A free-standing, flexible and bendable lithium-ion anode materials with improved performance

Xueqian Zhanga, Xiaoxiao Huang*a, Yingfei Zhanga, Long Xiab, Bo Zhongb, Xiaodong Zhangb, Nan Tiana, Tao Zhangb and Guangwu Wen*ab
aSchool of Materials Science and Engineering, Harbin Institute of Technology, Harbin 150001, China. E-mail: swliza@hit.edu.cn; g.wen@hit.edu.cn
bSchool of Materials Science and Engineering, Harbin Institute of Technology at Weihai, Weihai 264209, China

Received 31st July 2016 , Accepted 13th October 2016

First published on 31st October 2016


Abstract

Bendable, flexible and self-supported anode materials with excellent electrochemical properties have highly attractive for the high performance lithium-ion batteries (LIBs). We report a simple process to prepare a three dimension structure composed of carbon and SnO2 nanoparticles (CGN/SnO2). The non-woven cotton covered with conducting graphene is selected as a flexible and bendable skeleton. Ultrasmall SnO2 nanoparticles with a size of several nanometers uniformly anchor on both the graphene sheets and cotton fibers. The CGN/SnO2 composites exhibited a higher Li-battery performance than carbonized cotton/SnO2 (C/SnO2) and carbonized cotton covered by glucose/SnO2 (CG/SnO2). Especially, the CGN/SnO2 composites after bending 1000 times can maintain the same lithium storage property as before. Combining with this unique flexible, bendable structure, and the outstanding electrical and electrochemical performance, we demonstrate the great potential of CGN/SnO2 composites in high performance Li-ion battery field and the flexible, foldable and stretchable conductors application.


1. Introduction

An important trend of the advanced anode materials in rechargeable lithium-ions batteries (LIBs) for portable electronics and hybrid electric vehicles applications is mechanical flexibility, stability, simple and reproducible preparation process with high energy capacity and power density. Recent studies on the flexible anode materials have been reported.1–9 Among them, one useful method is to get mixture of the active materials, conductive carbon and binder on a flexible current collector, and then to prepare the anodes.1,10 However, this method is complicated and expensive, and some detrimental effect between electrolyte and inactive materials inevitably occur, thus influencing the stability of the electrode.

A major scientific challenge for the flexible anode materials with stable performance is developing the self-supported anode materials without polymer binders, carbon conductive and collectors. Some approaches have been exploited including graphene paper,11 graphene/CNT12 and graphene/SnO2 (ref. 13 and 14) composites. Liang et al. reported that they used a simple filtration method and following a thermal reduction to prepare the flexible free-standing graphene/SnO2 nanocomposites with the excellent electrochemical performance for electrode.13 Sun et al. vacuum-filtrated graphene oxide and CNT mixtures, and then annealed in Ar–H2 atmosphere to prepare graphene–CNT hybrid papers.12 Additionally, they prepared the test samples without the carbon additives, binders and metal current collectors, which would be different from the conventional slurry-casting method to prepare the LIBs anode.2,15,16 Including the essential prerequisite of flexible and wearable electronic devices. It is important to investigate the electrochemical properties stability before and after bending many times.17–19 So, to prepare the flexible anode materials keeping the flexible characterization with a simple technology, obtaining the excellent and stable electrochemical performance, especial before and after bending testing still remains a great challenge.

Herein, we report a rational strategy to synthesize well-dispersed ultrasmall SnO2 particles decorated on 3D non-woven cotton covered by graphene (CGN/SnO2) as an advanced anode material. The products are prepared by a simple in situ and thermal annealing process. The SnO2 nanoparticles with several manometers in size homogenously anchor on the graphene sheets, and act as spacers to keep the graphene sheets separate with each other. The graphene sheets can be firmly adsorbed on the cotton fibers to protect the cotton fibers in thermal annealing process. Furthermore, the graphene sheets keep the 3D composite conductivity and the excellent flexibility during the SnO2 nanoparticles crystallization process. This unique CGN/SnO2 composite electrode used as free-standing, binder-free anode excluding the current collectors, exhibits the excellent electrochemical performance. Especially, before and after bending many times, this anode material still remains a high reversible capacity, excellent cycling and a high rate capability. This work paves a new way in preparing the flexible anode materials for the flexible and wearable electronic devices application.

2. Experimental section

2.1 Materials synthesis

Graphene oxide (GO) was synthesized by Hummers method. 1 g of graphite powder and 0.5 g NaNO3 were added to 40 mL of H2SO4 solution and left stirring for 30 min at ambient temperature. Then 3 g of KMnO4 was slowly added, and stirring was continued for 4 h in the ice water. After increasing the temperature to 100 °C, 100 mL distilled water was added following the 30 min stirring. Finally, 10 mL H2O2 was added and the dispersion colour became golden yellow. For purification, the resulting mixture was centrifuged rinsed first with 5% HCl solution and then with deionized water three times. Aqueous dispersion of GO with 7 mg mL−1 was prepared by stock dispersion. After GO dispersion bath ultrasonic for 1 h, the non-woven cotton piece was immersed into the GO dispersion for 30 min. The products were subsequently frozen at −50 °C in cryogenic refrigerator for 24 h, and freeze-drying for 48 h to obtain cotton/graphene oxide (CGO) composite. For comparison, the non-woven cotton piece was immersed into a 0.5 M glucose, and then dried for 12 h at 60 °C to obtain cotton–glucose composite.

In a typical synthesis of carbon/SnO2 (C–SnO2) composite, 3 g SnCl2·2H2O was dispersed in 100 mL ethanol and stirred for 4 h at 60 °C. Then the non-woven cotton piece, CG and CGO composites were immersed in the above solution for 24 h at room temperature. Afterward, the non-woven cotton piece, cotton–glucose and CGO composites were taken out and dried for 12 h at 60 °C. Finally, cotton/SnO2 (C/SnO2), cotton glucose/SnO2 (CG/SnO2), and cotton covered by graphene/SnO2 (CGN/SnO2) composites were obtained by a tube furnace in N2 atmosphere for 1 h at 500 °C. For comparison, the non-woven cotton piece, cotton–glucose and CGO with same annealing treatment were carried out to prepare carbonized cotton, carbonized cotton–glucose (CG), and carbonized cotton/graphene (CGN) composite.

2.2 Characterization

Crystallographic information of the as-prepared composites was characterized by X-ray diffraction (XRD) machine (D&A25ADVANCE, BRUKER, Germany), using Cu Kα radiation (λ = 0.15418 nm). The microstructures of the as-prepared samples were observed by scanning electron microscope (SUPRA™55, ZEISS, United Kingdom). Transmission electron microscopy (TEM) images and the corresponding selected-area electron diffraction (SAED) pattern were captured on a microscope (Tecnai G2 F30, FEI, America). The thermal stability, using thermogravimetric analysis (TGA) (TGA/SDTA851#, METTLER TOLEDO, Switzerland) was carried on a heating rate of 10 °C min−1 in air from room temperature to 800 °C. The microstructures were investigated by Raman spectroscopy (inVia, Renishaw, England) with an excitation wavelength of 532 nm from 800 to 2000 cm−1.

2.3 Electrochemical measurement

All the samples were cut into discs (10 mm in diameter) as the working electrodes without any additional conductive agent, binder and current collector. The weight and thickness of electrodes were 2.0–2.2 mg and 100–120 μm for C/SnO2, 2.2–2.4 mg and 100–120 μm for CG/SnO2, and 3.9–4.1 mg and 140–160 μm for CGN/SnO2, respectively. The electrolyte was 1 mol L−1 solution of LiPF6 in a mixture of ethylene carbonate (EC)–diethyl carbonate (DMC) (1[thin space (1/6-em)]:[thin space (1/6-em)]1, vol%). All the cells were assembled in an Ar-filled glove box with the C/SnO2, CG/SnO2, and CGN/SnO2 electrodes as the working electrode and lithium foil as the counter electrode. Galvanostatic discharge and charge were measured at current density of 260 μA cm−2 between 0.001 and 3 V using a battery test instrument (LAND CT2001A model, Wuhan SHENGLAN, China). Cyclic voltammogram (CV) tests were conducted on the workstation (CHI760E, Chenhua, China) at a scanning rate of 0.5 mV s−1 in a potential range of 0.005–3 V. A three-electrode cell was used for electrochemical impedance spectroscopy (EIS) test over a frequency range from 100 kHz to 0.01 Hz, where lithium foil acts as both the counter and reference electrodes. All the electrochemical tests were conducted at room temperature.

3. Results and discussion

The fabrication process of the CGN/SnO2 composite is illustrated in Fig. 1a. The non-woven cotton, a continuous structure with an interconnected 3D scaffold, is selected as the skeleton to prepare the self-supporting and flexible anode materials. Due to the electrostatic interaction, van der Waals' force and hydrogen bond exist between the oxygen-containing groups on GO and cotton.20,21 The non-woven cotton is homogenously coated with GO immersing into GO aqueous suspension (Fig. S1). Difference from other reports using oven dry technique,20,22 the freeze-drying technique is chosen to remove water from a frozen sample by sublimation and desorption under vacuum, maintaining the non-woven cotton original fluffy structure in the CGO. The colour of non-woven cotton changes from white to brown as shown in Fig. 1b and c. When the CGO is introduced into SnCl2 solution, the Sn2+ bonds with the oxygenated groups of graphene oxide and cotton fibers by electrostatic forces with the oxygenated groups. And then, the SnO2 nanoparticles could nucleate and simultaneously grow in situ on the graphene sheets and cotton fibers. As shown in Fig. 1d, after about 10 min, the colour of the CGO changes from brown to black, indicating the reduction of a certain degree of GO by Sn2+. To obtain the CGN/SnO2, a thermal reduction under N2 atmosphere is critical to prepare the free-standing CGN/SnO2. The composites keep the flexible characterization and the SnO2 nanoparticles homogeneously decorate on the both sides of the graphene and the surface of the original cotton fibers, although a small shrinkage of the non-woven cotton skeleton occurring (Fig. 1e). The CGN/SnO2 is completely recovered back to its original shape with no mechanical failure even under the excessive bending and warping (ESI Movie).
image file: c6ra19347a-f1.tif
Fig. 1 Schematic illustration of synthesis of the CGN/SnO2 (a), photograph of non-woven cotton (b), CGO (c), non-woven covered by graphene with SnO2 (d), and CGN/SnO2 (e).

Fig. 2a shows the XRD patterns of above C–SnO2 composites. All the XRD patterns display several main diffraction peaks, which could be indexed to the faces of (110), (101), (200), (111), (211), (220), (002), (310), (112), (301) and (321) corresponding to the rutile SnO2 (JCPDS card no. 41-1445).23,24 No peaks of possible impurities, such as SnO or Sn, indicating the high purity of the obtained samples. The absence of carbon peaks suggests that carbonized cotton with the lower order structure, and graphene layer is exfoliated completely.23,25 Fig. 2b–d show the Raman spectra of C–SnO2 composites with different carbon framework. The Raman spectra in the region 800–2000 cm−1 show the characteristic peaks of carbonaceous materials. The D peak around 1340 cm−1 is associated with the structural defects in the graphitic plane. The G peak around 1580 cm−1 is attributed to the sp2 bonded graphitic carbons.26,27 The G peaks of the C/SnO2 (1587.6 cm−1), CG/SnO2 (1586.2 cm−1) and CGN/SnO2 (1584.8 cm−1) have blue shifts comparing with carbonized cotton (1585.7 cm−1), CG (1582.5 cm−1) and CGN (1583.8 cm−1), respectively, revealing the p-type doping effect on carbon framework.28,29 The p-type doping effect on carbon framework discloses the electronic interactions between SnO2 nanoparticles and carbon framework, which facilitates the formation of desirable 3D electronic networks and favours electron transportation between SnO2 and carbon framework.30 The area intensity ratios of D peak to G peak (ID/IG) of C/SnO2 (2.07), CG/SnO2 (1.95) and CGN/SnO2 (1.91) are larger than cotton (1.93), CG (1.84) and CGN (1.76), respectively. Suggesting doping SnO2 could decrease the average size of the sp2 domains upon the carbon framework.31 The higher ID/IG value indicates the smaller carbon size and more disorder structure of carbon framework.32,33 Moreover, the disorder structure is caused by vacancies and topological defects, which decreases in the thermal stability of materials at high temperatures. The cotton covered by graphene framework has the lowest value of ID/IG, indicating the graphene can decrease the production of vacancies and topological defects, and increase thermal stability.


image file: c6ra19347a-f2.tif
Fig. 2 XRD patterns of C/SnO2, CG/SnO2 and CGN/SnO2 (a), Raman spectra of carbonized cotton and C/SnO2 (b), CG and CG/SnO2 (c), CGN and CGN/SnO2 (d) composites, respectively.

The morphology and microstructure of the as-synthesized C–SnO2 samples are examined by SEM. The non-woven cotton, as a self-supporting skeleton, mainly consists of flower-shaped hub cotton fibers with diameters ranging from 10 to 20 μm (Fig. S2). After immersing in SnCl2 solution and carbonization, the morphologies of the C/SnO2 cotton fibers have no obvious changes, but some aggregation bulks on the fibers surface can be easily observed (Fig. S3a). The high magnification SEM image and corresponding EDX mapping show the aggregation bulks mainly consist of Sn and O element from SnO2 (Fig. S3b–e). For the cotton fibers covered by glucose, as shown in Fig. S3f–j, nanosized SnO2 particles with a much lesser degree of aggregation are uniformly deposited on the surface of CG framework. Glucose is a hydroxyl-enriched molecule, and a plenty of –OH would be available to facilitate the SnO2 crystallization, which would obstruct the particles aggregation.34 The preparation process shows great versatility in controlling both the macrostructure and the microstructure of CGN/SnO2 composites. From Fig. 3a, the graphene coating exists not only on the surface of the fibers but also filling the gap. The graphene is an effective connection bridge between the adjacent cotton fibers. The high magnification image (Fig. 3b) can clearly illustrate graphene wraps on the surface of cotton fiber, and many graphene wrinkles connect with each other. Moreover, through the cross section of CGN/SnO2 fibers, graphene wrinkles with hollow structures can be observed (Fig. S4a and e), which makes a lot of connecting pipes. Those structures greatly increase the electrode/electrolyte contact areas, and accelerate the electron and Li ions transport speed. Neither bulks nor particles can be observed on the CGN surface. Further investigation of EDS mapping shown in Fig. 3c–e indicates that all C, Sn and O elements are homogeneously distributed on the CGN framework. The as-formed ultra-small SnO2 nanocrystals may act as an active matrix to stabilize the structure. Compare with glucose, GO has more oxygen containing groups, such as hydroxyl, epoxy, carbonyl and carboxylic on sheets. This oxygen groups not only promote the connection with cotton fibers, but also are used to decorate and separate ultrasmall SnO2 nanoparticles.32 Element mappings of the CGN/SnO2 in cross section reveal the uniformly distribution of C, Sn and O elements (Fig. S4b–d). We can reasonably conclude that the Sn2+ diffuses inside the cotton fiber and forms SnO2 nanoparticles. The SnO2 nanoparticles with ultrasmall sizes on CGN framework can greatly shorten the solid-state diffusion distance of Li ions, leading to faster charge transfer reaction at the electrode/electrolyte interfaces.35,36


image file: c6ra19347a-f3.tif
Fig. 3 SEM images of the CGN/SnO2 with low (a), and high magnification (b), EDS mapping images of C (c), Sn (d), O (e) elements.

Furthermore, the morphology and crystal structure of the C–SnO2 samples are characterized by TEM and HRTEM. As shown in Fig. 4a and c, some large particles or particles agglomerations are deposited on the fiber of C/SnO2 and CG/SnO2. The ring-like mode in the selected-area electron diffraction (SAED) pattern (inset images of Fig. 4a and c) and HRTEM images (Fig. 4b and d) confirm the presence of polycrystalline SnO2 with a tetragonal rutile-like crystal structure. It may be caused by less attractive force between cotton or glucose and Sn2+, and the particles are more likely to deposit or aggregate on the fibers surface. The aggregation of the SnO2 particles will be further reunited during Li ions insertion/extraction, causing volume expansion and capacity fading. The TEM image and SAED pattern of CGN/SnO2 in Fig. 4e and g show that the cotton fibers and graphene sheets are uniformly covered by ultrasmall SnO2 nanoparticles. Before test, the composite is treated under a strong ultrasonic treatment for preparing the TEM samples. The SnO2 nanoparticles can be firmly decorated on the CGN framework, indicating the SnO2 good contact with CGN. The HRTEM image shown in Fig. 4h reveals graphene sheets with multiple layers. Ultrasmall SnO2 nanoparticles with several nanometers in size are uniformly dispersed on the cotton and graphene sheets (Fig. 4f and h). Clear lattice fringes with d spacing of 0.21 and 0.33 nm are observed, which can be attributed to the (210) and (110) planes of rutile SnO2, respectively.37 The graphene has a beneficial effect on the nucleation, growth and formation of ultrasmall SnO2 nanoparticles with a uniformly mono-dispersion on cotton fibers and graphene layers. The nanostructures of SnO2 are helpful for the electrolyte permeating into the inner SnO2 particles and further increasing the contact areas of the electrode/electrolyte. In this way, more active materials could effectively participate in the faradic reactions at the interface and significantly improve the specific capacity.35,37 The theoretical specific capacity of the free-standing C/SnO2, CG/SnO2 and CGN/SnO2 composite can be calculated by TGA (Fig. S5). The SnO2 content in the C/SnO2, CG/SnO2 and CGN/SnO2 determined by TGA is determined to be 40.11 wt%, 36.11 wt% and 45.80 wt%, and the carbon content is 59.89 wt%, 63.89 wt% and 54.20 wt%, corresponding to the theoretical specific capacity of 403.0, 340.8 and 454.0 mA h g−1, respectively. The specific capacities of carbonized cotton, CG, and capacity = (carbon content × related capacity + SnO2 content × 782 mA h g−1). The specific capacities of carbonized cotton, CG, and CGN are shown in Fig. S6.


image file: c6ra19347a-f4.tif
Fig. 4 TEM images of C/SnO2 (a), CG/SnO2 (c), CGN/SnO2 cotton part (e), and CGN/SnO2 graphene part (g), and HRTEM images of C/SnO2 (b), CG/SnO2 (d), CGN/SnO2 cotton part (f), and CGN/SnO2 graphene part (h).

The sheets resistance of fabrics were measure using a standard four-point probe method in previous reports, which should be called the surface resistance.20,38 Because the non-woven cotton is fluffy, it is hard to use four-point probe method test the resistance. We test the bulk resistance of the flexible bulk CGN/SnO2 materials resistance using two copper electrodes in the upper and lower position as shown in Fig. 5a. According to the equation: ρ = RS/L, the R, L and S are the resistance, the depth of carbon chip and the area of the carbon chip, respectively. The calculated resistivity of C/SnO2, CG/SnO2 and CGN/SnO2 is 5.26 × 109, 6.01 × 104 and 3.31 × 102 Ω cm, respectively (Fig. 5b). Compared with the C/SnO2, the resistivity of the CG/SnO2 decreases five orders of magnitude due to the cotton covered by a carbon film after heat treatment. Compared with CG/SnO2, the CGN/SnO2 resistivity is even smaller than two orders. There may be three possible reasons for the improved conductivity. First, graphene with excellent electrical conductivity coating on the cotton fibers is an important reason for improving the resistance of material. Secondly, the change in the contacts modes between fibers is another important factor for improving the electrical properties. In the C/SnO2 and CG/SnO2, the fibers contacts with each other should be point to point. However, in the CGN/SnO2, graphene acting as a buffer layer covers on the cotton fibers to increase the contact area with the contact mode of surfaces to surfaces, like Fig. 5c. Third, as shown in Fig. 5d, for those parallel fibers, there are no direct contact with each other, so the electron transport kinetics of the C/SnO2 and CG/SnO2 composite can only be in one direction. In contrast, for the CGN/SnO2 composite, graphene sheets fill in the cotton fibers interspace to connect the parallel fibers, resulting in the electron transportation in any dimension with increased conductivity. As can be seen from Fig. 5b, the resistance of the CGN/SnO2 even after bending 1000 times is still low, which will have a positive influence on the electrochemical performance.


image file: c6ra19347a-f5.tif
Fig. 5 Schematic of bulk resistance test system (a), resistance of C/SnO2, CG/SnO2 and CGN/SnO2 composites (b), contact fibers schematic of C/SnO2 and CG/SnO2 (upper), and CGN/SnO2 (lower) (c), parallel fibers schematic of C/SnO2 and CG/SnO2 (upper), and CGN/SnO2 (lower) (d).

To identify the electrochemical process and then assembled into coin cells as a free-standing and binder-free working electrode without any additional conductive agent, binder or current collector. The CV measurement is performed as shown in Fig. 6. The CG/SnO2 and CGN/SnO2 have the similar CV curves (Fig. 6b and c). In the first cycle, two well defined cathodic peaks around 0.5 V and 0 V are observed. The peak around 0.5 V corresponds to the solid electrolyte interface (SEI) formation and the reductive transformation of SnO2 to Sn (4Li+ + SnO2 + 4e = Sn + 2Li2O).39 This cathodic peak shifted to higher voltages (around 1.0 V) in the subsequent cycles. The CG/SnO2 cathodic peak is more intensive than CGN/SnO2, which indicates to produce more SEI layer. The other cathodic peak nearly 0 V can be attributed to the formation of alloys (xLi+ + Sn + xe = LixSn) and lithium reaction with carbon framework (xLi+ + C + xe = LixC).40 In the anodic process, two oxidation peaks around 0.7 and 1.1 V are observed, representing the dealloying process of LixSn and partially reversible reaction of SnO2.35 There are no obvious SnO2 lithiation process peaks appears for C/SnO2 composite, only lithium reaction with carbon oxidation and reduction peaks in the CV curve (Fig. 6a), indicating the SnO2 of C/SnO2 contribute for lithium storage is negligible. From second cycle onwards, the CV curves of CG/SnO2 and CGN/SnO2 mostly overlap, indicating the good reversibility of the electrochemical reactions.


image file: c6ra19347a-f6.tif
Fig. 6 CV curves of C/SnO2 (a), CG/SnO2 (b), and CGN/SnO2 (c) composites.

To further comprehend the electrochemical performance of as-prepared samples, galvanostatic discharge/charge measurements are investigated at the current density of 260 μA cm−2 in the potential windows of 0.001–3.0 V. The typical discharge/charge curves in first cycle are shown in Fig. 7a, the CGN/SnO2 composites deliver a capacity of 5848.0 μA h cm−2 for the first discharge and a reversible capacity of 3087.9 μA h cm−2. The first coulombic efficient of CGN/SnO2 is calculated to be 52.8%, higher than that of C/SnO2 (16.1%) and CG/SnO2 (36.2%). The high capacity can be attributed to a synergetic effect between CGN and the SnO2 in the 3D framework, where Li ions are interspacing between neighbouring SnO2 nanoparticles,41 and graphene sheets are the most candidate to consume large lithium to form much SEI.42 The initial irreversibility of the electrode composite could be due to SEI formation and reduction of SnO2 to Sn, as the same analysis with CV.40,43,44


image file: c6ra19347a-f7.tif
Fig. 7 Charge–discharge curves (a), cycling performance (b), EIS impedance (c) of C/SnO2, CG/SnO2 and CGN/SnO2 composites; original and bending 1000 times of CGN/SnO2 rate test properties (d).

The cycling performance of the C/SnO2, CG/SnO2 and CGN/SnO2 samples at a current density of 260 μA cm−2 is shown in Fig. 7b. The CGN/SnO2 showed the best electrochemical performance, delivering a high reversible capacity of 2512.5 μA h cm−2 after 60 cycles, which is about 80.0% capacity retention compared with that in the second cycle. It is worth to note that from the third cycle onwards, the coulombic efficient is higher than 97%, demonstrating that the CGN/SnO2 composite of quick reaching stable cycling. The specific capacity of CGN composite is only 307.5 μA h cm−2 after 60 cycles (Fig. S7), indicating the high capacity of CGN/SnO2 is major contribute from SnO2 nanoparticles. The SEM image in Fig. S8(c) demonstrate that the CGN/SnO2 maintains their overall morphology after cycling. The CG/SnO2 composites show higher than 3000 μm A cm−2 in first cycle, however, relatively capacity drastic decay to 275.2 μm A cm−2 after 60 cycles, only 23.1% capacity retention compared with the second cycle. The loss of capacity after extended cycling may be due to the severe volume variation of the aggregated particles and electrode pulverization.45 As seen from Fig. S8b, the CG/SnO2 electrode materials become small breaks after cycling, this is caused by huge volume expansion/contraction up cycling.46,47 Although the C/SnO2 have a good cycling stability, but the reversible capacity is only 294.9 μA h cm−2 after 60 cycles, which is much lower than that of CGN/SnO2. As shown in Fig. S8a, there is obvious deformation on the cotton fibers, where SnO2 aggregation bulks have been split, so the storage capacity is only contributed from carbonized cotton framework. In another point, the CGN/SnO2 composites areal capacity is much higher than that of materials reported, such as CoO nanowires/carbon cloth (1950 μA h cm−2 after 90 cycles),48 SnO2@Si nanowire/carbon cloth (1390 μA h cm−2 after 50 cycles),49 Cu–Sn nanowires (70 μA h cm−2 after 100 cycles),50 Si-CNT nanocomposites (1000 μA h cm−2 after 50 cycles),51 CoNiO nanowire/TiO2 nanotubes (360 μA h cm−2 after 40 cycles),52 SnO2/Fe2O3 nanotubes (730 μA h cm−2 after 50 cycles),53 and carbon-coated porous silicon anodes (180 μA h cm−2 after 70 rate cycles).54

To better understand the beneficial effect of composites, the electrochemical impedance measurements of C/SnO2, CG/SnO2 and CGN/SnO2 are performed. The spectra are shown in Fig. 7c, and a simplified equivalent circuit model is constructed to analyze the impedance spectra shown inset of Fig. 7c. It is worth to note that the CG/SnO2 and CGN/SnO2 have two semicircles, a high-frequency semicircle and a medium-frequency semicircle. However, the C/SnO2 only has one semicircle from high to medium frequency due to the high-frequency and medium-frequency semicircle partially coincidence. The hard distinguish high-frequency and medium-frequency may be caused by the low conductive and SnO2 bulks hard for electron transport and charge-transfer of C/SnO2. An intercept at the Zreal axis in high frequency corresponds to the ohmic resistance (RΩ), which represents the resistance of the electrolyte. The high-frequency semicircle is attributed to the Li-ion migration resistance (RSEI) through the SEI film. The semicircle in the medium-frequency region is assigned to charge-transfer resistance (Rct). The inclined line is associated with Warburg impedance (W) corresponding to the lithium-ion diffusion process. The equivalent series resistance, RΩ and R = RSEI + Rct, for the CGN/SnO2 anode is thus estimated to be 7.3 Ω and 119.1 Ω, which is much smaller that the CG/SnO2 (12.1 Ω and 155.2 Ω) and C/SnO2 (24.6 Ω and 169.1 Ω). The decrease in R resistance of CGN/SnO2 electrode can be attributed to the high electronic conductivity of CGN and fine contact of SnO2 nanoparticles on the CGN framework. This result is also in agreement with the superior cycling performance of CGN/SnO2, which is beneficial for lithium ions insertion/extraction into the anodes under the high charge–discharge cycling conditions.23,45,55,56

The rate properties of CGN/SnO2 composite characterized at various current densities are shown in Fig. 7d. The discharge capacity after 5 cycles is 2060.4 μA h cm−2 at a current density of 220 μA cm−2. As obvious increase in current density from 440 to 3520 μA cm−2, the discharge capacities of 1744.1, 1491.2, 1124.4 and 575.0 μA h cm−2 are obtained, respectively. Even under such rigorous testing conditions, as the current density is reduced to 220 μA cm−2, the capacity almost recovers to the original value, indicating the CGN/SnO2 composite is quite suitable for large current charge and discharge. Of importance for the flexible electronic devices is that performance stability of the capacitive characteristic before and after bending many times for the samples. The CGN/SnO2 composite is bent from 0° to 180° as shown in the video file, and back to the initial state for 1000 times. The capacity shows a slight decrease at different current density after 1000 times bending (Fig. 7d). It is reasonable to conclude that the CGN/SnO2 composite electrode should be a good candidate for textile flexible and wearable electronic devices applications.

The superior reversible capacity, excellent cycling, high coulombic efficiency, and good rate capacity performance of CGN/SnO2 composite can be attributed to the following several merits: (1) the non-woven cotton covered by the flexible graphene acts as flexible 3D carbon supports serving as a buffer to release the stress during cyclic processes. In addition, the CGN framework showing the excellent electronic conductivity decreases the inner resistance of the LIBs, facilitate good transport of electrons ions, resulting in a higher specific capacity. (2) The ultrasmall sizes of the SnO2 nanoparticles uniformly disperse on the graphene sheets and cotton fibers. Due to the short transport length and more accessible sites for the active materials, which improves the coulombic efficient, excellent cycling and high reversible capacity performance. (3) The SnO2 nanoparticles tightly decorate on the cotton covered by graphene framework, which can significantly reduce the strain and agglomeration during the lithiation/delithiation processes. (4) The interconnected pipeline structure can provide large electrode/electrolyte contact area, and rapid mass and electron transport of Li ion, which facilitates the lithium diffusion kinetics and could lead to excellent rate capability. So it can be concluded that CGN/SnO2 composites have great potential applicable for flexible electrochemical devices application.

4. Conclusions

In summary, using the non-woven cotton as a flexible and bendable skeleton, we developed a facile strategy to synthesize a self-supporting 3D structure of electrically conductive graphene anchored with SnO2 nanoparticles as an advanced anode material for high performance LIBs. The graphene coating on the non-woven cotton protects cotton fibers and shows the excellent electronic conductivity, which facilitates electrons and electrolyte transport. The ultrasmall SnO2 nanoparticles are uniformly anchored on the graphene sheets and cotton fibers, and reduce the strain and agglomeration in lithium ions insertion/extraction. Even after bending 1000 times, the capacitive performance of the CGN/SnO2 composites is almost no obvious loss. Most importantly, this synthesize method presented in this work may be extended to other active materials with CGN framework for large-scale promising anode material in high-performance energy storage field.

Acknowledgements

This project was supported by the National Natural Science Foundation of China (51172050, 51102060, 51102063, 51302050, and 51372052), Shandong Province Young and Middle-Aged Scientists Research Awards Fund (BS2013CL003), and the Fundamental Research Funds for the Central Universities (HIT. ICRST. 2010009).

Notes and references

  1. L. Hu, H. Wu, F. La Mantia, Y. Yang and Y. Cui, ACS Nano, 2010, 4, 5843–5848 CrossRef CAS PubMed.
  2. G. Zhou, F. Li and H.-M. Cheng, Energy Environ. Sci., 2014, 7, 1307–1338 CAS.
  3. M. Koo, K. I. Park, S. H. Lee, M. Suh, D. Y. Jeon, J. W. Choi, K. Kang and K. J. Lee, Nano Lett., 2012, 12, 4810–4816 CrossRef CAS PubMed.
  4. X. Jia, C. Yan, Z. Chen, R. Wang, Q. Zhang, L. Guo, F. Wei and Y. Lu, Chem. Commun., 2011, 47, 9669–9671 RSC.
  5. X. Zhao, C. M. Hayner, M. C. Kung and H. H. Kung, ACS Nano, 2011, 5, 8739–8749 CrossRef CAS PubMed.
  6. L. Shen, Q. Che, H. Li and X. Zhang, Adv. Funct. Mater., 2014, 24, 2630–2637 CrossRef CAS.
  7. J. F. Ihlefeld, P. G. Clem, B. L. Doyle, P. G. Kotula, K. R. Fenton and C. A. Apblett, Adv. Mater., 2011, 23, 5663–5667 CrossRef CAS PubMed.
  8. J. Jiang, J. Zhu, R. Ding, Y. Li, F. Wu, J. Liu and X. Huang, J. Mater. Chem., 2011, 21, 15969 RSC.
  9. N. Li, G. Zhou, F. Li, L. Wen and H.-M. Cheng, Adv. Funct. Mater., 2013, 23, 5429–5435 CrossRef CAS.
  10. R. B. Rakhi, W. Chen, D. Cha and H. N. Alshareef, Adv. Energy Mater., 2012, 2, 381–389 CrossRef CAS.
  11. F. Liu, S. Song, D. Xue and H. Zhang, Adv. Mater., 2012, 24, 1089–1094 CrossRef CAS PubMed.
  12. Y. Hu, X. Li, J. Wang, R. Li and X. Sun, J. Power Sources, 2013, 237, 41–46 CrossRef CAS.
  13. J. Liang, Y. Zhao, L. Guo and L. Li, ACS Appl. Mater. Interfaces, 2012, 4, 5742–5748 CAS.
  14. A. N. Gildea, D. Wang and G. G. Botte, J. Appl. Electrochem., 2015, 45, 217–224 CrossRef CAS.
  15. L. F. Cui, L. Hu, J. W. Choi and Y. Cui, ACS Nano, 2010, 4, 3671–3678 CrossRef CAS PubMed.
  16. J. Ji, J. Liu, L. Lai, X. Zhao, Y. Zhen, J. Lin, Y. Zhu, H. Ji, L. L. Zhang and R. S. Ruoff, ACS Nano, 2015, 9, 8609–8616 CrossRef CAS PubMed.
  17. H. Wu, Q. Meng, Q. Yang, M. Zhang, K. Lu and Z. Wei, Adv. Mater., 2015, 27, 6504–6510 CrossRef CAS PubMed.
  18. K. Rana, S. D. Kim and J. H. Ahn, Nanoscale, 2015, 7, 7065–7071 RSC.
  19. M.-S. Balogun, M. Yu, Y. Huang, C. Li, P. Fang, Y. Liu, X. Lu and Y. Tong, Nano Energy, 2015, 11, 348–355 CrossRef CAS.
  20. L.-L. Xu, M.-X. Guo, S. Liu and S.-W. Bian, RSC Adv., 2015, 5, 25244–25249 RSC.
  21. G. Yu, L. Hu, M. Vosgueritchian, H. Wang, X. Xie, J. R. McDonough, X. Cui, Y. Cui and Z. Bao, Nano Lett., 2011, 11, 2905–2911 CrossRef CAS PubMed.
  22. I. A. Sahito, K. C. Sun, A. A. Arbab, M. B. Qadir and S. H. Jeong, Electrochim. Acta, 2015, 173, 164–171 CrossRef CAS.
  23. L. Liu, M. An, P. Yang and J. Zhang, Sci. Rep., 2015, 5, 9055 CrossRef CAS PubMed.
  24. X. Zhao, B. Liu and M. Cao, RSC Adv., 2015, 5, 30053–30061 RSC.
  25. B. Li, J. Zai, Y. Xiao, Q. Han and X. Qian, CrystEngComm, 2014, 16, 3318 RSC.
  26. H.-P. Cong, S. Xin and S.-H. Yu, Nano Energy, 2015, 13, 482–490 CrossRef CAS.
  27. S. J. Prabakar, Y. H. Hwang, E. G. Bae, S. Shim, D. Kim, M. S. Lah, K. S. Sohn and M. Pyo, Adv. Mater., 2013, 25, 3307–3312 CrossRef CAS PubMed.
  28. C. Xu, J. Sun and L. Gao, J. Mater. Chem., 2011, 21, 11253 RSC.
  29. B. Das, R. Voggu, C. S. Rout and C. N. Rao, Chem. Commun., 2008, 5155–5157 RSC.
  30. C. Tan, J. Cao, A. M. Khattak, F. Cai, B. Jiang, G. Yang and S. Hu, J. Power Sources, 2014, 270, 28–33 CrossRef CAS.
  31. Y. Li, X. Lv, J. Lu and J. Li, J. Phys. Chem. C, 2010, 114, 21770–21774 CAS.
  32. H. Liu, L. Zhang, Y. Guo, C. Cheng, L. Yang, L. Jiang, G. Yu, W. Hu, Y. Liu and D. Zhu, J. Mater. Chem. C, 2013, 1, 3104 RSC.
  33. W. Luo, B. Wang, C. G. Heron, M. J. Allen, J. Morre, C. S. Maier, W. F. Stickle and X. Ji, Nano Lett., 2014, 14, 2225–2229 CrossRef CAS PubMed.
  34. T. Zhu, J. S. Chen and X. W. Lou, J. Phys. Chem. C, 2011, 115, 9814–9820 CAS.
  35. Y. Li, H. Zhang and P. Kang Shen, Nano Energy, 2015, 13, 563–572 CrossRef CAS.
  36. J. S. Chen and X. W. Lou, Small, 2013, 9, 1877–1893 CrossRef CAS PubMed.
  37. S. H. Choi, J.-K. Lee and Y. C. Kang, Nano Res., 2015, 8, 1584–1594 CrossRef CAS.
  38. I. A. Sahito, K. C. Sun, A. A. Arbab, M. B. Qadir and S. H. Jeong, Carbohydr. Polym., 2015, 130, 299–306 CrossRef CAS PubMed.
  39. J. Cheng, H. Xin, H. Zheng and B. Wang, J. Power Sources, 2013, 232, 152–158 CrossRef CAS.
  40. C. Xu, J. Sun and L. Gao, J. Mater. Chem., 2012, 22, 975–979 RSC.
  41. C.-T. Hsieh, J.-S. Lin, Y.-F. Chen and H. Teng, J. Phys. Chem. C, 2012, 116, 15251–15258 CAS.
  42. O. Vargas, A. Caballero, J. Morales and E. Rodriguez-Castellon, ACS Appl. Mater. Interfaces, 2014, 6, 3290–3298 CAS.
  43. Z.-S. Wu, G. Zhou, L.-C. Yin, W. Ren, F. Li and H.-M. Cheng, Nano Energy, 2012, 1, 107–131 CrossRef CAS.
  44. X. Wang, X. Cao, L. Bourgeois, H. Guan, S. Chen, Y. Zhong, D.-M. Tang, H. Li, T. Zhai, L. Li, Y. Bando and D. Golberg, Adv. Funct. Mater., 2012, 22, 2682–2690 CrossRef CAS.
  45. W. Li, D. Yoon, J. Hwang, W. Chang and J. Kim, J. Power Sources, 2015, 293, 1024–1031 CrossRef CAS.
  46. D. Larcher, S. Beattie, M. Morcrette, K. Edström, J.-C. Jumas and J.-M. Tarascon, J. Mater. Chem., 2007, 17, 3759 RSC.
  47. L. Zhang, G. Zhang, H. B. Wu, L. Yu and X. W. Lou, Adv. Mater., 2013, 25, 2589–2593 CrossRef CAS PubMed.
  48. S. Zhao, J. Guo, F. Jiang, Q. Su, J. Zhang and G. Du, J. Alloys Compd., 2016, 655, 372–377 CrossRef CAS.
  49. W. Ren, C. Wang, L. Lu, D. Li, C. Cheng and J. Liu, J. Mater. Chem. A, 2013, 1, 13433 CAS.
  50. G. F. Ortiz, M. C. López, R. Alcántara and J. L. Tirado, J. Alloys Compd., 2014, 585, 331–336 CrossRef CAS.
  51. Q. Xiao, Y. Fan, X. Wang, R. A. Susantyoko and Q. Zhang, Energy Environ. Sci., 2014, 7, 655–661 CAS.
  52. J. Yao, P. Xiao, Y. Zhang, M. Zhan, F. Yang and X. Meng, J. Alloys Compd., 2014, 583, 366–371 CrossRef CAS.
  53. W. Zeng, F. Zheng, R. Li, Y. Zhan, Y. Li and J. Liu, Nanoscale, 2012, 4, 2760–2765 RSC.
  54. E. Biserni, M. Xie, R. Brescia, A. Scarpellini, M. Hashempour, P. Movahed, S. M. George, M. Bestetti, A. Li Bassi and P. Bruno, J. Power Sources, 2015, 274, 252–259 CrossRef CAS.
  55. B. Jin, H.-B. Gu and K.-W. Kim, J. Solid State Electrochem., 2007, 12, 105–111 CrossRef.
  56. H. C. Shin, W. I. Cho and H. Jang, Electrochim. Acta, 2006, 52, 1472–1476 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra19347a

This journal is © The Royal Society of Chemistry 2016