jia and X. Ji, Adv. Funct. Mater., 2015, 25, 6793 CrossRef CAS.
  33. Y. He, J. Peng, W. Chu, Y. Lia and D. Tong, J. Mater. Chem. A, 2014, 2, 1721 CAS.
  34. C. Wang, C. Bottcher, D. Bahnemann and J. Dohrmann, J. Mater. Chem., 2003, 13, 2322 RSC.
  35. M. Zhou, J. Yu and B. Cheng, J. Hazard. Mater., 2006, 137, 1838 CrossRef CAS PubMed.
  36. X. Pan, M. Yang, X. Fu, N. Zhang and Y. Xu, Nanoscale, 2013, 5, 3601 RSC.
  37. T. Ohsaka, F. Izumi and Y. Fujiki, J. Raman Spectrosc., 1978, 7, 321 CrossRef.
  38. J. Parker and R. Siegel, Appl. Phys. Lett., 1990, 57, 943 CrossRef CAS.
  39. H. Huang, Y. Xia, X. Tao, J. Du, J. Fang, Y. Gan and W. Zhang, J. Mater. Chem., 2012, 22, 10452 RSC.
  40. X. Tao, X. Zhang, L. Zhang, J. Cheng, F. Liu, J. Luo, Z. Luo and H. J. Geise, Carbon, 2006, 44, 1425 CrossRef CAS.
  41. Y. Xia, W. Zhang, Z. Xiao, H. Huang, H. Zeng, X. Chen, F. Chen, Y. Gan and X. Tao, J. Mater. Chem., 2012, 22, 9209 RSC.
  42. H. Tüysüz, Y. Liu, C. Weidenthaler and F. Schüth, J. Am. Chem. Soc., 2008, 130, 14108 CrossRef PubMed.
  43. J. Callahan, K. Hool, J. Reynolds and K. Cook, Anal. Chem., 1988, 60, 714 CrossRef CAS.
  44. J. Zheng, Front. Phys. China, 2008, 3, 269 CrossRef.
  45. Y. Kim, J. Lee, S. Cho, Y. Kwon, I. In, N. You, E. Reichmanis, H. Ko and K. Lee, ACS Nano, 2014, 8, 6701 CrossRef CAS PubMed.
  46. L. Fransson, T. Eriksson, K. Edstrom, T. Gustafsson and J. Thomas, J. Power Sources, 2001, 101, 1–9 CrossRef CAS.
  47. S. Wang, L. Pan, J. Song, W. Mi, J. Zou, L. Wang and X. Zhang, J. Am. Chem. Soc., 2015, 137, 2975 CrossRef CAS PubMed.
  48. T. Xia, W. Zhang, Z. Wang, Y. Zhang, X. Song, J. Murowchick, V. Battaglia, G. Liu and X. Chen, Nano Energy, 2014, 6, 109 CrossRef CAS.
  49. M. Zukalová, M. Kalbáč, L. Kavan, I. Exnar and M. Graetzel, Chem. Mater., 2005, 17, 1248 CrossRef.
  50. L. Kavan, M. Kalbac, M. Zukalova, I. Exnar, V. Lorenzen, R. Nesper and M. Gratzel, Chem. Mater., 2004, 16, 477 CrossRef CAS.
  51. A. Armstrong, C. Arrouvel, V. Gentili, S. Parker, M. Islam and P. Bruce, Chem. Mater., 2010, 22, 6426 CrossRef CAS.
  52. A. Dalton, A. Belak and A. Van der Ven, Chem. Mater., 2012, 24, 1568 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra22105j

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.