Jun Young Cheong‡
,
Chanhoon Kim‡,
Ji Soo Jang and
Il-Doo Kim*
Department of Materials Science and Engineering, Korea Advanced Institute of Science and Technology, 291 Daehak-ro, Yuseong-gu, Daejeon, 305-701, Republic of Korea. E-mail: idkim@kaist.ac.kr
First published on 23rd December 2015
Ultra thin TiO2 layer (2 nm)-coated SnO2 nanotubes (NTs) wrapped by reduced graphene oxide (rGO) sheets (SnO2@TiO2@rGO) were synthesized as high capacity anode materials for lithium-ion batteries. The rationally designed anodes exhibited superior rate capability while maintaining a high discharge capacity of over 840 mA h g−1 at a current density of 500 mA g−1 after 50 cycles.
However, several drawbacks exist for SnO2 anodes in LIBs including: (i) severe volume expansion (>300%) during cycling which leads to poor cycling retention; (ii) formation of unstable solid electrolyte interphase (SEI) layer; and (iii) low electrical conductivity.17,18 In order to resolve these issues, numerous attempts have been made.19–22 For instance, different researches on Sn–C composite materials have been initially devised and suggested, and they were aimed at utilizing carbon to increase the low electrical conductivity of SnO2 and Sn.23–25 With this approach, however, increasing loading amount of carbon results in decreasing loading amount of either Sn or SnO2, which lead to lower capacity. Another approach has been made to utilize titanium dioxide (TiO2) with tin oxide to make a composite anode material. Combining with TiO2, it forms stable solid-electrolyte interphase (SEI) layer, thereby contributing to more stable nanostructure that maintains structural integrity during lithiation/de-lithiation, with better cycle retention for longer cycles.26,27 Different research strategies have been suggested such as TiO2 nanotubes@SnO2 nanoflakes, TiO2/SnO2/carbon hybrid nanofiber using thermal pyrolysis, SnO2@TiO2 double-shell nanotubes, and SnO2@TiO2 nanowires through atomic layer deposition (ALD).27–30 Nevertheless, the limitation for these approaches is that general synthetic methods were rather complicated series of steps and not applicable for large-scale synthesis.
In this research, we have successfully synthesized ultra thin TiO2 layer (2 nm)-coated SnO2 NTs (hereafter, SnO2@TiO2 NTs) via electrospinning method, subsequent calcination process, and sol–gel process. With uniform TiO2 coating layer on SnO2 NTs, it forms a stable SEI layer, which then enhances cycle retention and structural integrity. In addition, upon reduced graphene oxide (rGO) wrapping, rGO-wrapped SnO2@TiO2 NTs (SnO2@TiO2@rGO NTs) were formed, where the electrical conductivity was significantly increased. With facile and simple synthetic sol–gel methods distinct from the previous works on SnO2/TiO2, it forms stable SEI layer, resulting in better cycle retention. Subsequently, rate capability is also improved by rGO wrapping on SnO2@TiO2 NTs.
The synthesis of SnO2@TiO2@rGO NTs was prepared via electrospinning process, calcination treatment, sol–gel process, and solution-based reduced graphene oxide (rGO)-wrapping process, as illustrated in Fig. 1. Using single-nozzle electrospinning process, Sn precursor/poly(vinylpyrrolidone) (PVP) composite nanofibers (NFs) were fabricated. Upon calcination process at 600 °C for 1 h and 700 °C for 1 h, SnO2 NTs were formed by Kirkendall effect in the following steps:31 (i) as temperature increased, PVP was decomposed and tin (Sn) precursor on the outer surface of the fibers was oxidized and became SnO2; (ii) Sn precursor at the core is also subsequently oxidized and with the Ostwald ripening allows the oxidized Sn at the core to move to the outer SnO2 with larger grain sizes, in addition to the fact that CO2 gas and H2O gas from decomposition of PVP also moves Sn precursor at the core to the outer side; (iii) at the same time, inward flow of vacancies moved to the core at the place where Sn precursor moved outward; (iv) finally, upon subsequent temperature increase, SnO2 underwent nanograin formation, growth, and reorganization, as shown in the previous literature.32
Surface morphology of as-spun Sn precursor/PVP NFs, SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs is shown in Fig. 2. The scanning electron microscopy (SEM) image of Sn precursor/PVP composite NFs with diameter ranged between 300 and 500 nm, is shown in Fig. 2a. The Sn precursor/PVP composite NFs were converted to SnO2 NTs consisting of SnO2 nanograins that have mean grain sizes of 30.6 nm calculated from Scherrer's eqn (1) for the (110) peak in Fig. S1† after calcination (Fig. 2b):
![]() | (1) |
To clearly analyze the surface morphology and crystal structure, we carried out transmission electron microscopy (TEM) analysis. Fig. 2d shows the overall morphology of SnO2@TiO2 NT with uniform thickness across different areas. It can be shown from Fig. 2e that about 2 nm (solid yellow line) of TiO2 layer was uniformly present on the surface of SnO2. To illustrate the spatial distribution of SnO2 and TiO2, transmission electron microscopy energy-dispersive X-ray spectroscopy (TEM-EDS) elemental mapping was performed on as-synthesized SnO2@TiO2 NT (Fig. 2f). According to the elemental mapping images of SnO2@TiO2 NT, TiO2 layers were conformally coated on the surface of SnO2 nanograins. In the selected area electron diffraction (SAED) pattern of SnO2@TiO2 NT (Fig. 2g), however, crystal structure of TiO2 was not clearly shown, as it can be speculated that the concentration of TiO2 was too minimal to have enough intensity to have a clear diffraction pattern, as shown from previous literature.30
To clearly confirm the crystal structure of TiO2, another sample with five times higher concentration of Ti precursor (denoted as SnO2@TiO2 (×5)) was prepared. Fig. S3a† shows the TEM images of SnO2@TiO2 (×5) NT and Fig. S3b† shows its magnified image of red box. The marked interplanar d spacing of 0.352 nm corresponds to the (101) lattice plane of anatase TiO2. The SAED pattern of SnO2@TiO2 (×5) sample indicates the polycrystalline nature of the as-synthesized sample and the diffraction rings can be readily assigned to the SnO2 and dominant peaks of anatase TiO2 such as (101), (004), and (204), which were shown in Fig. S3c.† This anatase structure was also confirmed through X-ray diffraction (XRD) pattern of SnO2@TiO2 (×5) (Fig. S3d†), where anatase peaks were clearly apparent in the lattice planes of (101) and (200) at 25.3° and 48.0°.33,34
Upon rGO wrapping, rGO layers were coated on TiO2 layer and TEM image of SnO2@TiO2@rGO is shown in Fig. 2h, where the thin layers of rGO are shown on the outer side of SnO2@TiO2 NT. Lattice spacing of SnO2 in SnO2@TiO2@rGO NT was shown in Fig. 2i, which was also in agreement with the XRD pattern of SnO2 NTs.35–37 Based on thermogravimetric analysis (TGA) in air of the SnO2@TiO2@rGO NTs, it was calculated that the rGO content was ca. 3.10 wt% in the as-synthesized sample (Fig. 3a).38,39
Fig. 3b shows the Raman spectrum of SnO2@TiO2@rGO NTs. Two peaks at 1343 cm−1 and 1599 cm−1, which corresponded to the D and G band of characteristic carbon peaks, were found and the dimensional ratio of the D band to the G band for the sample was estimated to be 1.09, which indicates the formation of amorphous carbon which are in good agreement with Raman peaks of rGO in similar designs of materials.40–42
Surface chemical states of SnO2@TiO2 and SnO2@TiO2@rGO NTs were also analyzed using X-ray photoelectron spectroscopy (XPS). XPS spectra of Sn and Ti in SnO2@TiO2 NTs are presented in Fig. 4a and b, respectively. Sn4+ 3d3/2 and 3d5/2 have binding energy peaks at 495.0 and 486.6 eV, which are similar to the previously reported values.28 Ti4+ 2p1/2 and 2p3/2 have binding energy peaks at 464.5 and 458.8 eV, which are also similar to the previously reported values.28,43 XPS spectra of Sn and Ti in SnO2@TiO2@rGO NTs are presented in Fig. 4c and d, respectively. Sn4+ 3d3/2 and 3d5/2 have binding energy peaks at 495.5 and 487.1 eV, which are similar to the previously reported values.28 Ti4+ 2p1/2 and 2p3/2 have binding energy peaks at 465.0 and 459.3 eV, which are also similar to the previously reported values.28,44 In summary, XPS spectra for SnO2@TiO2 and SnO2@TiO2@rGO NTs confirmed that SnO2 and TiO2 are present.
A series of electrochemical characterizations were carried out to investigate the electrochemical properties of the SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs based electrodes using 2032-type coin cells. The detailed cell fabrication procedure was described in ESI.† Fig. S4† shows cyclic voltammogram (CV) curves of the first, second and third cycles of SnO2@TiO2 NTs within the range of 0.01–3.0 V vs. Li/Li+ at a scan rate of 0.1 mV s−1. Typically, it is widely known that electrochemical reaction of SnO2 involves two steps (reduction of SnO2 and alloying and de-alloying of Sn) and Li insertion/de-insertion reaction of TiO2 is an intercalation reaction, as shown in the reactions below:30,45,46
SnO2 + 4Li+ + 4e− → Sn + 2Li2O (reduction of SnO2) |
xLi+ + xe− + Sn ↔ LixSn (0 ≤ x ≤ 4.4) (alloying/de-alloying of Sn) |
TiO2 + xLi+ + xe− ↔ LixTiO2 (Li insertion/deinsertion of TiO2) |
In the first cycle, a cathodic peak of 0.75 V was observed in the discharge process, which can be ascribed to the reduction of SnO2 to Sn and the formation of Li2O as well as formation of SEI layer, in accordance with the previous literature.45,52 This process usually occurs at a voltage higher than 0.6 V and is generally known to be irreversible.47–50 Second cathode peak of 0.3 V was observed in the discharge process, which accounts for the alloying of Sn to LixSn. Alloying process occurs at a lower voltage and is known to be reversible.51 With the formation of Li2O and SEI layer in the 1st cycle, the general reaction becomes reversible after the 1st cycle, where CV curves do not vary significantly from the 2nd to 3rd cycle in the discharge process, indicating that the reaction has proceeded with the alloying reaction. At the same time, small peaks at 1.0 V and 1.5 V are mainly attributed to the reaction between Li and TiO2, which is broadly coated on the surface of SnO2.53 This concurs with the previous study that initial CV curve of TiO2 starts with −0.4 mA at 1.0 V, and shows one broad peak at 1.5 V.52 From the 2nd to 3rd cycle in the charge process, two peaks at 0.6 V and 1.3 V are related to the de-alloying of LixSn and oxidation of Sn to SnO2 and it is suggested that small peak at 1.75 V is mainly due to the single-phase Li-insertion/de-insertion reaction in TiO2.52 Presence of smooth, small CV peaks for TiO2 is in agreement with the previous observation that nano-sized systems result in smaller, broader peaks rather than sharp peaks.54
To examine the initial irreversible capacity loss as a result of conversion reaction (reduction of SnO2), the first charge and discharge profiles of SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs were analyzed in the voltage range of 0.005–3.0 V at a current density of 50 mA g−1, as shown in Fig. 5a. SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs have initial discharge capacity of 1429.2 mA h g−1, 1312.5 mA h g−1, and 1531.7 mA h g−1 and initial charge capacity of 944.3 mA h g−1, 873.8 mA h g−1, and 1029.2 mA h g−1, which correspond to the initial coulombic efficiency of 66.1%, 66.6%, and 67.2% for SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs, with similar efficiencies. The capacity values are higher than the theoretical capacity of SnO2 (782 mA h g−1) and theoretical capacity of SnO2@TiO2@rGO (CSnO2@TiO2@rGO = CSnO2*% SnO2 mass + CrGO*% rGO mass = (782) × (0.969) + (372) × (0.031) = 769.3 mA h g−1), which is attributed to the synergistic effects of different components.
To understand the effect of coated TiO2 and wrapped rGO layer on cycle retention, SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs were cycled at a current density of 500 mA g−1 for 70 cycles, as presented in Fig. 5b. Electrochemical properties of SnO2 NTs made under different calcination temperatures were tested, and it was shown that the condition for calcination described above was the optimal condition, as shown in the Fig. S5.† Comparing the cycle retention of SnO2 NTs, SnO2@TiO2, and SnO2@TiO2@rGO from the 2nd cycle to 70th cycle, it is evident that SnO2@TiO2@rGO exhibits better cycle retention (79.7%) compared with that of pristine SnO2 (20.3%) and SnO2@TiO2 (67.1%) NTs. Coulombic efficiency of SnO2@TiO2@rGO NTs maintains a stable state up to 70 cycles, which suggests that enhanced capacity retention is mainly attributed to the hybrid structure that forms stable SEI layer by coating TiO2 layer and increases electrical conductivity by wrapping rGO layers. In addition, the charge and discharge profile of SnO2@TiO2@rGO and SnO2@TiO2 NTs at a current density of 500 mA g−1 are shown in Fig. 5c and S6.†
In order to demonstrate that thin TiO2 layer on the surface of SnO2 helps to form more thin and stable SEI layer, we conducted ex situ SEM analysis of electrodes after 50 cycles at 500 mA g−1. The thinner and smoother SEI layer was formed on the surface of SnO2@TiO2 after 50 cycles at 500 mA g−1 than that of SnO2 in the SEM images (Fig. S7a and b†). Furthermore, we observed that thinner and smoother SEI layer was formed on the surface of SnO2@TiO2@rGO than both SnO2 and SnO2@TiO2 (Fig. S7c†). Owing to the formation of thin and stable SEI layer on the surface, SnO2@TiO2@rGO shows much improved cycle retention and reversible capacity.
The initial irreversible capacity of both SnO2@TiO2 and SnO2@TiO2@rGO NTs comes from the formation of Li2O, an amorphous matrix, and SEI layer.47,48 After the 2nd cycle, both the discharge and charge capacity of SnO2@TiO2 NTs decreased from 812.7 mA h g−1 to 655.3 mA h g−1 and from 769.4 mA h g−1 to 642.6 mA h g−1, showing capacity retention less than 85%. On the other hand, after the 2nd cycle, both the discharge and charge capacity of SnO2@TiO2@rGO NTs only slightly decreased from 974.8 mA h g−1 to 843.3 mA h g−1 and from 930.8 mA h g−1 to 825.8 mA h g−1, maintaining capacity above 800 mA h g−1 with better capacity retention even after 50 cycles.
To understand the effect of rGO layer in enhancement of electrical conductivity, we evaluate the rate capabilities of SnO2, SnO2@TiO2, and SnO2@TiO2@rGO NTs at various current densities (Fig. 5d). SnO2@TiO2@rGO NTs exhibited better rate capabilities compared with both SnO2@TiO2 and SnO2 NTs, especially at higher current densities, which can be suggested from the improved electrical conductivity with rGO-wrapping process that can be demonstrated by impedance tests. Fig. S8† shows the Nyquist plot of each sample after 50th cycle. Semicircles in different frequency regions are ascribed to different factors: semicircle in high frequency region is attributed to the contact resistance by SEI layer formation; semicircle in middle frequency region is attributed to the charge transfer resistance; the straight line in low frequency region describes the mass transfer of lithium-ions.55,56 After the 50th cycle, it is clearly shown that the charge-transfer resistance (Rct) of SnO2@TiO2@rGO is much smaller compared with those of SnO2@TiO2 and SnO2 NTs. Such trends of semicircle areas in Nyquist plots can be suggested that SnO2@TiO2@rGO, upon rGO-wrapping, greatly improved the electrical conductivity of SnO2@TiO2, maintained even after 50 cycles.57,58
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra23704a |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2016 |