Xiaoyan Hana,
Ran Lib,
Shengqiang Qiub,
Xiaofang Zhanga,
Qing Zhang*ac and
Yingkui Yangabc
aKey Laboratory of Resources Green Conversion and Utilization of State Ethnic Affairs Commission & Ministry of Education, South-Central University for Nationalities, Wuhan 430074, China. E-mail: qingzhang@mail.scuec.edu.cn
bSchool of Materials Science and Engineering, Hubei University, Wuhan 430062, China
cHubei Engineering Technology Research Centre for Energy Polymer Materials, School of Chemistry and Materials Science, South-Central University for Nationalities, Wuhan 430074, China
First published on 18th February 2019
SnO2/graphene nanocomposite was successfully synthesized by a facile sonochemical method from SnCl2 and graphene oxide (GO) precursors. In the sonochemical process, the Sn2+ is firstly dispersed homogeneously on the GO surface, then in situ oxidized to SnO2 nanoparticles on both sides of the graphene nanosheets (RGO) obtained by the reduction of GO under continuous ultrasonication. Graphene not only provides a mechanical support to alleviate the volume changes of the SnO2 anode and prevent nanoparticle agglomeration, but also serves as a conductive network to facilitate charge transfer and Li+ diffusion. When used as a lithium ion battery (LIB) anode, the SnO2/graphene nanocomposite exhibits significantly improved specific capacity (1610 mA h g−1 at 100 mA g−1), good cycling stability (retaining 87% after 100 cycles), and competitive rate performance (273 mA h g−1 at 500 mA g−1) compared to those of bare SnO2. This sonochemical method can be also applied to the synthesis of other metal-oxide/graphene composites and this work provides a large-scale preparation route for the practical application of SnO2 in lithium ion batteries.
To alleviate volume change during lithiation/delithiation process and enhance structural stability of SnO2, an effective strategy is to design nanostructured SnO2.14–17 However, the agglomeration of nanostructured SnO2 during the cycling process usually leads to fast capacity attenuation.18,19 Recently, nanocomposites have drawn wide attention for their superiorities to integrate multiple properties of different nanoscale building blocks to improve mechanical and electronic properties.20–24 Particularly, graphene-based nanocomposites with metal, metal oxides or polymers have also demonstrated particular mechanical, electronic, electrochemical and catalytic properties.25–28 Hence, various SnO2/graphene nanocomposites with enhanced electrochemical performance have been fabricated, in which graphene supplies a mechanical support to prevent volume changes and enhances electrical conductivity of composites.29–31 In addition, flexible graphene support with high surface area, porosity, electrical conductivity and chemical inertness can prevent nanoparticle agglomeration to achieve a stable electrode structure and provide good electric conductivity for the composite.32–34
Various methods have been used to prepare SnO2/graphene nanocomposites and each method has its own merits and demerits. For example, Lee et al. prepared SnO2/3D graphene hybrid material by a simple hydrothermal process,35 Wang and co-workers developed a ternary self-assembly approach to construct layered SnO2-graphene nanocomposites,36 and Zhang's group synthesized SnO2@graphene nanocomposites via a wet mechanochemical method.37 Compared to these reported methods, sonochemical methods are more favorable for the construction of nanostructures without involving any surfactants and complex processes, and the morphology and particle size can be regulated by ultrasonic waves.38–41 In fact, the basic principle of sonochemistry method is the acoustic cavitation phenomenon in a liquid, which arises high temperature and pressure, and high cooling rate instantaneously during the cavitation process.38,41 Such extreme conditions can trigger the formation of metal oxides on a nanometer scale with homogeneous particle size distribution.42–45 Herein, SnO2/graphene nanocomposite (SnO2/RGO) was synthesized via a simple and effective sonochemical method using SnCl2 and graphene oxide (GO) as precursors. When used as a LIBs anode, the as-prepared SnO2/RGO nanocomposite displays large specific capacity (1610 mA h g−1 at 100 mA g−1), good cycling performance (87% retention after 100 cycles) and competitive rate capability (273 mA h g−1 at 500 mA g−1).
In this work, a sonochemical method was employed to synthesize SnO2/RGO nanocomposite by using SnCl2 as reducing agent and graphene oxide (GO) as oxidizing agent. The possible reaction mechanism is proposed according to the following equations:37
SnCl2 + H2O ↔ Sn(OH)Cl + HCl | (1) |
Sn(OH)Cl + GO → SnO2/graphene + HCl | (2) |
In the sonochemical reaction, SnCl2 is firstly hydrolyzed to Sn(OH)Cl as shown in reaction (1), subsequently Sn(OH)Cl is oxidized to SnO2 and GO is reduced to graphene (RGO) under continuous ultrasonic forces. As shown in Fig. 1, Sn(OH)Cl can be dispersed homogeneously on the surface of two-dimensional graphene nanosheets in ethanol solution under constant ultrasonication. The ultrasound treatment triggers the redox reaction between SnCl2 and GO, and in situ synthesis of SnO2 nanoparticles onto the graphene surface.
The electrochemical performance was carried out using a 2032-type coin cells (CR2032). A lithium foil was used as the counter electrode and a Celgard-2400 microporous membrane was used as the separator. The working electrode consists of 80 wt% active material, 10 wt% acetylene black and 10 wt% polyvinylidene fluoride (PVDF) binder. A copper foil acted as the current collector. 1.0 M LiPF6 solution dissolved in ethylene carbonate (EC) and dimethylcarbonate (DMC) (1:1, v/v) was used as the electrolyte. The cells were assembled in an argon-filled glove box. The galvanostatic discharge/charge tests were studied between 0.001 and 2.0 V versus Li+/Li on a multichannel battery testing system (LAND CT2001A). Cyclic voltammetry (CV) experiments and electrochemical impedance spectroscopy (EIS) were performed on an electrochemical working station (CHI770E) using model cells. CV were recorded at scan rate of 0.2 mV s−1 in the potential range of 0.001–2.0 V. EIS was examined in a frequency range from 0.01 Hz to 100 kHz under AC amplitude of 5 mV.
Fig. 2 (a) TEM image of GO. (b) SEM image of bare SnO2. (c and d) SEM images of SnO2/RGO. The inset in (b) shows a TEM image of bare SnO2. |
XRD patterns of GO, bare SnO2 and SnO2/RGO are given in Fig. 3. GO displays only one broad peak located at 10.8°, indexing to the (001) plane of GO with an interlayer distance of 8.18 Å.47 Bare SnO2 and SnO2/RGO exhibit similar diffraction peaks, which are well indexed to the pure SnO2 (JCPDS no. 41-1445), indicating that the reduction of GO does not affect the crystal structure of SnO2. However, the typical stacking peak of graphene nanosheets located at 26° is also absent in the pattern of SnO2/RGO nanocomposite. This indicates that SnO2 particles are formed on the both sides of RGO nanosheets, which prevents graphene sheets from restacking.48 In addition, compared to sharp diffraction peaks of SnO2, SnO2/RGO nanocomposite displays broader XRD peaks, which is attributed to the decrease of particle size (bare SnO2: 10.8 nm, SnO2/RGO: 7.8 nm, calculated by Scherrer formula: D = Kλ/Bcosθ), suggesting that the growth of SnO2 nanoparticles on the surface of RGO nanosheets can prevent nanoparticles agglomeration. Furthermore, no other peaks can be observed, indicating the complete transformation of SnCl2 into SnO2 upon sonochemical and heat treatment process.
Fig. 4a shows FT-IR spectra of bare SnO2, GO and SnO2/RGO. FT-IR spectrum of GO is in good agreement with previous works.49 GO shows the O–H deformation peak at 1401 cm−1, the O–H stretching vibrations peak at 3420 cm−1 and the CO stretching vibrations peak at 1735 cm−1. The peaks at 1624, 1220, and 1052 cm−1 are due to CC bending vibrations, C–OH stretching vibrations, and C–O stretching vibrations, respectively. Those peaks weaken or almost disappear for the SnO2/RGO nanocomposite, implying that GO is completely reduced to RGO nanosheets during the formation of nanocomposite. Additionally, a strong peak at 624 cm−1 is assigned to Sn–O stretching vibrations, confirming the presence of SnO2 in the composite.48,49 The FT-IR results confirm the reduction of GO and the oxidation of Sn2+ to SnO2, suggesting the formation of SnO2/RGO nanocomposite.
Fig. 4 (a) FT-IR of GO, bare SnO2, and SnO2/RGO nanocomposite. (b) Raman spectra of GO, RGO, and SnO2/RGO nanocomposite. (c) TGA curve of SnO2/RGO nanocomposite in air atmosphere. |
Fig. 4b gives Raman spectra of GO, RGO and SnO2/RGO. The peaks at 1588 and 1344 cm−1 correspond to the G and D band, respectively. The G band is related to the presence of isolated double bonds in a 2-dimensional hexagonal lattice, while the D band is linked with the defects and disorder in the hexagonal graphitic layers. The relative intensity ratio of D to G bands (ID/IG) can represent the reduction degree of carbonaceous materials. The defect density is calculated to be 2.0 for SnO2/RGO nanocomposite, larger than that of GO (0.9) and RGO (1.0). This enhancement of defect density could be ascribed to the exfoliation and reduction of GO with the attachment of SnO2 nanoparticles on the surface.50
To obtain the content of RGO in the composite, thermogravimetric analysis (TGA) was measured under air as shown in Fig. 4c. In general, the negligible weight loss below 400 °C is attributed to adsorbed water molecules and oxygen-containing groups of graphene in the sample. The significant weight loss between 400 and 600 °C is attributed to the complete combustion of carbon components from RGO. Above 600 °C, the weight of the sample is constant, indicating that the combustion is completed and only SnO2 is left. The weight loss of SnO2/RGO nanocomposite is 18.2% (RGO component), which is different from pre-designed feed ratio of GO (30%). This deviation can be ascribed to the loss of oxygen-containing groups in GO during thermal reduction process.
XPS was further employed to confirm the reduction of GO and formation of SnO2 during the sonochemical process. Fig. 5a shows the survey spectra of GO, bare SnO2 and SnO2/RGO. The spectrum of GO only shows two distinguishable peaks of carbon (C 1s, 285 eV) and oxygen (O 1s, 530 eV). In contrast, the spectrum of SnO2/RGO shows all the peaks of desired carbon, oxygen and SnO2. The Sn 3d peaks of SnO2/RGO (Fig. 5b) could be resolved into 487.0 and 495.4 eV with an 8.4 eV peak-to-peak separation, corresponding to Sn 3d5/2 and Sn 3d3/2, respectively, which are consistent with the previous reports about SnO2/C composites.48,51 Additionally, the peaks of C 1s spectra of SnO2/RGO and GO could be resolved into four binding energies (Fig. 5c and d), corresponding to four types of carbon atoms (C–C/CC: 284.7 eV, C–O: 286.7 eV, CO: 288.7 eV, and O–CO: 289.6 eV). The C–C/CC content in SnO2/RGO (77.8%) is larger than that of GO (36.7%). Meanwhile, C–O amount in SnO2/RGO declines sharply due to successful removal of oxygen-containing groups in GO, which further confirms the completely reduction of GO.
Fig. 5 (a) XPS spectra of GO, bare SnO2 and SnO2/RGO. (b) Sn 3d XPS spectra of SnO2/RGO. C 1s XPS spectra of (c) SnO2/RGO and (d) of GO. |
Fig. 6a shows CV curves of SnO2/RGO for the first three cycles. It has been established that two-step reactions are involved in the SnO2-based electrodes:
SnO2 + 4Li+ + 4e− → Sn + 2Li2O | (3) |
Sn + xLi+ + xe− ↔ LixSn (0 ≤ x ≤ 4.4) | (4) |
There is an obvious cathodic peak at 0.72 V in the first cycle of the SnO2/RGO, which is attributed to the reduction of SnO2 to metallic Sn and formation of amorphous Li2O as indicated in reaction (3) as well as the formation of solid electrolyte interphase (SEI).31,50 The broad cathodic peak at about 0.25 V and anodic peak at about 0.65 V correspond to reversible reaction as indicated in reaction (4). The anodic peak at about 1.32 V represents partial reversibility of the reaction as shown in reaction (3).31 The CV curves of the second and third cycles are almost overlapped, implying an excellent cycling stability and reversibility of the SnO2/RGO. Furthermore, charge/discharge profiles of the SnO2/RGO are displayed in Fig. 6b. There appears two discharge plateaus around 0.8 V and 0.3 V in the first discharge process, originating from the formation of SEI film and alloying of LixSn. A charge plateau at 1.2 V in the first cycle is linked with dealloying of LixSn,13 which is consistent with the oxidation–reduction peaks of CV curves. The initial discharge and charge capacities of the SnO2/RGO electrode are 1610 and 765 mA h g−1, respectively. This large initial irreversible loss (52%) is common for SnO2 materials, which could be assigned to the irreversible transformation reaction as described in eqn (3) and the formation of solid electrolyte interphase (SEI) film.
Fig. 6c compares the cycling performances of SnO2/RGO with bare SnO2 and RGO. Apparently, bare RGO shows good cycling stability but low reversible capacity. After 100 cycles, the discharge capacity decreases to 263 mA h g−1. Even though bare SnO2 delivers a high discharge capacity (1024 mA h g−1), but the capacity tends to decay quickly to below 95 mA h g−1 after 100 cycles. In contrast, the cycling performance of SnO2/RGO electrode is substantially improved. Although there is still some capacity decay within the initial 20 cycles, the capacity of SnO2/RGO tends to level off in following cycles. It is important to note that SnO2/RGO delivers initial discharge capacity as high as 1610 mA h g−1 and a reversible capacity of 450 mA h g−1 can be retained over 100 cycles (C100th/C20th = 87%) with the coulombic efficiency of 100%. These results indicate that SnO2/RGO shows improved cycling stability and higher reversible capacity than bare SnO2. Moreover, to fully estimate the electrochemical performance of SnO2/RGO electrode, the rate performance at current densities of 100, 200, 300, 400 and 500 mA g−1 are also shown in Fig. 6d. As the current density increases from 100 to 500 mA g−1, the reversible capacity varies from 650 to 512, 360, 311 and 273 mA h g−1, respectively. Additionally, when the current density returns to 100 mA g−1, a reversible capacity of 639 mA h g−1 is recovered, thus demonstrating a competitive rate capability compared to the previously reported SnO2-based composites as shown in Table S1.†52–55 The above results show that the introducing of graphene nanosheets combines the respective advantages of SnO2 and RGO, thus greatly enhances the overall electrochemical capacity and cycling stability of the SnO2/RGO. In order to investigate the influence of GO content on the electrochemical performance of the composite, we also synthesized a series of SnO2/RGO nanocomposites with various component ratios via regulating the feeding ratios of GO and SnCl2 in the sonochemical reaction systems as shown in Fig. S1.† Considering that large amounts of graphene would in turn inevitably reduce the energy density of finished batteries, thus SnO2/RGO (30 wt%) is considered to exhibit the best comprehensively electrochemical performance.
EIS measurements were performed for the bare SnO2, RGO, and SnO2/RGO electrode to further elucidate the influence of graphene nanosheets on the electrochemical behavior of the composite (Fig. 7). The inset is the equivalent circuit used for fitting the impedance spectrum. Rs represents the ohmic resistance from the system and Cd is the double-layer capacity.28 In the Nyquist plots, a semicircle in higher frequency region and an inclined line of approximately 45° slope at lower frequency region were observed for all materials. The semicircle can be ascribed to charge transfer resistance (Rct) at the interface between the electrolyte and electrode, while the inclined line corresponds to Li+ diffusion process in the bulk of the electrode and represents the Warburg impedance (W).48 The values of Rct for the bare SnO2, RGO and SnO2/RGO are calculated to be 1088, 597, and 464 Ω, respectively. Apparently, SnO2/RGO shows smaller Rct than that of bare SnO2, which is ascribed to the enhanced electronic conductivity provided by the graphene substrate. Moreover, the high slope of inclined line for SnO2/RGO is a characteristic of low Zw, suggesting fast Li+ diffusion in the composite electrode. The results confirm that the presence of graphene nanosheets indeed facilitates charge transfer and Li+ diffusion during the electrochemical lithiation/delithiation process, which accounts for the enhanced electrochemical performance.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ra00554d |
This journal is © The Royal Society of Chemistry 2019 |