A facile post-process method to enhance crystallinity and electrochemical properties of SnO2/rGO composites with three-dimensional hierarchically porous structure

Fei Tiana, Xiaobing Wang*a, Zhenyu Chena, Yuming Guoa, Huijun Liangab, Zhansheng Luc, Dong Wanga, Xiangdong Lou*a and Lin Yang*a
aCollaborative Innovation Center of Henan Province for Green Manufacturing of Fine Chemicals, Key Laboratory of Green Chemical Media and Reactions, Ministry of Education, School of Chemistry and Chemical Engineering, Henan Normal University, Xinxiang, Henan 453007, P. R. China. E-mail: wxb95@163.com; yanglin1819@163.com; louxd16@163.com
bCollege of Chemistry and Chemical Engineering, Xinxiang University, Xinxiang, Henan 453003, P. R. China
cCollege of Physics and Information Engineering, Henan Normal University, Xinxiang, Henan 453007, P. R. China

Received 18th September 2016 , Accepted 31st October 2016

First published on 2nd November 2016


Abstract

In this work, three SnO2/reduced graphene oxide (SnO2/rGO) composites with a three-dimensional hierarchically porous structure were synthesized via freeze drying and different annealing temperatures in an air atmosphere. The results show that the crystallinity of SnO2 can be improved with increasing annealing temperature in air atmosphere, but the crystallinity quality of rGO presents a volcanic type change with increasing annealing temperature. The cyclic voltammetry curves indicate that the reaction of Li+ intercalation into SnO2 and rGO can only occur in the 120 °C sample, so that the cycling performance of the 120 °C sample is the best. Its initial discharge capacity is 1610 mA h g−1 at current densities of 0.5 A g−1. After 100 cycles, the specific discharge capacity is 429 mA h g−1. The XPS results reveal that SnO2 nanoparticles are anchored on the surface of rGO via Sn–O–C bonds. During the charge/discharge process, the volume expansion of SnO2 nanoparticles can be prevented through enhancing the crystallinity of ultra-small SnO2 nanoparticles (about 3.2 nm). This might pave the way for the commercial production of the SnO2/rGO electrode because these post-processing technologies are more easily implemented in industrial production.


1. Introduction

Lithium ion batteries (LIBs) with high capacities and long cycle lives have attracted considerable research activity due to the rapid development of electric vehicles (EVs) and portable electronic devices.1,2 As important components, anode materials are crucial for the next generation of high-performance rechargeable LIBs.3 Although graphite is the earliest commercial anode material in LIBs, its theoretical specific capacity is only 372 mA h g−1, which is insufficient to satisfy the increasing demand for batteries with higher capacity.4,5 Tin oxide (SnO2) based nanomaterials have many potential advantages as compared to graphite, such as high specific capacity (1494 mA h g−1), low cost, abundance and environmental benignity, etc.6 However, when SnO2 is employed as an anode material in LIBs, it will undergo an irreversible conversion and form a reversible Li4.4Sn alloy.7 This process shows a volume change of about 300% during charge/discharge process,8,9 which results in a poor cycling performance,10 rapid capacity decrease,11 and hindering the practical applications of SnO2 based anode materials in LIBs.12

To address these issues, one strategy is to design various SnO2 nanostructures, such as nanotube,13 nanobox,14 nanosheet,15 and hollow sphere,16 etc. These nanoarchitectures can overcome the large volume change of SnO2 through the space reserved and specific surface area. The other strategy is to minimize the physical strain through nanosized SnO2.17,18 Previous reports have indicated that SnO2 nanoparticles of ∼3 nm size could not aggregate into larger Sn clusters, and have a superior capacity and cycling stability.19 However, it is difficult to synthesize the ultra-small size of SnO2,17 and pure SnO2 nanoparticles are easy to aggregate owing to high surface energy of nanoparticles.20 Therefore, new approach is to integrate SnO2 with carbonaceous materials, such as SnO2/porous carbon,21 SnO2/carbon nanotubes22 and SnO2/reduced graphene oxide (SnO2/rGO),23 etc. The porous nature of carbonaceous materials is used to accommodate huge volume changes of SnO2, and the electronic conductivity of the electrode can be enhanced by the carbonaceous materials,24 so that SnO2/carbonaceous composite materials have attracted considerable attentions for the applications of energy and sensors.25,26

Among all carbonaceous materials, graphene has been regarded as a good matrix for SnO2 anode since it has ultrathin graphitic layers, excellent electronic conductivity, highly specific surface areas (2600 m2 g−1), good mechanical properties, and good chemical stabilities.27,28 Up to now, many chemical methods have been attempted to synthesize SnO2/rGO, such as one-pot route,29 electrostatic spray deposition,30 atomic layer and in situ deposition,7 hydrothermal/solvothermal,25,31 and so on. For the synthesis of SnO2/rGO composites, it is a facile method that graphene oxide (GO) is reduced by stannous ions without any reductants. However, the crystallinity of SnO2 and crystallization quality of rGO is poor in the SnO2/rGO, so that the SnO2/rGO composites remain highly resistive, and it is not optimal for energy storage applications.32 Although a high-temperature annealing process is frequently used to enhance the crystallinity of SnO2 and crystallization quality of rGO, the protective atmosphere is essential to prevent the decomposition of rGO in the most of previous reports.33 This high-temperature annealing process is a major disadvantage in the industrial production, because it needs a relatively high equipment requirements and operation condition as compared to low temperature and air atmosphere. Therefore, it is still important for how to improve the crystallinity of SnO2 and crystallization quality of rGO in industrial production. In addition, the volume expansion and aggregation of SnO2 nanoparticles can not be prevented by the SnO2/rGO composites,7 and SnO2 based nanoparticles could be easily peeled off from the graphene and agglomerate on the anode. All of these results show that SnO2/rGO composites still suffer from the large contact resistance,34 self-discharge, capacity loss and even electrode failure.2

Herein, we demonstrated that a facile method improve the crystallinity of SnO2 and crystallization quality of rGO, which is performed by a low-temperature calcining process in air. Before calcining, Li+ can only intercalate into rGO. However, the crystallinity of SnO2 and crystallization quality of rGO can be improved at 120 °C in air atmosphere. In this case, Li+ can intercalate into SnO2 and rGO. With increasing temperature to 280 °C, although the crystallinity of SnO2 can be improved, the crystallization quality of rGO decrease, and the reaction of the Li+ intercalation into rGO disappears. The morphology and structural characterization reveal that SnO2 nanoparticles are anchored by Sn–O–C bonds on graphene skeleton, and three-dimensional (3D) hierarchically porous structure of SnO2/rGO composites can be retained. Moreover, the volume expansion of SnO2 can be prevented through enhancing crystallinity of ultra-small SnO2 nanoparticles size (about 3.2 nm) at 120 °C. This approach could be an effective way to improve the electrochemical performances of SnO2/rGO composites in the industrial production, because it is simple and low-cost.

2. Experimental section

Tin dichloride dihydrate (SnCl2·2H2O, analytical grade purity) was purchased from the Tianjin Deen Chemical Reagent Co., Ltd (China). All other chemicals were of analytical grade and used without further purification.

2.1. Synthesis of SnO2/rGO composites

GO was synthesized from natural graphite powders by a modified Hummer's method.35 Then the as-prepared GO was dispersed in deionized water through the aid of ultrasound, the concentration of GO suspension was controlled at 1.7 mg mL−1. In a typical process, 2.0 g of SnCl2·2H2O and several tin metal grains were added to 200 mL deionized water under magnetic stirring for 15 min. Then, 200 mL of GO suspension was added to the above solution and stirred at room temperature for 12 h, and the black precipitate was obtained by centrifugation, washed with deionized water several times. The resultant product was collected after freeze-drying (−52 °C 24 h), and denoted as SnO2/rGO. As a comparison, the SnO2/rGO was calcined at 120 °C for 4 h in air, and named as SnO2/rGO-120. The SnO2/rGO was calcined at 280 °C for 2 h in air, and named as SnO2/rGO-280.

2.2. Characterizations

The microstructures and morphologies were observed using a SUPRA-40VP field emission scanning electron microscopy (FESEM, ZEISS, Germany) and a JEM-2100 transmission electron microscope (TEM, JEOL, Japan). The X-ray diffraction (XRD) patterns were obtained on the Bruker advance-D8 XRD instrument using CuKα radiation at 40 kV and 100 mA. X-ray photoelectron spectra (XPS) was detected with a VG Scientific ESCALABMKLL spectrometer using Al Kα X-ray source (10 mA, 15 kV). Nitrogen adsorption/desorption isotherms were determined with an ASAP 2020 (Micromeritics Instruments). Surface-area determination and pore distribution were evaluated by using the Brunauer–Emmet–Teller (BET) method and the Barrett–Joyner–Halenda (BJH) method. Raman spectra was collected using an inVia (Renishaw, England) with an excitation wavelength of 532 nm. The thermogravimetric and differential thermal analysis (TG-DTA) were performed on a STA449C Analyzer (NETZSCH, Germany) with a heating rate of 2.4 °C min−1 under air atmosphere. Fourier transform infrared spectrographs (FT-IR) were recorded on a Bio-Rad FTS-40 Fourier transform infrared spectrophotometer with KBr pellets.

2.3. Electrochemical measurements

The electrochemical properties were evaluated by coin-type cells (2016-type). Lithium metal foil was used as a counter electrode. The working electrodes were prepared by mixing 90 wt% of active materials (SnO2/rGO) and 10 wt% polyvinylidene fluoride (PVDF) dissolved in N-methylpyrrolidinone (NMP) to form a slurry, which was then coated onto a copper foil, dried in air at 120 °C overnight. The cell was assembled in an Ar-filled glove box (O2 and H2O < 0.1 ppm). The electrolyte was 1 M LiPF6 in a mixture of ethylene carbonate (EC) and dimethyl carbonate (DMC) (1[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v). Cyclic voltammetry (CV) tests were performed using CHI660D Electrochemical Workstation at a scanning rate of 0.1 mV s−1. The electrochemical impedance spectroscopy (EIS) was recorded on a CHI660D Electrochemical Workstation by applying a sine wave with amplitude of 5.0 mV over the frequency range from 100 kHz to 0.01 Hz. Galvanostatic charge/discharge measurement was carried out on a LAND CT2001A battery-testing instrument at various current rates. The voltage range was from 0.01 to 3.0 V for all tests.

3. Results and discussion

3.1. Morphology and structural characterization

The FESEM and TEM images of the three samples (SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280) are shown in Fig. 1. The FESEM images (Fig. 1a, e and i) of the three samples show a disordered highly opened macroporous structure, and these macroporous are formed by rGO thin films with dispersive SnO2 nanoparticles. However, the SnO2/rGO-280 (Fig. 1i) presents a collapse structure. According to the previous researches, these macroporous structure should come from the sublimation of the ice crystals,36 and the collapse structure of SnO2/rGO-280 should be due to the high calcinating temperature, so that 3D hierarchically porous structure was destroyed. The high-magnification FESEM images of SnO2/rGO and SnO2/rGO-120 (Fig. 1b and f) reveal that there are many blocks of 10–50 nm on thin films of rGO, and these blocks link each other to form mesopores with diameter of 10–50 nm. The thickness of SnO2/rGO films ranges from 10 to 25 nm. The FESEM image of SnO2/rGO-280 (Fig. 1j) is fuzzy, and the blocks of 10–50 nm disappear, but its lamellar and porous structure can still be observed explicitly.
image file: c6ra23236a-f1.tif
Fig. 1 The FESEM (a, b, e, f, i, j), TEM (c, g, k) and HRTEM (d, h, l) images of SnO2/rGO (a–d), SnO2/rGO-120 (e–l) and SnO2/rGO-280 (i–l), respectively. The insets in (d), (h), (l) are the SAED patterns of samples, respectively.

The TEM images (Fig. 1c, g and k) of the three samples show that the SnO2/rGO (Fig. 1c) is a smooth sheet-like structure, but SnO2/rGO-120 (Fig. 1g) is a wrinkled sheet-like structure, and SnO2/rGO-280 (Fig. 1k) is stacked together. The HRTEM images (Fig. 1d, h and l) of the three samples show that SnO2 nanoparticles were uniformly loaded on the surface of graphene layers, and their size is less than 5 nm nanocrystals. According to the HRTEM images, the grain size distributions of SnO2 nanoparticles are shown in Fig. S1. SnO2/rGO and SnO2/rGO-120 exhibited a relatively narrow grain size distribution. The average grain sizes of SnO2 nanoparticles in the SnO2/rGO, SnO2/rGO-120, and SnO2/rGO-280 composites are 2.9 ± 0.5 (Fig. S1a), 3.2 ± 0.5 (Fig. S1b), and 3.8 ± 0.5 nm (Fig. S1c) respectively. Although the size of SnO2 nanoparticles increased gradually according to the grain size distributions (Fig. S1), considering the measure error with TEM, the SnO2 nanoparticles size of the three samples should be in the same range. The selected-area electron diffraction (SAED, insets of Fig. 1d, h and l) patterns of the three samples contain a series of well-defined Debye–Scherrer rings, confirming their polycrystalline nature. The calculated d-spacing values of 3.3, 2.6 and 1.7 Å correspond to Miller indices (110), (101) and (211) of rutile SnO2 (cassiterite).

The XRD patterns of the three samples are shown in Fig. 2. XRD patterns of SnO2/rGO (Fig. 2a) and SnO2/rGO-120 (Fig. 2b) exhibit lower the intensity of diffraction peaks, indicating the poor crystallinity of SnO2. The strong diffraction peaks of SnO2/rGO-280 (Fig. 2c) can be indexed to a tetragonal rutile SnO2 (JCPDS no. 41-1445).


image file: c6ra23236a-f2.tif
Fig. 2 XRD patterns of (a) SnO2/rGO, (b) SnO2/rGO-120 and (c) SnO2/rGO-280, respectively.

To further verify the successful reduction of GO by Sn2+ at room temperature, the chemical composition and surface state of the three samples were discreetly investigated by XPS. The survey spectra of the three samples are shown in Fig. 3a, only Sn, O, and C elements were found. Sn 3d high-resolution spectra of the three samples are shown in Fig. 3b, the Sn 3d peaks of SnO2/rGO could be fitted with two peaks centered at 495.3 (Sn 3d3/2) and 486.9 eV (Sn 3d5/2). For the SnO2/rGO-120 and SnO2/rGO-280, the Sn 3d3/2/3d5/2 were 495.7/487.3 eV and 495.6/487.2 eV, respectively. It can be seen that the peak-to-peak separations between the Sn 3d3/2 and the Sn 3d5/2 are 8.4 eV, indicating the Sn4+ oxidation state in the three samples.15,37,38 Meanwhile, the peak shifting of Sn 3d should be attributed to the influences of different functional groups on the surface of rGO. The C 1s high resolution spectra of the three samples are shown in Fig. 3c, C 1s binding energies of SnO2/rGO appear at 283.8 (C–H),39 284.2 (C–C),5,40 284.6 (C[double bond, length as m-dash]C),41,42 285.1 (C–OH),5 286.2 eV (C[double bond, length as m-dash]O),43,44 and 288.7 eV (O–C[double bond, length as m-dash]O).29,39 After calcinating at 120 °C, the C–H peaks (283.8 eV) disappear, and the peak area ratio of C–OH, C[double bond, length as m-dash]O and O–C[double bond, length as m-dash]O are decreased. However, the peak area ratio of C–OH and O–C[double bond, length as m-dash]O is increased when the calcinating temperature reached to 280 °C. The O 1s peak of SnO2/rGO can be resolved into six peaks that correspond to Sn–OH, C–OH, Sn–O–Sn, Sn–O–C, Sn[double bond, length as m-dash]O and C[double bond, length as m-dash]O with binding energies of 530.4, 530.8, 531.3, 531.8, 532.3, and 533.1 eV, respectively.41,45–47 The Sn–OH peak (530.4 eV) disappeared after calcinating at 120 °C, but the peak area ratio of Sn–O–Sn, Sn–O–C, Sn[double bond, length as m-dash]O (SnO2/rGO-120 is 532.1 eV, SnO2/rGO-280 is 532.2 eV) are gradually increase with increasing calcinating temperature. It means that there are some Sn–O–C and Sn–OH bonds in SnO2 nanoparticles of SnO2/rGO, and they may also result in the poor crystallinity of SnO2 in three samples (Fig. 2). Meanwhile, it also indicates that SnO2 nanoparticles could be anchored on the surface of rGO via the Sn–O–C bond, and the amount of Sn–O–C bond is gradually increased according to the peak area ratio of Sn–O–C.


image file: c6ra23236a-f3.tif
Fig. 3 (a) XPS survey spectrum, (b) high resolution XPS spectra of Sn 3d, (c) C 1s, and (d) O 1s in the SnO2/rGO, SnO2/rGO-120, and SnO2/rGO-280, respectively.

TG-DTA analysis was performed to determine the appropriate processing temperature and the mass percentage of SnO2 nanoparticles in the SnO2/rGO composite. Fig. 4 shows the TG-DTA curves of SnO2/rGO in air atmosphere. The mass loss of about 7% before 120 °C is attributed to the removal of adsorbed water and crystal water in SnO2/rGO composite.4 The mass loss of about 12% from 120 to 500 °C is attributed to the oxidation of rGO.48 It also means that the mass percentage of SnO2 and rGO are about 81% and 12% in SnO2/rGO composite, respectively. After calcinating at 120 °C, the mass percentage of SnO2 could reach to about 87% [percentage = 81%/(81% + 12%)]. In addition, there are two platforms in DTA curve (blue line). One is from 226 to 263 °C, and the other is from 315 to 345 °C. We speculate that the first platform may correspond to the crystal phase transition of SnO2 or rGO, and the second platform may correspond to the oxidation of functional groups on the surface of rGO.


image file: c6ra23236a-f4.tif
Fig. 4 TG-DTA curves of SnO2/rGO composite in air atmosphere.

The changes in molecular structure can be revealed by the FT-IR spectra of SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280. As shown in Fig. 5, several absorption bands of oxygen functional groups are observed in the three samples, such as the absorption bands at about 3430 (O–H stretching vibrations), 1729 (C[double bond, length as m-dash]O stretching), 1587 (O–H bending), 1227 (C–O stretching), 1044 (C–O–C stretching), and 565 cm−1 (Sn–O–Sn stretching).1,29,49 It suggests that there are some oxygen functional groups in the three samples, they should come from GO. However, for the intensity of absorption bands at 1729, 1587, and 1227–1044 cm−1, the SnO2/rGO-280 significantly enhance compared with SnO2/rGO and SnO2/rGO-120. The result is consistent with the XPS and TG-DTA analysis.


image file: c6ra23236a-f5.tif
Fig. 5 FT-IR spectra of (a) SnO2/rGO, (b) SnO2/rGO-120 and (c) SnO2/rGO-280.

The crystallinity of SnO2 and crystallization quality of rGO have been investigated by Raman spectroscopy. As shown in Fig. 6, for all samples, two characteristic peaks can be observed at 1342 and 1601 cm−1, corresponding to the D and G band of rGO.5,43,47 The D band and G band in Raman spectra are related to the defects/disorder in the graphitic layers and stretching vibration of sp2 C in a graphene layer, respectively.50 The intensity ratios of D and G band (ID/IG) of three samples are 1.13, 1.20 and 1.04, respectively. It can be seen that the ID/IG of SnO2/rGO-120 is the biggest, which was suggestive of a partial reduction of GO in SnO2/rGO-120.5 With further increasing the calcinating temperature to 280 °C, the value of ID/IG decreases to 1.04, suggestive of further reduction of the rGO in SnO2/rGO-280.26 A possible reason for the higher ID/IG ratio of the graphene generated at 120 °C is that more oxygen-containing groups (e.g., C–OH, C[double bond, length as m-dash]O, and O–C[double bond, length as m-dash]O) decompose to CO2 at 120 °C, leaving more defects in the graphene.43,51 Meanwhile, the decomposing of oxygen-containing groups also generate more rGO in SnO2/rGO-120. This speculation can be confirmed by the XPS analysis of C and O (Fig. 3c and d). As shown in Fig. 3, the peak area ratio of C–OH (285.1 eV), C[double bond, length as m-dash]O (286.2 eV), and O–C[double bond, length as m-dash]O (288.7 eV) decrease (Fig. 3c) after calcinating at 120 °C, and the C–H peaks (283.8 eV) disappear. Moreover, the amount of C–O–C band increase (Fig. 3d, SnO2/rGO-120) in this process, which should arise from the esterification or condensation reaction of oxygen-containing groups on the surface of GO. However, the absorption bands intensity of C[double bond, length as m-dash]O and C–O bonds (Fig. 5c) significantly enhance at SnO2/rGO-280, which means that new and more oxygen-containing groups are again generated on the surface of rGO. It also suggests that the crystallization symmetrical hexagonal graphitic lattice may have been destroyed at 280 °C in air, so that crystallization quality of rGO decreases. Meanwhile, the SnO2/rGO-120 and SnO2/rGO-280 composites show a weak peak at ∼478 cm−1 corresponding to the Eg vibrational mode of rutile SnO2,52–56 and this peak intensity is enhanced gradually with increasing the calcinating temperature. It means that the crystallinity of SnO2 can be enhanced by increasing the calcinating temperature. This is also consistent with the XRD and XPS results. In addition, the peaks at ∼2674 and 2948 cm−1 are attributed to the 2D and D + G band of single-layer graphene, respectively.


image file: c6ra23236a-f6.tif
Fig. 6 Raman spectra of (a) SnO2/rGO, (b) SnO2/rGO-120, (c) SnO2/rGO-280.

The porous nature of SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280 was demonstrated by BET measurement. The N2 adsorption–desorption isotherms and the pore size distribution curves are shown in Fig. 7, and the results are listed in Table 1. The three samples show the typical IV adsorption isotherm (Fig. 7a), which indicates that all of samples contain mesoporous structure. The pore size distributes (Fig. 7b) demonstrate that the as-prepared samples have three kinds of pore structure in the range of 2–130 nm.


image file: c6ra23236a-f7.tif
Fig. 7 (a) N2 adsorption–desorption isotherms of three samples. (b) The pore size distribution calculated from the desorption branch.
Table 1 The specific surface area, pore volume, and average pore size of the products
Sample Specific surface area [m2 g−1] Pore volume [cm3 g−1] Average pore size [nm]
SnO2/rGO 179 0.34 7.6
SnO2/rGO-120 183 0.32 7.2
SnO2/rGO-280 142 0.23 6.5


The first is mesopore, the pore diameter centers are about 30, 37 and 18 nm in the SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280, respectively. The second is mesoporous with 3–5 nm, and the third is microporous with less than 2 nm. The mesopore mainly come from the cross-link between graphene layers. The 3D hierarchically porous structure was destroyed when the calcinating temperature reached to 280 °C (Fig. 1i), which result in the amount of mesopore is decreased (Fig. 7b). The mesoporous and microporous mainly come from the connection of SnO2 nanoparticles and the defects of graphene layers. As shown in Table 1, because the collapse of highly opened macroporous structure (Fig. 1), the pore volume and average pore sizes of as-prepared samples decrease gradually with increasing the calcinating temperature.

The distribution of C, Sn and O elemental can be further revealed by element mapping of composite. For example, the element mapping images of SnO2/rGO-120 are shown in Fig. S2, it is clear that C, Sn and O are evenly distributed onto the surface of SnO2/rGO-120. The C, O and Sn atomic percent in the SnO2/rGO composite is about 29.6%, 54.17% and 16.23% according to the elemental analysis, respectively.

All the results indicate that SnO2 nanoparticles could be anchored on graphene skeleton via Sn–O–C bonds. The crystallinity of SnO2 and crystallization qualities of rGO can be enhanced through calcinating at 120 °C in air atmosphere, and the amount of C–OH, C[double bond, length as m-dash]O and O–C[double bond, length as m-dash]O are decreased (Fig. 3c). But rGO can also be decomposed with further increasing calcinating temperature to 280 °C. Therefore, the fabrication process can be illustrated in Scheme 1. It is an advantage for the industrial production of SnO2/rGO composites as compared to a high-temperature annealing process in protective atmosphere, because it is not necessary for a relatively high equipment requirements and strict operation condition, and its cost is also relatively low at low temperature and air atmosphere.


image file: c6ra23236a-s1.tif
Scheme 1 Scheme illustration for the fabrication of SnO2/rGO composites at different calcinating temperature.

3.2. Electrochemical performance study

The electrochemical reactivity of the SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280 was further evaluated by cyclic voltammetry in lithium-ion half-cell. Fig. 8 displays the cyclic voltammetry curves (CVs) of three samples on the first 3 cycles, which were obtained between 0 and 3.0 V at a scanning a rates of 0.1 mV s−1. The CVs of SnO2/rGO are shown in Fig. 8a, in the first cycle, an irreversible broad peak at about 0.84 V is observed, which should be attributed to the formation of a solid electrolyte interphase and decomposition of SnO2 to form Sn.57–59 The subsequent CVs show that it (about 0.84 V) becomes reversible peak at about 0.94 V, indicating the formation of various LixSn species.57 During the anodic process, two oxidation peaks at around 0.64 and 1.34 V can be attributed to the dealloying of LixSn and the partial conversion of Sn to SnO2,11,60–62 respectively. The electrochemical process of SnO2/rGO electrode can be described by the following reactions:28,29,63
 
Li+ + e + electrolyte → SEI (1)
 
SnO2 + 4Li+ + 4e → 2Li2O + Sn (2)
 
Sn + xLi+ + xe ↔ LixSn (0 ≤ x ≤ 4.4) (3)
 
xLi+ + C (graphene) + xe ↔ LixC (4)

image file: c6ra23236a-f8.tif
Fig. 8 Cyclic voltammetry (CVs) of the different SnO2/rGO electrodes at a scan rate of 0.1 mV s−1. (a) SnO2/rGO, (b) SnO2/rGO-120, (c) SnO2/rGO-280.

As shown in Fig. 8a, the peaks of ∼0.01 V can be ascribed to Li+ intercalation into graphene to form LiC6 (eqn (4)).64,65 According to the previous reports, the formation of LixSn (eqn (3)) should form a peak at about 0.25 V,65 but we can not find this peak in Fig. 8a. It means that the formation of alloy between Sn and Li was very weak owing to the poor crystallinity of SnO2. As shown in Fig. 8b, the broad peak of SnO2/rGO-120 has almost no change at about 0.84 V on the first 3 cycles. It suggests that the decomposition of SnO2 is a stable process in the SnO2/rGO-120 electrode. Meanwhile, the irreversible reduction peak at about 0.19 V was divided into two reversible peaks at 0.06 and 0.26 V, corresponding to Li intercalation into graphene (eqn (4)) and the formation of alloy between Sn and Li (eqn (3)).64,65 Furthermore, the subsequent CVs show good reproducibility with several cathodic and anodic peak pairs, suggesting a very high degree of reversibility of the multistep conversion. It also suggests that both rGO and SnO2 play the activity materials role in the SnO2/rGO-120.

For the SnO2/rGO-280 electrode (Fig. 8c), the peak position drift to about 0.90 V, and the intensity increase significantly in the subsequent CVs, it should be due to that more SnO2 are decomposed into Sn (eqn (2)). Meanwhile, the reduction peak at about 0.19 V is very stable in the SnO2/rGO-280 electrode. It suggests that the reaction of Li intercalation into graphene is a very weak process in the SnO2/rGO-280 electrode. All these CVs results indicate that the intercalation object of Li+ is different in the SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280. It should come from the change of SnO2 crystallinity and the amount of rGO.

The Nyquist plots of SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280 are shown in Fig. 9a. After fitting the Nyquist plots with the typical circuit model (inset of Fig. 9a),66–68 the resistance parameters are listed in Table 2. The transfer resistances (R3) of three samples are 490, 90.1, and 3.2 × 10−10 Ω, respectively, suggesting that the electron transference for the latter is prior to that of the former.69 Especially SnO2/rGO-280, the electron transference has almost no resistance. According to the XPS (Fig. 3d) and Raman spectroscopy (Fig. 6), the crystallinity of SnO2 and the amount of Sn–O–C bonds are the highest in SnO2/rGO-280. These factors should be advantage for enhancing the electron transference of composites, so that R3 of SnO2/rGO-280 is very small. When the crystallization quality of rGO is also considered, we speculate that the crystallization quality and the amount of rGO could be small in the as-prepared SnO2/rGO-280 sample (Fig. 6), so the reaction of Li intercalation into graphene is a very weak process in the SnO2/rGO-280 electrode. Therefore, the cycling capacities of SnO2/rGO-280 should be the smallest in the three samples because only SnO2 plays the active materials role in the charge–discharge process.


image file: c6ra23236a-f9.tif
Fig. 9 (a) Nyquist plots of the different SnO2/rGO electrodes. The inset shows equivalent circuit for the impedance spectrum, the solid lines are the fitted curves. (b) Charge–discharge profiles of the different SnO2/rGO electrodes at a current rate of 0.2 A g−1 for the first cycle. (c) Cycling performance of the different SnO2/rGO electrodes at 0.5 A g−1. (I)–(III) are corresponding to the SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280 in (a–c). Cycling performances of SnO2/rGO (d), SnO2/rGO-120 (e) and SnO2/rGO-280 (f) at the different current densities. All of the measurements were conducted using a voltage window of 0.01–3.0 V.
Table 2 The circuit used to fit the EIS spectra and the fitted values for the corresponding components in the equivalent circuit
Samples R1 R2 R3 Rcell
SnO2/rGO 19.5 437.1 490 946.6
SnO2/rGO-120 10.5 128.2 90.1 228.8
SnO2/rGO-280 13.3 53.4 3.2 × 10−10 66.7


The first discharge–charge curves of SnO2/rGO, SnO2/rGO-120 and SnO2/rGO-280 are shown in Fig. 9b. For the three samples, the initial discharge capacity could reach to about 2046, 1947, and 1596 mA h g−1 at current densities of 0.2 A g−1, and the first charge capacity is about 1258, 1166, and 900 mA h g−1, respectively. Thus their coulombic efficiency is 61%, 60%, and 56%, respectively. As shown in Fig. 9c–f, the charge–discharge capacities of the SnO2/rGO and SnO2/rGO-120 in the first 3 cycles is unstable at the different current densities. On the contrary, the SnO2/rGO-280 is stable. In addition, the cycling performances of SnO2/rGO-120 is the best in the three samples, such as at current densities of 0.5 A g−1, the initial discharge capacity of three samples are 1194, 1610 and 1364 mA h g−1, respectively. After 100 cycles, the specific discharge capacity of three samples are 250, 429 and 172 mA h g−1, respectively. According to the structural characterization (Section 3.1), CVs (Fig. 8) and impedance analysis (Fig. 9a), good cycling performance and high rate capability of SnO2/rGO-120 should be attributed to the appropriate calcinating temperature. On the one hand, it provides a novel approach to modify SnO2/rGO composites. On the other hand, it decreases the resistance of SnO2/rGO-120 electrode. Meanwhile, we note that the SnO2/rGO is dried in the drying oven at 120 °C (electrode), and the SnO2/rGO-120 is calcined in the muffle furnace, but they are difference in the electrochemical properties. It means that the effect of heat treatment mode is important for the rGO composites. However, many studies have ignored this in previous reports.

Fig. 10 shows the HRTEM images of three samples and the grain size distributions of SnO2 nanoparticles after 100 cycles at current densities of 0.2 A g−1. The HRTEM images (Fig. 10a–c) of three samples show that the dispersion of SnO2 nanoparticles is still good. The grain size distributions of SnO2 nanoparticles (Fig. 10d–f) show that the average sizes of SnO2 nanoparticles are 4.9 ± 0.5, 3.2 ± 0.5 and 4.0 ± 0.5 nm in three samples, respectively. The volume expansion of SnO2 is significant for SnO2/rGO during the charge/discharge process. The reason is that the SnO2 crystallinity is poor in SnO2/rGO. On the contrary, with increasing the SnO2 crystallinity, the volume expansion of SnO2 can be prevented. The previous reports have indicated that the SnO2 nanoparticle size is one of the key factors for the stable cycling performance of SnO2, where smaller particle size could help to prevent gradual aggregation of Sn into large clusters.70 In this work, the SnO2 nanoparticles size of 3.2 nm demonstrates that it also help to prevent the volume expansion of SnO2 during the charge/discharge process.


image file: c6ra23236a-f10.tif
Fig. 10 The HRTEM images and the grain size distributions of SnO2 nanoparticles in the three samples. (a, d) SnO2/rGO, (b, e) SnO2/rGO-120, (c, f) SnO2/rGO-280. The samples were obtained after 100 discharge–charge cycles at a current rate of 0.2 A g−1.

4. Conclusion

In summary, three 3D hierarchically porous SnO2/rGO composite was synthesized through stannous ions reduced GO. A facile post-process approach can be used to improve the crystallinity of SnO2 and crystallization quality of rGO in SnO2/rGO composites, and SnO2 NPS can be anchored on the surface of rGO substrate by Sn–O–C bond. The electrochemical performance results indicate that the crystallinity of SnO2 and the crystallization quality of rGO can effect on the intercalation object of Li+, conductivity of composites, cycling performances and volume expansion of SnO2, etc. Since this post-process approach can be performed at low temperature in air atmosphere rather than high temperature in protecting atmosphere, it is beneficial for the synthesis of large quantities of SnO2/rGO composites and improving the electrochemical properties of SnO2/rGO composites in LIBs applications.

Acknowledgements

This work was financially supported by Science and Technology Research Project of Henan Province (No. 152102310311, 152102210083 and 142102210455), the Key Scientific Research Project of Colleges and Universities in Henan (No. 16A150032, 15A150057) and Henan Normal University Innovation Funds for Postgraduate (No. YL201407, No. YL201512). We would also like to thank for the High Performance Computing Center of Henan Normal University, Engineering Technology Research Center of Motive Power and Key Materials for providing a LAND CT2001A battery-testing instrument. F. T. and X. W. contributed equally to this work.

Notes and references

  1. M. J. Reddy, S. H. Ryu and A. M. Shanmugharaj, Nanoscale, 2016, 8, 471–482 RSC.
  2. Z. Li, G. Wu, S. Deng, S. Wang, Y. Wang, J. Zhou, S. Liu, W. Wu and M. Wu, Chem.–Eng. J., 2016, 283, 1435–1442 CrossRef CAS.
  3. 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.
  4. R. Liu, Y. Liu, Q. Kang, A. Casimir, H. Zhang, N. Li, Z. Huang, Y. Li, X. Lin, X. Feng, Y. Ma and G. Wu, RSC Adv., 2016, 6, 9402–9410 RSC.
  5. C. Gong, Y. Zhang, M. Yao, Y. Wei, Q. Li, B. Liu, R. Liu, Z. Yao, T. Cui, B. Zou and B. Liu, RSC Adv., 2015, 5, 39746–39751 RSC.
  6. J. Tang, J. Yang, X. Zhou, H. Yao and L. Zhou, J. Mater. Chem. A, 2015, 3, 23844–23851 CAS.
  7. X. Li, X. Meng, J. Liu, D. Geng, Y. Zhang, M. N. Banis, Y. Li, J. Yang, R. Li, X. Sun, M. Cai and M. W. Verbrugge, Adv. Funct. Mater., 2012, 22, 1647–1654 CrossRef CAS.
  8. W. Zhou, J. Wang, F. Zhang, S. Liu, J. Wang, D. Yin and L. Wang, Chem. Commun., 2015, 51, 3660–3662 RSC.
  9. X. Huang, Z. Zeng, Z. Fan, J. Liu and H. Zhang, Adv. Mater., 2012, 24, 5979–6004 CrossRef CAS PubMed.
  10. X. W. Lou, C. M. Li and L. A. Archer, Adv. Mater., 2009, 21, 2536–2539 CrossRef CAS.
  11. Z. Zhang, L. Wang, J. Xiao, F. Xiao and S. Wang, ACS Appl. Mater. Interfaces, 2015, 7, 17963–17968 CAS.
  12. J. S. Chen and X. W. Lou, Small, 2013, 9, 1877–1893 CrossRef CAS PubMed.
  13. P. Wu, N. Du, H. Zhang, J. Yu, Y. Qi and D. Yang, Nanoscale, 2011, 3, 746–750 RSC.
  14. Z. Wang, D. Luan, F. Y. Boey and X. W. Lou, J. Am. Chem. Soc., 2011, 133, 4738–4741 CrossRef CAS PubMed.
  15. Y. Zhu, H. Guo, H. Zhai and C. Cao, ACS Appl. Mater. Interfaces, 2015, 7, 2745–2753 CAS.
  16. C. Wei, G. Zhang, Y. Bai, D. Yan, C. Yu, N. Wan and W. Zhang, Solid State Ionics, 2015, 272, 133–137 CrossRef CAS.
  17. H. Song, X. Li, Y. Cui, D. Xiong, Y. Wang, J. Zeng, L. Dong, D. Li and X. Sun, Int. J. Hydrogen Energy, 2015, 40, 14314–14321 CrossRef CAS.
  18. G. Zhang, J. Zhu, W. Zeng, S. Hou, F. Gong, F. Li, C. C. Li and H. Duan, Nano Energy, 2014, 9, 61–70 CrossRef CAS.
  19. W.-S. Kim, Y. Hwa, J.-H. Jeun, H.-J. Sohn and S.-H. Hong, J. Power Sources, 2013, 225, 108–112 CrossRef CAS.
  20. 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.
  21. M. Wu, J. Liu, M. Tan, Z. Li, W. Wu, Y. Li, H. Wang, J. Zheng and J. Qiu, RSC Adv., 2014, 4, 25189–25194 RSC.
  22. Y. Wang, H. C. Zeng and J. Y. Lee, Adv. Mater., 2006, 18, 645–649 CrossRef CAS.
  23. L. Wang, D. Wang, Z. Dong, F. Zhang and J. Jin, Nano Lett., 2013, 13, 1711–1716 CrossRef CAS PubMed.
  24. S. M. Paek, E. Yoo and I. Honma, Nano Lett., 2009, 9, 72–75 CrossRef CAS PubMed.
  25. Y. Deng, C. Fang and G. Chen, J. Power Sources, 2016, 304, 81–101 CrossRef CAS.
  26. Z. Song, Z. Wei, B. Wang, Z. Luo, S. Xu, W. Zhang, H. Yu, M. Li, Z. Huang, J. Zang, F. Yi and H. Liu, Chem. Mater., 2016, 28, 1205–1212 CrossRef CAS.
  27. Z. Li, G. Wu, D. Liu, W. Wu, B. Jiang, J. Zheng, Y. Li, J. Li and M. Wu, J. Mater. Chem. A, 2014, 2, 7471–7477 CAS.
  28. D. Wang, X. Li, J. Wang, J. Yang, D. Geng, R. Li, M. Cai, T.-K. Sham and X. Sun, J. Phys. Chem. C, 2012, 116, 22149–22156 CAS.
  29. W. Li, D. Yoon, J. Hwang, W. Chang and J. Kim, J. Power Sources, 2015, 293, 1024–1031 CrossRef CAS.
  30. Y. Jiang, T. Yuan, W. Sun and M. Yan, ACS Appl. Mater. Interfaces, 2012, 4, 6216–6220 CAS.
  31. J. Lin, Z. Peng, C. Xiang, G. Ruan, Z. Yan, D. Natelson and J. M. Tour, ACS Nano, 2013, 7, 6001–6006 CrossRef CAS PubMed.
  32. F. Yan, X. Tang, Y. Wei, L. Chen, G. Cao, M. Zhang and T. Wang, J. Mater. Chem. A, 2015, 3, 12672–12679 CAS.
  33. R. Tian, Y. Zhang, Z. Chen, H. Duan, B. Xu, Y. Guo, H. Kang, H. Li and H. Liu, Sci. Rep., 2016, 6, 19195 CrossRef CAS PubMed.
  34. X. Liu, J. Cheng, W. Li, X. Zhong, Z. Yang, L. Gu and Y. Yu, Nanoscale, 2014, 6, 7817–7822 RSC.
  35. W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS.
  36. C. Botas, D. Carriazo, G. Singh and T. Rojo, J. Mater. Chem. A, 2015, 3, 13402–13410 CAS.
  37. X. Wang, Y. Wang, F. Tian, H. Liang, K. Wang, X. Zhao, Z. Lu, K. Jiang, L. Yang and X. Lou, J. Phys. Chem. C, 2015, 119, 15963–15976 CAS.
  38. S. J. Choi, B. H. Jang, S. J. Lee, B. K. Min, A. Rothschild and I. D. Kim, ACS Appl. Mater. Interfaces, 2014, 6, 2588–2597 CAS.
  39. M. Liu, X. Wang, Y. Chen and L. Dai, RSC Adv., 2015, 5, 61481–61485 RSC.
  40. Y. Zhu, C. Li and C. Cao, RSC Adv., 2013, 3, 11860 RSC.
  41. X. Xie, D. Su, J. Zhang, S. Chen, A. K. Mondal and G. Wang, Nanoscale, 2015, 7, 3164–3172 RSC.
  42. S. Chen, Y. Qiao, J. Huang, H. Yao, Y. Zhang, Y. Li, J. Du and W. Fan, RSC Adv., 2016, 6, 25198–25202 RSC.
  43. Y. Hong, J. Zhang, Z. Wang, J. J. Stankovich and X. Jin, RSC Adv., 2014, 4, 64402–64409 RSC.
  44. R. Shen, Y. Hong, J. J. Stankovich, Z. Wang, S. Dai and X. Jin, J. Mater. Chem. A, 2015, 3, 17635–17643 CAS.
  45. S. Singkammo, A. Wisitsoraat, C. Sriprachuabwong, A. Tuantranont, S. Phanichphant and C. Liewhiran, ACS Appl. Mater. Interfaces, 2015, 7, 3077–3092 CAS.
  46. A. Bhaskar, M. Deepa, M. Ramakrishna and T. N. Rao, J. Phys. Chem. C, 2014, 118, 7296–7306 CAS.
  47. Y. G. Zhu, Y. Wang, J. Xie, G.-S. Cao, T.-J. Zhu, X. Zhao and H. Y. Yang, Electrochim. Acta, 2015, 154, 338–344 CrossRef CAS.
  48. L. Yin, D. Chen, X. Cui, L. Ge, J. Yang, L. Yu, B. Zhang, R. Zhang and G. Shao, Nanoscale, 2014, 6, 13690–13700 RSC.
  49. R. Huang, L. Wang, Q. Zhang, Z. Chen, Z. Li, D. Pan, B. Zhao, M. Wu, C. M. L. Wu and C.-H. Shek, ACS Nano, 2015, 11, 11351–11361 CrossRef PubMed.
  50. S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff, Carbon, 2007, 45, 1558–1565 CrossRef CAS.
  51. X. Yu, Y. Kang and H. S. Park, Carbon, 2016, 101, 49–56 CrossRef CAS.
  52. G. Qin, F. Gao, Q. Jiang, Y. Li, Y. Liu, L. Luo, K. Zhao and H. Zhao, Phys. Chem. Chem. Phys., 2016, 18, 5537–5549 RSC.
  53. B. Zhao, B. Fan, Y. Xu, G. Shao, X. Wang, W. Zhao and R. Zhang, ACS Appl. Mater. Interfaces, 2015, 7, 26217–26225 CAS.
  54. N. Bhardwaj, A. Pandey and S. Mohapatra, J. Alloys Compd., 2016, 656, 647–653 CrossRef CAS.
  55. J. Wei, S. Xue, P. Xie and R. Zou, Mater. Lett., 2015, 159, 489–492 CrossRef CAS.
  56. Y. M. Lu, J. Jiang, C. Xia, B. Kramm, A. Polity, Y. B. He, P. J. Klar and B. K. Meyer, Thin Solid Films, 2015, 594, 270–276 CrossRef CAS.
  57. J. Lin, Z. Peng, C. Xiang, G. Ruan, Z. Yan, D. Natelson and J. M. Tour, ACS Nano, 2013, 7, 6001–6006 CrossRef CAS PubMed.
  58. M. Mohamedi, S.-J. Lee, D. Takahashi, M. Nishizawa, T. Itoh and I. Uchida, Electrochim. Acta, 2001, 46, 1161–1168 CrossRef CAS.
  59. L. Noerochim, J.-Z. Wang, S.-L. Chou, H.-J. Li and H.-K. Liu, Electrochim. Acta, 2010, 56, 314–320 CrossRef CAS.
  60. J. Sun, L. Xiao, S. Jiang, G. Li, Y. Huang and J. Geng, Chem. Mater., 2015, 27, 4594–4603 CrossRef CAS.
  61. F. Han, W. C. Li, M. R. Li and A. H. Lu, J. Mater. Chem., 2012, 22, 9645–9651 RSC.
  62. L. Noerochim, J.-Z. Wang, S.-L. Chou, D. Wexler and H.-K. Liu, Carbon, 2012, 50, 1289–1297 CrossRef CAS.
  63. Y. Zhang, L. Jiang and C. Wang, Phys. Chem. Chem. Phys., 2015, 17, 20061–20065 RSC.
  64. X. Li, Y. Zhang, T. Li, Q. Zhong, H. Li and J. Huang, Electrochim. Acta, 2014, 147, 40–46 CrossRef CAS.
  65. H. Liu, J. Huang, X. Li, J. Liu, Y. Zhang and K. Du, Appl. Surf. Sci., 2012, 258, 4917–4921 CrossRef CAS.
  66. T. Xia, W. Zhang, Z. H. Wang, Y. L. Zhang, X. Y. Song, J. Murowchick, V. Battaglia, G. Liu and X. B. Chen, Nano Energy, 2014, 6, 109–118 CrossRef CAS.
  67. H. Han, T. Song, E.-K. Lee, A. Devadoss, Y. Jeon, J. Ha, Y.-C. Chung, Y.-M. Choi, Y.-G. Jung and U. Paik, ACS Nano, 2012, 6, 8308–8315 CrossRef CAS PubMed.
  68. J. Y. Liao, D. Higgins, G. Lui, V. Chabot, X. Xiao and Z. Chen, Nano Lett., 2013, 13, 5467–5473 CrossRef CAS PubMed.
  69. L. Zeng, C. Zheng, L. Xia, Y. Wang and M. Wei, J. Mater. Chem. A, 2013, 1, 4293–4299 CAS.
  70. Y. Chen, B. Song, R. M. Chen, L. Lu and J. Xue, J. Mater. Chem. A, 2014, 2, 5688–5695 CAS.

Footnote

Electronic supplementary information (ESI) available: The grain size distributions of SnO2 nanoparticles, elemental mapping images. See DOI: 10.1039/c6ra23236a

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