Skip to main content

Capacity Contribution Mechanism of rGO for SnO2/rGO Composite as Anode of Lithium-ion Batteries

Abstract

Compared with ordinary graphite anode, SnO2 possesses higher theoretical specific capacity, rich raw materials and low price. While the severe volume expansion of SnO2 during lithium-ion extraction/intercalation limits its further application. To solve this problem, in this work the reduced graphene oxide (rGO) was introduced as volume buffer matrix of SnO2. Herein, SnO2/rGO composite is obtained through one-step hydrothermal method. Three-dimensional structure of rGO could effectively hinder the polymerization of SnO2 nanoparticles and provide more lithium storage sites attributed to high specific surface area and density defects. The initial discharge capacity of the composite cathode is 959 mA·h·g-1 and the capacity remained at 300 mA·h·g-1 after 1000 cycles at 1 C. It proved that the rGO added in the anode has a capacity contribution to the lithium-ion battery. It changes the capacity contribution mechanism from diffusion process dominance to surface driven capacitive contribution. Due to the addition of rGO, the anode material gains stable structure and great conductivity.

1 Introduction

Lithium-ion batteries have realized large-scale application in electric and hybrid vehicles. In the past few decades, SnO2 has attracted extensive attention as electrode material. Compared with ordinary graphite anode materials, SnO2 possesses higher theoretical specific capacity (790 mA·h·g-1), rich raw materials and low price. While the volume expansion of SnO2 during lithium-ion extraction/intercalation reaches more than 50%, the high stress caused by volume expansion may lead to the fracture failure of electrode material which makes the specific capacity of the material decrease rapidly, while the rate performance and cycle stability decreasing [1,2,3]. Therefore, the key to realize the application of SnO2 cathode materials in lithium-ion batteries is to find ways to effectively slow down the volume dilatation effect of SnO2 during the lithium-ion extraction/intercalation process and upgrade the electrochemical performance of electrode materials [4,5,6,7]. Graphene is a 2D material with good conductivity and high specific capacity. Graphene nanosheets can not only effectually prevent the volumetric change and particle aggregation of SnO2, but also improve the conductivity. As an active material for lithium storage, graphene nanosheets keep the structural integrity of electrode [8,9,10]. Chen et al. synthesize a reduced graphene and SnO2 nanospheres composite and obtain good performance in the application of lithium-ion batteries or sodium-ion batteries [11]. Zhou et al. built a SnO2/GO structure by hydrothermal method [12]. It’s worth noting that they researched the influence of GO dosage in SnO2/GO anode and found that the lithium ions storage capability increases with the raise of GO dosage. Lu et al. focused on the drying process after hydrothermal process [13]. By comparing anodes treated by spray-drying and freeze-drying, novel composition of SnO2 and GO achieved after spray-drying process obtained favorable lithium-ion transmission during charging/discharging process.

The combination of rGO and anode materials of lithium-ion batteries or sodium-ion batteries can increase the capacitance contribution, which has been confirmed by much research [14,15,16,17]. While, for SnO2, the influence of rGO on capacitance contribution rate has not been studied. In this work, we report to synthesize SnO2/rGO composite by the sample single step method. We analyze the lithiation/delithiation process and find that the addition of rGO will significantly improve the mobility of lithium ions. When the scan rate comes to 1 mV·s-1, the capacitive contribution rate raises from 56% to 71% after the addition of rGO. It confirms that the addition of rGO changes the capacity contribution mechanism from diffusion process dominance to surface driven capacitive contribution dominance attributed to high specific surface area, abundant defects and high conductivity of rGO.

2 Materials and Methods

2.1 Preparation of SnO2/rGO Composite

The GO dispersion was prepared by the oxidation of graphite powder in acidic environment according to a modified Hummers method [18,19,20]. The as-obtained aqueous suspension of GO was dispersed under ultrasonication for 6 h and purified with deionized water. In a typical synthesis, 0.904 g (4 mmol) SnCl2 2H2O, 2.3528 g (8 mmol) sodium citrate, 10 ml ethanol and 15 ml GO dispersion (2 mg·mL-1) were mixed together to form a uniform solution on a magnetic stirrer. During magnetic stirring, the mixture will be ultrasonically treated for a period of time in order to make the solute fully dispersed. Then the mixture was transferred to a 30 ml reaction kettle for hydrothermal treatment at 180 ℃ for 6 h. Thereafter, the cooled down product was washed by continuous centrifugation method. Finally, the as-obtained product was dried at 60 ℃ for the whole day, the ratio of rGO to SnO2 is about 4:1. The synthesizing procedure of SnO2 nanoparticles is just in common with that of SnO2/rGO in which GO dispersion was replaced by equal amount of deionized water.

2.2 Electrochemical Measurement

CR2025 coin cells were assembled for the following electrochemical investigation. In order to prepare the anode, active materials, conductive carbon black and polyvinylidene difluoride were mixed together to form slurry by 8:1:1. The active material loaded on the Cu foil electrode was about 1.6 mg cm-2. 1 M LiPF6 in ethylene carbonate was added to dimethyl carbonate (1:1, V/V) as the electrolyte during the battery assembly process. Constant current charge and discharge test at various rates were performed by LAND CT2001A battery testing system within a voltage range of 0.001–3 V and the electrochemical characterizations were performed by electrochemical workstation (CHI660C, Shanghai).

2.3 Materials Characterization

The microstructure and composition were analyzed by X-ray diffraction (XRD, Rigaku D/max 2500 pc X-ray diffractometer), Fourier transform infrared spectroscopy (FTIR, Nicolet-6700 Thermofisher), Raman spectra (HORIBA Jobin Yvon LabRAM HR800) and X-ray photoelectron spectroscopy (XPS, Thermo Scientific, Waltham, MA, USA). The microstructure was explored by scanning electron microscopy (SEM, JEOL JSM-6700F) and transmission electron microscopy (TEM, FEI, Talos F200). The specific surface area was tested by N2 adsorption/desorption (NOVA 2000, Quanta chrome).

3 Results and Discussion

For the sake of analyzing the morphology and nanostructure of hydrothermal product, SEM test was carried out on SnO2 and SnO2/rGO powders. Figure 1a shows the pure SnO2 obtained by hydrothermal treatment presents spherical morphology with a diameter of around 1 to 3 microns. Figure 1b illustrates that SnO2 spheres (with a diameter of approximately 3 μm) are composed of several spherical nanoflowers with a diameter of about 200 nm. As for SnO2/rGO (Figure 1c and d), spherical SnO2 are equably absorbed on the surface of rGO nanosheets, meanwhile the SnO2 spherical structure remains unchanged. It’s obviously that SnO2 nanoflowers grown on rGO nanosheets, while the size is obviously smaller than that of SnO2 powders which is about 100 nm. The reduction of the diameter of the nanosphere means the increase of the specific surface area.

Figure 1
figure 1

a Low-magnification and b high-magnification of SEM images of SnO2 nanoparticles, c Low-magnification and d high-magnification of SEM images of SnO2/rGO composite, e TEM and f HRTEM images of SnO2/rGO composite

TEM spectra (Figure 1e, f) further illustrate the morphology and nanostructure of as prepared materials. It is highly consisted with the results obtained in previous SEM test. SnO2 nanoparticles are absorbed onto the rGO nanosheets with favorable connection. As for high-resolution TEM spectra (Figure 1f), SnO2 crystal grains are relatively visible and the rGO nanosheets could be obviously observed and the lattice spacing of 0.33 nm, corresponding to the (110) planes of tetragonal SnO2. X-ray diffraction (XRD) is tested for the component analysis of SnO2/rGO and SnO2 powder (Figure 2a). As for the XRD spectroscopy of SnO2/rGO, diffraction peaks in the curve could correspond to that of tetragonal tin dioxide with cassiterite structure (JCPDS Card No.41e1445). While, it’s of particular interest that, the XRD spectroscopy of SnO2 failed to correspond the standard peak one-to-one. This is because the crystallinity of SnO2 obtained after 6 h of hydrothermal reaction is not high enough, which shifts the peak of 33.8° to the left [22]. Since SnO2/rGO failed to detect the diffraction peak of rGO, the XRD spectrum results could not prove the existence of rGO.

Figure 2
figure 2

a XRD patterns and b Raman spectra and c FT-IR spectra of SnO2 and SnO2/rGO composite, d Nitrogen sorption isotherms of SnO2 particles, SnO2 nanoflower and SnO2/rGO composite

For the common homonuclear diatomic pairs in carbon-based materials, their Raman activity will be strong, so their Raman peaks can be easily detected in Raman spectra, thus the Raman spectroscopy was used for further composition analysis (Figure 2b). There are two obvious peaks appearing in ~ 1350 cm-1 and ~ 1580 cm-1 corresponding to the D and G bands of graphene respectively. The ID/IG significantly increases from 0.98 of GO to 1.76 of SnO2/rGO, which indicates that the sp2 domain of SnO2/rGO composites is smaller. Meanwhile it illustrates that oxygen content of GO reduced during hydrothermal process [21].

In order to further analyze the components and verify the partial reduction of GO during the hydrothermal process. FT-IR spectra of as-prepared nanocomposites are plotted (Figure 2c). Both spectra show the peak of 3440 cm-1 and 1629 cm-1 which correspond to O-H stretching vibrations of absorbed water molecules and O-H bending vibrations, the absorption band of around 1220 cm-1 corresponding to C-O. After compounding with SnO2, similar oxygen-containing groups are found in the SnO2/rGO spectra, which illustrates that GO is partially reduced in hydrothermal treatment, and a strong peak at 669 cm-1 is attributed to O-Sn-O stretching mode [22, 23]. N2 adsorption/desorption test is carried out to analyze the specific surface area of SnO2 particles (purchase from enterprise), SnO2 nanoflower and SnO2/rGO composite in Figure 2d. The BET specific surface area of SnO2 particles is only 23.16 cm2·g-1, while it comes to 63.17 cm2·g-1 when the morphology of SnO2 turns into spherical nanoflower. As for SnO2/rGO composite, it reaches 126.74 cm2·g-1, which is much higher than that of SnO2. It illustrates that the assistance of rGO could significantly increase the specific surface area [13]. The larger specific surface area means that SnO2/rGO can have better contact with electrolyte and absorb more lithium ions at the same time.

To obtain the exact composition of SnO2/rGO powders, we tested it with X-ray photoelectron spectroscopy (XPS), since XPS could display the chemical environment of Sn, O and C elements. Figure 3a displays the XPS survey spectra of SnO2/rGO powders. In C1s fitting patterns (Figure 3b), there are C-C bond at 284.7 eV, C-O bond at 285.6 eV and C=O bond at 286.8 eV, in which C-C bond takes an absolute dominant position. It corresponds to a common rGO chemical bond composition as mentioned in previous study [24]. In O1s spectroscopy (Figure 3c), Sn-O bond at 531.6 eV and C-O-Sn bond at 532.8 eV make different contribution in fitting patterns. The obvious C-O-Sn bond fitting peak indicates that there exists surpassing chemical bonding among SnO2 and rGO [25]. In Sn3d spectrum (Figure 3d), there are two obvious peaks located at 495.9 eV and 487.6 eV, which correspond to Sn3d 3/2 and Sn3d 5/2 respectively [26]. The peaks generally consist with that reported in pure SnO2 powder; it illustrates that major constituent of SnO2/rGO powders is SnO2.

Figure 3
figure 3

a XPS Survey spectrum of SnO2/rGO powders; High-resolution XPS spectra of SnO2/rGO composite b C1s; c O1s and d Sn3d

For the sake of researching the influence on lithium-ion battery performance after compounding with rGO, the electrochemical performance tests were carried out. Figure 4a display the initial three cycle cyclic voltammetry (CV) profiles of the electrode which chose SnO2/rGO composites as active materials in the voltage range from 0.01 to 3.0 V. In the first cycle, the cathodic peak at around 0.78 V could be observed while it disappeared in the subsequent cycles. This situation occurs due to the irreversible reduction of SnO2 and the formation of solid electrolyte interface (SEI), as described in Eq. (1) [27]. Furthermore, it will lead to significant capacity loss in the first cycle. An obvious reduction peak appeared at the range of 0.08 and 0.13 V, which correspond to the alloying reaction of Sn and Li (Eq. (2)) [28]. Meanwhile it is relevant to the intercalation of lithium [29]. As for the following two cycles, the obvious cathodic peaks at around 1.2 V correspond to the transformation between SnO2 and Sn [30]. In the charge curve, the anodic peak at around 0.57 V correspond to the reverse process of the alloying reaction mentioned before, which means the decomposition of LixSn [31]. There were two weak oxidation peaks at 1.27 and 1.9 V, which corresponded to reversible reaction of SnO2 and Li2O. In comparison of the CV curves for the first three cycles, the crest value of alloying/dealloying reaction showed slight decline as cycle number increased, it indicated that SnO2 nanoparticles brought out further activation during the test period. The CV curves of later cycles had significant overlap ratio, which indicated that SnO2/rGO has excellent reversibility [28].

Figure 4
figure 4

a CV curves of SnO2/rGO composite at a scan rate of 0.1 mV·s-1, b Discharge/charge profiles of SnO2/rGO composite at different cycles with current rate of 0.1 C, c Cycling performance and coulombic efficiency of SnO2/rGO composite and SnO2 anode at current rate of 0.1 C, d Rate capability of SnO2/rGO composite and SnO2 at gradient current rate, e Long cycle stability of SnO2/rGO anode

$$\text{SnO}_{2}+4\text{Li}+4\text{e}^{-}\to \text{Sn}+2\text{Li}_{2}\text{O},$$
(1)
$$\text{Sn}+x\text{Li}^{+}+x\text{e}^{-}\to \text{Li}_{x}\text{Sn}(0\le x\le 4.4).$$
(2)

When the specific current reaches 0.1 C, the galvanostatic discharge-charge curves of 1st, 2nd, 5th and 10th for the SnO2/rGO are shown in Figure 4b. In the first cycle, the discharge and charge capacity of the composite are 1114.3 mA·h·g-1 and 731.9 mA·h·g-1, respectively. Compare our study with other SnO2/rGO materials as the anode for Li ion batteries, the capacitance performance of SnO2/rGO at 0.1C (1114.3 mA·h·g−1) is superior to SnO2 NPs/rGO (1100 mA·h·g−1) [11], CGN/SnO2 composites (1048.7 mA·h·g−1) [32] and SnO2-RGO (665 mA·h·g−1) [33]. A voltage plateau appears at approximately 1 V in the first cycle, while it can’t be observed in the subsequent cycles. It mainly due to irreversible reduction of SnO2 and electrolyte decomposition. In the following profiles, the voltage plateaus can be observed in the ranges of 0.01–0.5 V and 0.5–1 V corresponding to combination reaction of Li-Sn alloy. These results are highly consistent with the verdicts obtained in previous CV tests.

The cyclic performance and coulombic efficiency of two materials under current density of 0.1 C is displayed in Figure 4c. The initial cycle of SnO2/rGO galvanostatic discharge-charge give out 959.2 mA·h·g-1 for discharge capacity. In contrast, SnO2 delivers a discharge capacity of 754.6 mA·h·g-1 with coulombic efficiency of 59.04%. It’s obviously that, SnO2/rGO composite possess much higher discharge capacity and coulombic efficiency than that of SnO2 nanoparticles. Besides, the decline of discharge capacity with the increase of cycle number for SnO2/rGO composite is significantly faster than that of SnO2 nanoparticles. When the cycle number comes to 100, SnO2/rGO remains discharge capacity of 469.6 mA·h·g-1, and SnO2 nanoparticles reach to 121.1 mA·h·g-1. The above phenomenon is mainly caused by the following two reasons. Firstly, the SnO2/rGO composite possess much smaller grain size, which will lead to more active materials participating in the reaction during charge and discharge. Secondly, rGO provides an attached framework for SnO2. This structure puts out an effective buffer for the volume change in the reaction process. Meanwhile it facilitates the insertion and detachment of lithium ions.

The rate performance of two materials at gradient current densities are displayed in Figure 4d. After 10 cycles charge-discharge process, SnO2/rGO composite remains capacity of 773.7 mA·h·g-1 at 0.1 C, while SnO2 nanoparticles come up to 500.9 mA·h·g-1 at the same condition. With rate raising up from 0.1 C to 2 C, rate capability diminishes gradually, moreover, the lithium ions storage capability difference of two materials becomes much larger. The reversible capacity of SnO2/rGO composite come up with 210 mA·h·g-1, demonstrating admirable reversible capacity and structure stability under large steady current density [12]. When the rate turns back to 0.1 C, the rate capability of SnO2/rGO composite is basically restored indicating the great reversible capacity. As for Figure 4e, SnO2/rGO displays great long-cycle stability even at current charge and discharge rate of 1 C. When the cycle number comes to 1000, the discharge capacity still remains about 300 mA·h·g-1. Obviously, the excellent rate performance and cycle stability are attributed to the addition of rGO. It remarkably improves the electrode material stability and provides more lithium ions storage sites [34].

EIS spectra were performed to determine the electrical conductivity and electrochemical kinetics of two electrodes (Figure 5a). The composition of the equivalent circuit was shown above the fitting plot, which can be used for quantitative analysis. A lower charge transfer resistance (Rct) value of SnO2/rGO composite (65.7 ohms) indicates that the electrode gains better kinetics after introducing rGO into SnO2 nanoparticles. The reason is that rGO can amend the connections among SnO2 nanoparticles and present sufficient electronic transmission channels which is favorable for electron transfer [12]. For the sake of further exploring the lithium-ion transfer mechanism, linear fitting of Z’ and ω-1/2 at low frequency were carried out in Figure 5b. Lithium-ion diffusion coefficient is a momentous parameter for evaluating the performance of electrode materials. The following equation is applied to calculate the diffusion coefficients (DLi) of lithium ions (Eq. (3)).

$${D}_{Li}={R}^{2}{T}^{2}/2{A}^{2}{n}^{4}{F}^{4}{C}^{2}{\sigma }^{2},$$
(3)

where, R represents gas constant, T stands for Kelvin temperature, A represents the area of electrode, n is the number of electrons transferred during the reaction. F is Faraday constant, σ is the slope of Z’~ω-1/2 and C is lithium-ion phase concentration [35]. From the date in Table 1, we can find that the DLi of SnO2/rGO electrode is much larger than that of SnO2. It illustrates that the structure of SnO2/rGO composite is conducive to the electrolyte diffusion and facilitates Li+ migration during lithiation/delithiation process.

Figure 5
figure 5

a EIS spectra of SnO2/rGO composite and SnO2 anode, b Liner fitting of Z’ and ω -1/2 at low frequency of SnO2/rGO composite and SnO2 anode

Table 1 Kinetic parameters of SnO2/rGO and SnO2 electrodes

The lithium storage kinetics of SnO2/rGO anode were researched by contrasting CV profiles at different scan rates. By estimating the pseudocapacitive contribution, the high rates performance of SnO2/rGO anode can be investigated. Meanwhile the influence in complexing rGO in SnO2 would be further discussed. Figure 6a is the CV profiles of SnO2/rGO composite under the larger current range as scanning rate raises from 0.1 mV·s-1 to 2 mV·s-1, the capacity contribution mechanism can be expressed by Eq. (4).

$$i=a{v}^{b},$$
(4)
$$\text{log}\left(i\right)=b\text{log}\left(v\right)+\text{log}(a),$$
(5)

where, b value could be acquired by drawing log(i)–log(v) curves and fitting the line to obtain the slope (Eq. 5). Generally, when b = 0.5, the capacity contribution is mainly determined by diffusion process, if b = 1, it represents to the typical capacitive charge storage for the surface faradaic redox reaction [36]. The b value obtained from the fitting line of cathodic and anodic peaks are 0.85 and 0.80 for SnO2/rGO composite in Figure 6b. It indicates that the capacitance contribution of capacitor driving in electrochemical behavior is larger than that of diffusion process for SnO2/rGO anode. The b value of SnO2 anode is much closer to 0.5, suggesting that the capacitive contribution is mainly acquired by diffusion process for SnO2 anode. The significantly higher b value of SnO2/rGO anode indicates that much larger specific surface area obtained by adding three-dimensional rGO is the main reason for the obvious increase of capacitive contribution. For the sake of investigating the capacity contribution mechanism, total current (I) is separated into two parts by Eqs. (6) and  (7).

Figure 6
figure 6

CV profiles of SnO2/rGO composite at various scan rates from 0.1 mV·s-1 to 2 mV·s-1, b The index b values of the anodic and cathodic scan, c CV curves of SnO2 and SnO2/rGO composite with scan rate of 1 mV·s-1d Capacitive contribution ratio of SnO2/rGO composite and SnO2 anode at various scan rates

$$I={k}_{1}v+{k}_{2}{v}^{1/2},$$
(6)
$$I/{v}^{1/2}={k}_{1}{v}^{1/2}+k,$$
(7)

where, k1v and k2v1/2 represents the current attributed to conversion behavior and diffusion-controlled reaction respectively [37]. As shown in Figure 6c and d, when the scan rate comes to 1 mV·s-1, the capacitive contribution of SnO2/rGO anode is 71%, which is significantly larger than that of SnO2 anode (56%). This result shows that among two lithium storage mechanisms in SnO2/rGO anode, the capacitive drive process takes an absolute dominant position owing to large specific surface area [38]. The capacitive contribution gets bigger and bigger in the wake of scan rate enlargement as shown in Figure 6d, it is in accord with the charge/discharge profiles and rate performance mentioned before. This result indicates that superior capacitive contribution with large current densities is in favor of clipping Li+ insertion-extraction reaction [36]. Previous discussions prove that the lithium storage behavior of SnO2/rGO anode is mainly driven by capacitive behaviors. Meanwhile it is beneficial to cycling stability and rate performance [21]. Besides, it certifies that the structure of rGO could inhibit the aggregation of SnO2 nanoparticles, which is consistent with the result of SEM.

The charge-discharge mechanism diagram is shown in the Figure 7, and the previous experiments have obtained good electrochemical performance for the following reasons. The addition of rGO provides more loading sites for SnO2 and can effectively prevent SnO2 from falling off during the charging and discharging process. rGO has good electrical conductivity, which is conducive to charge transfer. Besides, rGO itself also has the function of storing lithium ions.

Figure 7
figure 7

Charge-discharge simulation diagram

4 Conclusions

  1. (1)

    In this paper, via a one-step hydrothermal synthesis, graphene was added to SnO2 to alleviate the volume expansion of SnO2 in the charging-discharging process of Li ion batteries.

  2. (2)

    Taking the advantage of specific surface area, high conductivity and density defects of graphene, the aggregation of SnO2 can alleviate structurally.

  3. (3)

    The capacity contribution mechanism changed from diffusion dominated to surface driven capacitive contribution, which provided more sites for storing lithium ions.

  4. (4)

    Even at high current density of 1C and long cycle of 1000 times, the specific capacity of SnO2/rGO can maintain at 300 mA·h·g-1 which has a better practical perspective.

References

  1. H D Liu, J M Huang, X L Li, et al. SnO2 nanorods grown on graphite as a high-capacity anode material for lithium ion batteries. Ceramics International, 2012, 386 (6): 5145-5149.

    Article  Google Scholar 

  2. Q H Tian, Y Tian, W Zhang, et al. Impressive lithium storage of SnO2@TiO2 nanospheres with a yolk-like core derived from self-assembled SnO2 nanoparticles. Journal of Alloys and Compounds, 2017, 70(2): 99-105.

    Article  Google Scholar 

  3. J Zhang, H Ren, J Y Wang, et al. Engineering of multi-shelled SnO2 hollow microspheres for highly stable lithium-ion batteries. Journal of Materials Chemistry A, 2016, 44(5): 17673-17677.

    Article  Google Scholar 

  4. J J Liang, C C Yuan, H H Li, et al. Growth of SnO2 nanoflowers on N-doped carbon nanofibers as anode for Li- and Na-ion batteries. Nano-Micro Letter, 2018, 10(2): 1-9.

    Google Scholar 

  5. N T Wu, W Z Du, X Gao, et al. Hollow SnO2 nanospheres with oxygen vacancies entrapped by a N-doped graphene network as robust anode materials for lithium-ion batteries. Nanoscale, 2018, 10(24): 11460-11466.

    Article  Google Scholar 

  6. Z C Liu, X H Yuan, S S Zhang, et al. Three-dimensional ordered porous electrode materials for electrochemical energy storage. NPG Asia Materials, 2019, 11(1): 1-21.

    Article  Google Scholar 

  7. X Min, B Sun, S Chen, et al. A textile-based SnO2 ultra-flexible electrode for lithium-ion batteries. Energy Storage Materials, 2019, 1(6): 597-606.

    Article  Google Scholar 

  8. L B Chen, M Zhang, W F Wei. Graphene-based composites as cathode materials for Lithium ion batteries, Journal of Nanomaterials, 2013, 2: 2.

    Google Scholar 

  9. X L Chen, Y Y Ma. Wearable Lithium ion batteries based on carbon nanotubes and graphene. Advanced Materials Technologies, 2018, 3(10): 1800041.

    Article  Google Scholar 

  10. X Y Chen, Y Tian. Review of graphene in cathode materials for Lithium-ion batteries. Energy Fuels, 2021, 355(5): 3572-3580.

    Article  Google Scholar 

  11. X X Chen, R Z Cai, P G Liu, et al. Preparation and electrochemical performance of reduced graphene and SnO2 nanospheres composite materials for Lithium-ion batteries and Sodium-ion batteries. ChemistrySelect, 2021, 6(13): 3192-3198.

    Article  Google Scholar 

  12. S Zuo, D Li, Z Wu, et al. SnO2/graphene oxide composite material with high rate performance applied in lithium storage capacity. Electrochimica Acta, 2018, 264: 61-68.

    Article  Google Scholar 

  13. Z Y Lu, Z Kong, L Y Jing, et al. Porous SnO2/graphene composites as anode materials for Lithium-ion batteries: Morphology control and performance improvement. Energy Fuels, 2020, 34(10): 13126-13136.

    Article  Google Scholar 

  14. D P Cai, B H Qu, Q H Li, et al. Reduced graphene oxide uniformly anchored with ultrafine CoMn2O4 nanoparticles as advance anode materials for lithium and sodium storage. Journal of Alloys and Compounds, 2017, 716: 30-36.

    Article  Google Scholar 

  15. S Gupta, N Dimakis. Elucidating the effects of oxygen- and nitrogen-containing functional groups in graphene nanomaterials for applied electrochemistry by density functional theory. Journal of Applied Physics, 2021, 130(8): 084902.

    Article  Google Scholar 

  16. J L Liu, J Wang, C H Xu, et al. Advanced energy storage devices: Basic principles, analytical methods, and rational materials design. Advanced Science, 2018, 5(1): 1700322.

    Article  Google Scholar 

  17. B Quan, A Jin, S H Yu, et al. Solvothermal-derived S-doped graphene as an anode material for Sodium-ion batteries. Advanced Science, 2018, 5(5): 1700880.

    Article  Google Scholar 

  18. L J C Jaemyung Kim, Franklin Kim, Wa Yuan, et al. Graphene oxide sheets at interfaces. Journal of the American Chemical Society, 2010, 132(23): 8180-8186.

  19. K S Novoselov, A K Geim, S V Morozov, et al. Two-dimensional gas of massless dirac fermions in graphene. Nature, 2005, 438(7065): 197-200.

    Article  Google Scholar 

  20. D Wang, J Yang, X Li, et al. Layer by layer assembly of sandwiched graphene/SnO2 nanorod/carbon nanostructures with ultrahigh lithium ion storage properties. Energy Environmental Science, 2013, 6(10): 2900-2906.

    Article  Google Scholar 

  21. G Zhang, S Zeng, L Duan, et al. The dual capacity contribution mechanism of SnSb〢nchored nitrogen doped 3D reduced graphene oxide enhances the performance of Sodiumon batteries. ChemElectroChem, 2020, 7(22): 4663-4671.

    Article  Google Scholar 

  22. H J Zhang, Q Q He, X D Zhu, et al. Surfactant-free solution phase synthesis of monodispersed SnO2 hierarchical nanostructures and gas sensing properties. Crystengcomm, 2012, 14(9): 3169-3176.

    Article  Google Scholar 

  23. S K Park, S H Yu, N Pinna, et al. A facile hydrazine-assisted hydrothermal method for the deposition of monodisperse SnO2 nanoparticles onto graphene for lithium ion batteries. Journal of Materials Chemistry, 2011, 22(6): 2520-2525.

    Article  Google Scholar 

  24. M Xin, A Bs, C A Shi, et al. A textile-based SnO2 ultra-flexible electrode for lithium-ion batteries. Energy Storage Materials, 2019, 16: 597-606.

    Article  Google Scholar 

  25. W Yao, S Wu, L Zhan, et al. Two-dimensional porous carbon-coated sandwich-like mesoporous SnO2/graphene/mesoporous SnO2 nanosheets towards high-rate and long cycle life lithium-ion batteries. Chemical Engineering Journal, 2019, 6: 234-241.

    Google Scholar 

  26. S Jiang, R Huang, W Zhu, et al. Free-standing SnO2@rGO anode via the anti-solvent-assisted precipitation for superior Lithium storage performance. Frontiers in Chemistry, 2019, 7: 878-886.

    Article  Google Scholar 

  27. H D Bian, Y R Tian, C Lee, et al. Mesoporous SnO2 nanostructures of ultrahigh surface areas by novel anodization. ACS Applied Materials & Interfaces, 2016, 8(42): 28862-28871.

    Article  Google Scholar 

  28. J Lin, Z W Peng, C S Xiang, et al. Graphene nanoribbon and nanostructured SnO2 composite anodes for Lithium ion batteries. ACS Nano, 2013, 7(7): 6001-6006.

    Article  Google Scholar 

  29. Y J Chao, X X Yuan, Z F Ma. Preparation and characterization of carbon cryogel (CC) and CC-SiO composite as anode material for lithium-ion battery. Electrochimica Acta, 2008, 5(39): 3468-3473.

    Article  Google Scholar 

  30. R Demir-Cakan, Y S Hu, M Antonietti, et al. Facile one-pot synthesis of mesoporous SnO2 microspheres via nanoparticles assembly and lithium storage properties. Chemistry of Materials, 2008, 20(4): 1227-1229.

    Article  Google Scholar 

  31. R Jia, J L Yue, Q Y Xia, et al. Carbon shelled porous SnO2-delta nanosheet arrays as advanced anodes for lithium-ion batteries. Energy Storage Materials, 2018, 13: 303-311.

    Article  Google Scholar 

  32. X Q Zhang, X X Huang, X Geng, et al. Flexible anodes with carbonized cotton covered by graphene/SnO2 for advanced lithium-ion batteries. Journal of Electroanalytical Chemistry, 2017, 794: 15-22.

    Article  Google Scholar 

  33. C C Hou, S Brahma, S C Weng, et al. Facile, low temperature synthesis of SnO2/reduced graphene oxide nanocomposite as anode material for lithium-ion batteries. Applied Surface Science, 2017, 413: 160-168.

    Article  Google Scholar 

  34. L L Gao, C P Gu, H B Ren, et al. Synthesis of tin(IV) oxide@reduced graphene oxide nanocomposites with superior electrochemical behaviors for lithium-ions batteries. Electrochimica Acta, 2018, 290: 72-81.

    Article  Google Scholar 

  35. C Wang, D Higgins, F F Wang, et al. Controlled synthesis of micro/nanostructured CuO anodes for lithium-ion batteries. Nano Energy, 2014, 9: 334-344.

    Article  Google Scholar 

  36. A Rba, B Gkv, C Sba, et al. Photonic split-second induced mesoporous TiO2-Graphene architectures for efficient sodium-ion batteries. Carbon, 2021, 178: 332-344.

    Article  Google Scholar 

  37. T Brezesinski, J Wang, S H Tolbert, et al. Ordered mesoporous alpha-MoO3 with iso-oriented nanocrystalline walls for thin-film pseudocapacitors. Nature Materials, 2010, 9(2): 146-151.

    Article  Google Scholar 

  38. M Khan, M N Tahir, S F Adil, et al. Graphene based metal and metal oxide nanocomposites: synthesis, properties and their applications. Journal of Materials Chemistry A, 2015, 3(37): 18753-18808.

    Article  Google Scholar 

Download references

Acknowledgements

The authors sincerely thanks to Professor Fushen Lu of Shantou University and Liying Wang of Changchun University of Technology for their critical discussion and reading during manuscript preparation.

Funding

Supported by National Natural Science Foundation of China (Grant No. 61774022), Natural Science Foundation of Guangdong Province (Grant No. 2022A1515011449), Special Program for Science Research Foundation of the Higher Education Institutions of Guangdong Providence (Grant No. 2020ZDZX2052), 2020 Li Ka Shing Foundation Cross-Disciplinary Research Grant (Grant No. 2020LKSFG01A), Research. Start-up Foundation of Shantou University (Grant No. NTF20024).

Author information

Authors and Affiliations

Authors

Contributions

LD was in charge of the whole trial; QL wrote the manuscript; GZ, YQ, ZZ, JW and MZ assisted with sampling and laboratory analyses. All the authors read and approved the final manuscript.

Authors’ information

Qi Li, born in 1993, is currently a master candidate at Advanced Institute of Materials Science & Department of Materials Science and Engineering, Changchun University of Technology, China.

Guoju Zhang, born in 1995, is currently a doctoral candidate at SEU-FEI Nano-Pico Center, Key Laboratory of MEMS of Ministry of Education, Southeast University, China.

Yuanduo Qu, born in 1995, is currently a doctoral candidate at Advanced Energy Materials Laboratory, Department of Chemistry, Shantou University, China.

Zihan Zheng, born in 2000, is currently a doctoral candidate at Advanced Energy Materials Laboratory, Department of Chemistry, Shantou University, China.

Junkai Wang, born in 1992, is a lecturer at Advanced Energy Materials Laboratory, Department of Chemistry, Shantou University, China.

Ming Zhu, born in 1997, is currently a master candidate at Advanced Institute of Materials Science & Department of Materials Science and Engineering, Changchun University of Technology, China.

Lianfeng Duan, born in 1981, is currently a Professor at Advanced Energy Materials Laboratory, Department of Chemistry, Shantou University, China. His research interests include new energy materials and devices. Tel: +86-0754-86503795; E-mail: lfduan@stu.edu.cn.

Corresponding author

Correspondence to Lianfeng Duan.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Q., Zhang, G., Qu, Y. et al. Capacity Contribution Mechanism of rGO for SnO2/rGO Composite as Anode of Lithium-ion Batteries. Chin. J. Mech. Eng. 35, 63 (2022). https://doi.org/10.1186/s10033-022-00739-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s10033-022-00739-8

Keywords