Next Article in Journal
Simultaneous In-Cylinder Flow Measurement and Flame Imaging in a Realistic Operating Engine Environment Using High-Speed PIV
Previous Article in Journal
Assessment of Crystallinity Development during Fused Filament Fabrication through Fast Scanning Chip Calorimetry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Enhanced Electrochemical Performance of Sb2O3 as an Anode for Lithium-Ion Batteries by a Stable Cross-Linked Binder

1
Collaborative Innovation Center of Nonferrous Metals of Henan Province, Henan Key Laboratory of High-Temperature Structural and Functional Materials, School of Materials Science and Engineering, Henan University of Science and Technology, Luoyang 471023, China
2
Henan Key Laboratory of Non-Ferrous Materials Science & Processing Technology, Henan University of Science and Technology, Luoyang 471023, China
3
School of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225009, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2019, 9(13), 2677; https://doi.org/10.3390/app9132677
Submission received: 26 May 2019 / Revised: 24 June 2019 / Accepted: 28 June 2019 / Published: 30 June 2019
(This article belongs to the Section Materials Science and Engineering)

Abstract

:
A binder plays an important role in lithium-ion batteries (LIBs), especially for the electrode materials which have large volume expansion during charge and discharge. In this work, we designed a cross-linked polymeric binder with an esterification reaction of Sodium Carboxymethyl Cellulose (CMC) and Fumaric Acid (FA), and successfully used it in an Sb2O3 anode for LIBs. Compared with conventional binder polyvinylidene fluoride (PVDF) and CMC, the new cross-linked binder improves the electrochemical stability of the Sb2O3 anode. Specifically, with CMC-FA binder, the battery could deliver ~611.4 mAh g−1 after 200 cycles under the current density of 0.2 A g−1, while with PVDF or CMC binder, the battery degraded to 265.1 and 322.3 mAh g−1, respectively. The improved cycling performance is mainly due to that the cross-linked CMC-FA network could not only efficiently improve the contact between Sb2O3 and conductive agent, but can also buffer the large volume charge of the electrode during repeated charge/discharge cycles.

1. Introduction

Nowadays, advanced energy conversion and storage devices, such as supercapacitor, lithium-ion batteries (LIBs) and solar cells, have gradually become research hotspots to manage the energy supply for future sustainable resources [1,2,3,4]. Among them, LIBs have been widely used in almost all aspects of our lives for their high energy density, long life span, and environmental friendliness [5,6,7,8,9,10,11]. Although the energy density of LIBs grew very quickly in recent years, it still could not meet people’s growing demand. For graphite, the anode material of a commercial lithium battery, the theoretical specific capacity of which is only about 372 mAh g−1, and this poses a great obstacle to the increase of lithium ion battery capacity [12,13,14,15,16,17,18,19,20,21,22]. Nevertheless, the post-transition metal oxides (PTMOs) have attracted much attention because of their high theoretical capacity (>600 mAh g−1) [23], such as SnO2 [24], Sb2O3 [25,26,27], Sb2O4 [28] and Sb6O13 [29]. Among these PTMOs, Sb2O3 has advantages of low cost, simple synthesis process, and easily availability of raw materials, making it a promising electrode material for rechargeable lithium-ion batteries [30]. However, the lithium storage mechanism of Sb2O3 is found to be phase conversion and alloying, which could cause great volume increases or shrinks during cycling [23]. Thus, Solid Electrolyte Interphases (SEI), on Sb2O3 electrode during initial cycle, would be disrupted, exposing new active site and forming new SEI during cycling, resulting in a decrease in capacity [31].
To stabilize an Sb2O3 electrode, many researchers have been working on controlling the morphology of Sb2O3, such as synthesizing porous Sb2O3 [32] and Sb2O3 nanowires [33], and decorating Sb2O3 with different carbon materials; for instance, carbon cloth [34], reduced graphene oxide [35] and so on. Even though these methods can improve the electrochemical performance of Sb2O3 electrodes, the synthesis processes are complex and of high cost, which limits their wide application. Moreover, the addition of carbon material will reduce the specific capacity of the electrode.
A binder plays a crucial role in affecting the performance of LIBs. Compared to traditional linear binder, cross-linked binder not only could improve the binding between active material and conductive agent, but also can accommodate volume expansion of the electrode during cycling [36,37,38]. The construction of three-dimensional crosslinking structure can improve the cycling stability of the battery, which has been described in many studies [38,39]. For example, Song and his colleagues obtained enhanced electrochemical performance of Si anode for lithium-ion batteries by using a PAA-PVA polymeric binder [39]. Recently, Liu et al. synthesized a stable cross-linking polymeric network of CMC and citric acid (CA) and used it in silicon-based anodes of LIBs, in which an enhanced cycling stability was obtained comparing to linear polymeric PVDF and CMC binders [36]. Moreover, a more stable silicon electrode was also obtained by designing corn starch–maleic anhydride cross-linked binder [40]. In these studies, various novel binder systems for stabilizing the silicon anode for LIBs were fabricated, enhancing their cycling stability.
It is known that Si and Sb2O3 are similar alloying mechanism in degradation of capacity during repeated lithium ion insertion/extraction [38,41]. In this work, we designed a cross-linked polymeric binder with esterification reaction of CMC and Fumaric Acid (FA), and successfully applied it in Sb2O3 anode for LIBs. Compared with conventional binder PVDF and CMC, the cross-linked CMC-FA polymeric network effectively enhanced the Sb2O3 anode electrochemical performance in LIBs. Specifically, with CMC-FA binder, the battery could deliver ~611.4 mAh g−1 (200 cycles, 0.2 A g−1), while with PVDF and CMC, the battery degraded to 265.1 and 322.3 mAh g−1, respectively. This effective solution for enhancing cycling stability of Sb2O3 in LIBs is simple and inexpensive, and could facilitate the commercialization of Sb2O3 for LIBs.

2. Materials and Methods

2.1. Synthesis of Sb2O3 Sample

Sb2O3 particles were synthesized through a self-propagating room temperature reaction method as described elsewhere [42]. Specifically, solutions with 3.310 g SbCl3 and 4.940 g NaOH in 40 mL distilled water were prepared separately. Then, NaOH solution was slowly added to SbCl3 solution (Operator should wear goggles and mask to prevent high temperature chemical liquid splashing). After stirring in beaker for several minutes, the solution was sealed; 3 h later, the obtained slurry was centrifuged and rinsed repeatedly with distilled water to dissolve the excess of NaOH. Finally, the Sb2O3 particle was obtained after 8 h of drying in the vacuum oven at 60 °C.

2.2. Materials Characterization

The crystalline structure of the as-synthesized samples was characterized by X-ray diffraction (XRD, Bruker D8 ADVANCE, Cu kα source), and the microstructures and morphology were observed by scanning electron microscope (SEM, JSM-5610LV, JEOL, Akishima, Japan), and transmission electron microscope (TEM, JSM-2100F, 200 kV, Hitachinaka, Naka, Japan) and high-resolution transmission electron microscope (HRTEM, FEI, TecnaiG2 F30, 200kV, Hitachinaka, Naka, Japan). The sample of cross-linked binder (CMC-FA) was prepared by the method same as preparing electrode without active materials and conductive materials, Fourier transform infrared (FT-IR) spectrascopy of CMC, FA and CMC-FA samples were carried out as diffuse reflectance measurements using a Bruker IFS 66/S FT-IR spectrometer and Spectra-Tech Collector II DRIFTS accessory. The specific surface area of the sample was calculate based on Brunauer–Emmett–Teller (BET) method through nitrogen gas adsorption isotherms measured by using a BEL mini instrument.

2.3. Preparation of Electrodes

The Sb2O3 electrodes using polyvinylidene fluoride (PVDF 5130, Solvay), Sodium Carboxymethyl Cellulose (CMC) and cross-linked Carboxymethyl Cellulose-Fumaric Acid (CMC-FA) as binders, were denoted as PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3, respectively. The electrode slurry was prepared by mixing active materials (Sb2O3), conductive materials (Super P, Shenzhen Kejingstar Technology LTD) and binder with a weight ratio of 6:2:2. The electrode was made by pasting the formed slurry on Cu foil and drying the anode in a vacuum oven at 80 °C for 8 h. For PVDF/CMC@Sb2O3 electrode, the binder was PVDF or CMC, and the solvent was N-Methyl-2-pyrrolidone (NMP) or deionized water, respectively. For CMC-FA@Sb2O3 electrode, the weight ratio of CMC:FA was 1:1, and the solvent was Dimethyl sulfoxide (DMSO). Different from PVDF/CMC@Sb2O3 electrode, CMC-FA@Sb2O3 electrode was dried at room temperature for 8 h in a vacuum oven and further thermally treated at 150 °C for 2 h to guarantee the esterification reaction between CMC and FA. The electrodes were rut into a 12 mm diameter disk, with a load of around 1–1.2 mg (0.9–1.06 mg cm−2) per disk, for electrochemical measurements.

2.4. Electrochemical Measurement

In order to evaluate electrochemical performance, CR2025-type coin cells were assembled in an Ar-filled glove box with low concentrations of water vapor and oxygen (<0.5 ppm). During assembly, as-prepared electrodes were used as cathodes, Lithium foil as anode, Celgard 2300 as separator, and the electrolyte was 1 M LiPF6 in ethylene carbonate (EC) and diethyl carbonate (DEC) with volume ratio 1:1. The electrochemical lithium storage performance were evaluated through a two-electrode system, in which lithium foil was employed as the counter and reference electrode, and the Sb2O3 electrodes were employed as working electrode. The galvanostatic charge–discharge performance were evaluated by a LAND battery-testing system (CT2001A, Wuhan, China) between 0.01–3.0 V vs. Li/Li+. Electrochemical impedance spectroscopy (EIS) was conducted ranging from 100 kHz to 0.01 Hz, and cyclic voltammetry was carried out in the range of 0.01–3.0 V (vs. Li/Li+) at a scan rate of 0.2 mV s−1. The above two electrochemical tests were carried out on an electrochemical workstation (CHI 660E).

3. Results

3.1. Material Characterization

The crystallographic phase of the synthesized Sb2O3 was characterized by X-ray diffraction, and the XRD pattern is depicted in Figure 1a, in which all the strong and sharp peaks are consistent with standard pattern of the Orthorhombic Sb2O3 phase (JCPDS card No. 71-0383), indicating the high crystalline Sb2O3 was obtained. Figure 1b shows the SEM image of Sb2O3, it is clear that this product is in cuboid-like shape with sizes of 200–400 nm in width and 0.6–2 μm in length, and the cuboid particles easily aggregated together to form bigger particles. The HRTEM image of the as-prepared Sb2O3 was exhibited in Figure 1c, the interplanar spacing is ~3.15 Å, this corresponds to the spacing between the (121) planes of Antimony Oxide. This is consistent with the X-ray diffraction pattern, as shown in Figure 1a.
The Sb2O3 sample was also characterized by Nitrogen adsorption–desorption isotherm measurement, and a typical IV-type isotherm was obtained, as shown in Figure 1d, indicating that there are nearly no micropores existing in the Sb2O3 sample [43]. A large amount of mesopores (most diameters are less than 40 nm) could be observed in the pore-size distribution curve (inset in Figure 1d). Moreover, the hysteresis hoop of isotherms ascribes that the Sb2O3 aggregated and formed a slit, consisting with SEM image of Sb2O3 particles [44]. According to the BET equation and Barrett–Joyner–Halenda (BJH) Plot, the specific surface area of the Sb2O3 sample was ~3.42 m2 g−1.
Figure 2a presents the formation process of cross-linking CMC-FA network by CMC and FA through the esterification reaction. A cross-linking CMC-FA network was formed via the condensation process of the hydroxyl group of CMC and carboxyl group of FA. To identify this chemical reaction, between CMC and FA, the FT-IR measurement was carried out (Figure 2b). For the FA spectrum, the hydroxyl bond dominates the range from 3700 to 2800 cm−1, and the absorption band between 3200 and 2800 cm−1 could be ascribed to hydroxyl bond between -COOH groups of FA [45]. By contrast, for the spectrum of CMC-FA, the nearly disappeared band between 3200 and 2800 cm−1 indicates that the hydroxyl bond in carboxyl of FA was greatly weakened after the cross-linking reaction with CMC and ~3400 cm−1 has a broad peak which may be attributed to the residual hydroxyl bond of CMC. Moreover, the FT-IR spectra in Figure 2b showed that the C=O band in -COOH of FA shifted from 1677 cm−1 to 1681 cm−1, which could be ascribed to an ester group formed via cross linking during condensation reaction between FA and CMC [37]. This esterification reaction of the binders in the electrode could form a cross-linked polymeric network, which could increase the connection within the electrode (Sb2O3, Super P, binder and Cu foil). At the same time, it provides an accommodation for the volume change of the Sb2O3 electrode.

3.2. Electrochemical Properties

The cyclic voltammetry (CV) of PVDF@Sb2O3 and CMC@Sb2O3 electrode at a scan rate of 0.2 mV s−1 between 0.01 V and 3 V was depicted in Figure 3a,b. Both PVDF@Sb2O3 and CMC@Sb2O3 electrodes have similar curves so that the cathodic peaks located at about 1 V could be assigned to the lithiation of Sb2O3 (6Li+ + Sb2O3 +6e → 2Sb + 3Li2O) during the first cycle [23,46], and the following reduction peaks at about 0.6 V may be associated with the multi-step lithiation of Sb (6Li+ + 2Sb + 6e → 2Li3Sb). The shift from ~0.8 V to ~1.5 V in the subsequent reduction peaks show an activation procedure in the initial lithium-ion insertion process [23]. As for oxidation peaks, the peaks centered at about 1.2 V and 1.5 V may be caused by de-alloying process (2Li3Sb → 6Li+ + 2Sb + 6e). As for the CV curves of CMC-FA@Sb2O3 electrode in Figure 3c, some characteristic peaks were detected where the triangle notation in the figure was caused by lithiation and delithiation process as described above. In addition, those extra reduction peaks and distorted oxidation peaks may be attributed to some side reactions under the effect of CMC-FA polymer network.
As the Figure 3d–f shows, the lithiation/delithiation plateaus of the PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 at 0.2 A g−1, were nearly consistent with the peaks in the cyclic voltammetry curve. The initial discharge capacity of CMC-FA@Sb2O3 is 1315.8 mAh g−1, which is higher than PVDF@Sb2O3 (1236 mAh g−1) and CMC@Sb2O3 (1141.3 mAh g−1). The initial coulombic efficiency of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 is 72.73%, 70.24% and 60.09%, respectively. The low coulombic efficiency of three electrodes may be attributed to a form of SEI and conversion reaction of Sb2O3 [47]. In addition, the charge–discharge curves in the second to fifth cycles of CMC-FA@Sb2O3 overlapped obviously, while the curves of PVDF@Sb2O3 and CMC@Sb2O3 dispersed, indicating that electrodes will have higher cycling stability with the CMC-FA polymer network.
The cycling performances of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 as anode materials at current density of 0.2 A g−1 were depicted in Figure 4a. During the first 50 cycles, the specific discharge capacity of CMC-FA@Sb2O3 gradually decreases to 400.5 mAh g−1, which may be attributed to some side reactions during the discharge stage of CMC-FA@Sb2O3 electrode (CMC-FA polymer binder has residual hydroxyl group, which may react with PF5 in electrolyte) [38]. From the 50th cycle, the discharge curve CMC-FA@Sb2O3 goes upward, after 200 cycles, the specific discharge capacity of CMC-FA@Sb2O3 gradually increases to 611.4 mAh g−1. The upward trend may be attributed to the activation of CMC-FA@Sb2O3 electrode, some active sites were covered for the reason of as-prepared Sb2O3 particles are easily aggregate, which we have discussed in SEM image and Nitrogen adsorption–desorption isotherms measurement of Sb2O3 particle. Changes in the volume of Sb2O3 particle during lithiation/delithiation may lead to exposure of new active sites and result in capacities increasing. In contrast, the specific discharge capacity of PVDF@Sb2O3 and CMC@Sb2O3 decreases to 286.3 and 272.3 mAh g−1 after 200 cycles, respectively. The abnormal cycling performance and coulomb efficiency of PVDF@Sb2O3 after 150 cycles may be caused by the collapse of the electrode structure with shedding of active materials. The excellent cycling stability of CMC-FA@Sb2O3 electrode may be ascribed to that the cross-linking network could reinforce the electrode, as well as accommodate the volume change of the electrode in repeated lithiation and delithiation process.
Figure 4b shows the rate capabilities of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 at current density ranging from 0.05 to 0.5 A g−1. At a current density of 50 and 100 mA g−1, capacity of CMC-FA@Sb2O3 is lower than PVDF@Sb2O3 and CMC@Sb2O3 under low current density. This may be due to the slow activation of electrodes under the effect of aggregated Sb2O3 and CMC-FA binder. PVDF@Sb2O3 and CMC@Sb2O3 have same downward trends even though they have higher specific discharge capacity than CMC-FA@Sb2O3, whose specific discharge capacity is relatively stable. When current density was increased to 0.2, 0.3 and 0.5 A g−1, the reversible capacities of CMC-FA@Sb2O3 were 382.9, 344.3 and 295.4 mAh g−1, respectively, which is better than PVDF@Sb2O3 and CMC@Sb2O3. Moreover, when the current density returns to the initial, the capacity of CMC-FA@Sb2O3 could recover to ~569 mAh g−1, implying the good rate capability.
The EIS measurements were conducted to test the electrochemical kinetics of different kinds of electrodes, and the test frequencies ranged from 0.1 to 100 kHz. Figure 4c,d show the Nyquist plots of three kinds of electrodes in initial states and after 200 cycles at 0.2 A g−1, respectively. The equivalent circuit (inset in Figure 4c,d) was obtained using Nova software. For the impedance spectra before the discharge as shown in Figure 4c, the semicircles in the high-frequency region correspond to charge-transfer resistance (Rct) in the interface of electrodes and electrolyte, and the inclined lines in the low frequency region are attributed to insertion of lithium-ions into the negative electrode [48]. Re was the internal resistance of the as-assembled cell, CPEl represents constant phase element of the electrode/electrolyte interface, and Zw represents the Warburg impedance. The Rct of the initial CMC-FA@Sb2O3 electrode is ~269 Ω, and the PVDF@Sb2O3 and CMC@Sb2O3 are about 733 and 593 Ω, respectively, much larger than the CMC-FA@Sb2O3. It can be seen in Figure 4d that in the impedance spectra after 200 cycles, the semicircle in the high frequency region corresponds to the resistance of SEI film, and the semicircle at middle frequency correspond to Rct, the inclined line at low frequency reflections are attributed to insertion of lithium-ions into the negative electrode [49]. The meaning of Rf and CPE2 in equivalent circuits were corresponding to the resistance and constant phase element of the SEI film. As we can see, although the discharge capacity of CMC-FA@Sb2O3 electrode is much higher than the other two electrodes after 200 cycles at 0.2 A g−1 (Figure 4a), they have quite similar overall impedance after 200 cycles, which could be ascribed to the Li anode contribution; the anode SEI file impedance and the anode charge-discharge resistance may have obvious change during cycling [50]. The Rct decreased to ~55 Ω, ~74.6 Ω and 90.1 Ω for CMC-FA@Sb2O3, PVDF@Sb2O3, CMC@Sb2O3 electrodes, respectively. This fact confirms that the CMC-FA binder could not only increase the connection within the electrode (Sb2O3, super p, binder and Cu foil), but also reduce the particle shedding due to volume changes of Sb2O3 electrode during cycling, enhancing the electrode’s electrochemical performance.
To further illustrate the performance of the binder, the morphology of electrodes was examined before and after 200 cycles at 0.2 A g−1, as depicted in Figure 5. The morphologies of different electrodes before cycling were shown in Figure 5a–c, in which they exhibited similar morphology. Interestingly, a flower-like polymer network could be found in Figure 5c, which may be associated with cross-linking reactions. After 200 cycles, it can be easily found that there are many large cracks and pulverization in the PVDF@Sb2O3 electrode and CMC@Sb2O3 electrode, shown in Figure 5d,e. As a comparison, the CMC-FA@Sb2O3 electrode has better structure integrity with smaller cracks as shown in Figure 5f, so that CMC-FA@Sb2O3 has better cycling stability at same charge/discharge conditions. This result indicates that CMC-FA polymer network can mitigate the strain caused by lithiation/delithiation during cycling (Figure 6).
Furthermore, we have compared the electrochemical properties of the CMC-FA@Sb2O3 electrode with a reported study on Sb2O3-based composites as anodes for LIBs in Table 1. First, the Synthesis of the Sb2O3 sample and cross-linked polymeric binder have the advantages of cheap raw materials, simple reaction and short reaction time, which can greatly reduce the production cost. Second, the superior capacity and long-term cycling performance of CMC-FA@Sb2O3 can be explained as that CMC-FA polymer network could improve the integrity of electrode and could buffer the strain caused by volume change of Sb2O3 during the recharging process.

4. Conclusions

In conclusion, we successfully designed a cross-linked polymeric binder through esterification reaction of CMC and FA and used it in Sb2O3 for LIBs. It was found that CMC-FA@Sb2O3 electrodes exhibited good cycling stability and rate capability, better than the electrodes with conventional binder. The results are ascribed to that the novel binder could improve the connection between active materials, conductive agent, and current collector, and also could accommodate the volume expansion of the Sb2O3 anode in the cycles. Further studies will probably focus on preventing side reactions, improving the cycling stability of initial cycles and solving the aggregation of Sb2O3 by decorating with functional materials, which may reduce the activation time and improve the early cycling performance of the CMC-FA@Sb2O3 electrode. Furthermore, this low-cost cross-linked binder may be generalized to other electrode materials for lithium ion batteries, which suffer from large volume-changes during cycling.

Author Contributions

Y.L., F.R. and H.P. conceived and designed the experiments; H.W., K.Y. and Y.Y. performed the experiments; H.W., J.M. and K.P. analyzed the data; H.W. wrote the paper, Y.L. and G.W. revised the paper. All the authors have discussed the results and commented on the manuscript.

Funding

This work was supported by Program for Changjiang Scholars and Innovative Research Team in University (IRT_16R21), the Chinese 02 Special Fund (2017ZX02408003), the Key Research Projects of Colleges and Universities in the Henan Province (15A430023), the Scientific and Technological Project of Henan Province (182102210297), Open Fund of National Joint Engineering Research Center for abrasion control and molding of metal materials (HKDNM201802, HKDNM201807), Scientific Research Starting Foundation for Ph.D. of Henan University of Science and Technology (13480065), Science Foundation for Youths of Henan University of Science and Technology (2013QN006) and the Student Research Training Plan of Henan University of Science and Technology (2018029).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chu, S.; Cui, Y.; Liu, N. The path towards sustainable energy. Nat. Mater. 2017, 16, 16–22. [Google Scholar] [CrossRef] [PubMed]
  2. Armand, M.; Tarascon, J.M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  3. Larcher, D.; Tarascon, J.M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19–29. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Y.; Zhai, X.L.; Yang, K.K.; Wang, F.; Wei, H.J.; Zhang, W.H.; Ren, F.Z.; Pang, H. Mesoporous NH4NiPO4 center dot H2O for High-Performance Flexible All-Solid-State Asymmetric Supercapacitors. Front. Chem. 2019, 7, 118. [Google Scholar] [CrossRef] [PubMed]
  5. Suga, T.; Konishi, H.; Nishide, H. Photocrosslinked nitroxide polymer cathode-active materials for application in an organic-based paper battery. Chem. Commun. (Camb. Engl.) 2007, 2017, 1730–1732. [Google Scholar] [CrossRef] [PubMed]
  6. Tang, Y.; Zhang, Y.; Li, W.; Ma, B.; Chen, X. Rational material design for ultrafast rechargeable lithium-ion batteries. Chem. Soc. Rev. 2015, 44, 5926–5940. [Google Scholar] [CrossRef] [PubMed]
  7. Arico, A.S.; Bruce, P.; Scrosati, B.; Tarascon, J.-M.; van Schalkwijk, W. Nanostructured materials for advanced energy conversion and storage devices. Nat. Mater. 2005, 4, 366–377. [Google Scholar] [CrossRef]
  8. Wu, N.T.; Du, W.Z.; Gao, X.; Zhao, L.; Liu, G.L.; Liu, X.M.; Wu, H.; He, Y.B. Hollow SnO2 nanospheres with oxygen vacancies entrapped by a N-doped graphene network as robust anode materials for lithium-ion batteries. Nanoscale 2018, 10, 11460–11466. [Google Scholar] [CrossRef] [PubMed]
  9. Whittingham, M.S. Lithium batteries and cathode materials. Chem. Rev. 2004, 104, 4271–4301. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, Y.Y.; Wang, Z.J.; Lei, D.N.; Lv, W.; Zhao, Q.; Ni, B.; Liu, Y.; Li, B.H.; Kang, F.Y.; He, Y.B. Spherical Li Deposited inside 3D Cu Skeleton as Anode with Ultrastable Performance. ACS Appl. Mater. Interfaces 2018, 10, 20244–20249. [Google Scholar] [CrossRef]
  11. Zheng, J.C.; Yang, Z.; He, Z.J.; Tong, H.; Yu, W.J.; Zhang, J.F. In situ formed LiNi0.8Co0.15Al0.05O2@Li4SiO4 composite cathode material with high rate capability and long cycling stability for lithium-ion batteries. Nano Energy 2018, 53, 613–621. [Google Scholar] [CrossRef]
  12. Yan, J.; Liu, X.; Li, B. Recent progress in Li-rich layered oxides as cathode materials for Li-ion batteries. RSC Adv. 2014, 4, 63268–63284. [Google Scholar] [CrossRef]
  13. Zhu, C.; Mu, X.; van Aken, P.A.; Maier, J.; Yu, Y. Fast Li Storage in MoS2-Graphene-Carbon Nanotube Nanocomposites: Advantageous Functional Integration of 0D, 1D, and 2D Nanostructures. Adv. Energy Mater. 2015, 5, 1401170. [Google Scholar] [CrossRef]
  14. Wang, H.; Feng, H.; Li, J. Graphene and Graphene-like Layered Transition Metal Dichalcogenides in Energy Conversion and Storage. Small 2014, 10, 2165–2181. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, J.; Wen, Y.; Wang, Y.; van Aken, P.A.; Maier, J.; Yu, Y. Carbon-Encapsulated Pyrite as Stable and Earth-Abundant High Energy Cathode Material for Rechargeable Lithium Batteries. Adv. Mater. 2014, 26, 6025–6030. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, H.; Liang, Q.; Wang, W.; An, Y.; Li, J.; Guo, L. Preparation of Flower-like SnO2 Nanostructures and Their Applications in Gas-Sensing and Lithium Storage. Cryst. Growth Des. 2011, 11, 2942–2947. [Google Scholar] [CrossRef]
  17. Tang, Y.; Zhang, Y.; Rui, X.; Qi, D.; Luo, Y.; Leow, W.R.; Chen, S.; Guo, J.; Wei, J.; Li, W.; et al. Conductive Inks Based on a Lithium Titanate Nanotube Gel for High-Rate Lithium-Ion Batteries with Customized Configuration. Adv. Mater. 2016, 28, 1567–1576. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, H.; Wu, Y.; Bai, Y.; Zhou, W.; An, Y.; Li, J.; Guo, L. The self-assembly of porous microspheres of tin dioxide octahedral nanoparticles for high performance lithium ion battery anode materials. J. Mater. Chem. 2011, 21, 10189–10194. [Google Scholar] [CrossRef]
  19. Liu, Y.; Wei, H.J.; Wang, C.; Wang, F.; Wang, H.C.; Zhang, W.H.; Wang, X.F.; Yan, C.L.; Kim, B.H.; Ren, F.Z. Nitrogen-Doped Carbon Coated WS2 Nanosheets as Anode for High-Performance Sodium-Ion Batteries. Front. Chem. 2018, 6, 236. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, F.; Liu, Y.; Zhao, Y.F.; Wang, Y.; Wang, Z.J.; Zhang, W.H.; Ren, F.Z. Facile Synthesis of Two-Dimensional Porous MgCo2O4 Nanosheets as Anode for Lithium-Ion Batteries. Appl. Sci. 2018, 8, 22. [Google Scholar] [CrossRef]
  21. Liu, G.L.; Cui, J.; Luo, R.J.; Liu, Y.; Huang, X.X.; Wu, N.T.; Jin, X.Y.; Chen, H.P.; Tang, S.Y.; Kim, J.K.; et al. 2D MoS2 grown on biomass-based hollow carbon fibers for energy storage. Appl. Surf. Sci. 2019, 469, 854–863. [Google Scholar] [CrossRef]
  22. Liu, G.; Liu, X.; Li, X.; Jin, X.; Xiao, H.; Wu, N.; Guo, D.; Tang, S. Coordination Engineering Construction of Si@ZnS@N,S-Doped Reduced Graphene Oxide Nanocomposite as Anode Material with Enhanced Lithium Storage Performance. Energy Technol. 2019, 7, 1900186. [Google Scholar] [CrossRef]
  23. Zhou, X.; Zhang, Z.; Lu, X.; Lv, X.; Ma, G.; Wang, Q.; Lei, Z. Sb2O3 Nanoparticles Anchored on Graphene Sheets via Alcohol Dissolution-Reprecipitation Method for Excellent Lithium-Storage Properties. ACS Appl. Mater. Interfaces 2017, 9, 34927–34936. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, J.X.; Li, W.; Wang, F.; Xia, Y.Y.; Asiri, A.M.; Zhao, D.Y. Controllable synthesis of SnO2@C yolk-shell nanospheres as a high-performance anode material for lithium ion batteries. Nanoscale 2014, 6, 3217–3222. [Google Scholar] [CrossRef] [PubMed]
  25. Tan, Y.; Chen, L.; Chen, H.; Hou, Q.; Chen, X. Synthesis of a symmetric bundle-shaped Sb2O3 and its application for anode materials in lithium ion batteries. Mater. Lett. 2018, 212, 103–106. [Google Scholar] [CrossRef]
  26. Hong, K.S.; Nam, D.H.; Lim, S.J.; Sohn, D.; Kim, T.H.; Kwon, H. Electrochemically Synthesized Sb/Sb2O3 Composites as High-Capacity Anode Materials Utilizing a Reversible Conversion Reaction for Na-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 17264–17271. [Google Scholar] [CrossRef] [PubMed]
  27. Hu, M.J.; Jiang, Y.Z.; Sun, W.P.; Wang, H.T.; Jin, C.H.; Yan, M. Reversible Conversion-Alloying of Sb2O3 as a High-Capacity, High-Rate, and Durable Anode for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2014, 6, 19449–19455. [Google Scholar] [CrossRef]
  28. Zhou, X.Z.; Zhang, Z.F.; Wang, J.W.; Wang, Q.T.; Ma, G.F.; Lei, Z.Q. Sb2O4/reduced graphene oxide composite as high-performance anode material for lithium ion batteries. J. Alloys Compd. 2017, 699, 611–618. [Google Scholar] [CrossRef]
  29. Zhou, X.Z.; Zhang, Z.F.; Xu, X.H.; Yan, J.; Ma, G.F.; Lei, Z.Q. Anchoring Sb6O13 Nanocrystals on Graphene Sheets for Enhanced Lithium Storage. ACS Appl. Mater. Interfaces 2016, 8, 35398–35406. [Google Scholar] [CrossRef]
  30. Deng, M.X.; Li, S.J.; Hong, W.W.; Jiang, Y.L.; Xu, W.; Shuai, H.L.; Zou, G.Q.; Hu, Y.C.; Hou, H.S.; Wang, W.L.; et al. Octahedral Sb2O3 as high-performance anode for lithium and sodium storage. Mater. Chem. Phys. 2019, 223, 46–52. [Google Scholar] [CrossRef]
  31. Gao, H.; Zhou, W.; Jang, J.-H.; Goodenough, J.B. Cross-Linked Chitosan as a Polymer Network Binder for an Antimony Anode in Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1502130. [Google Scholar] [CrossRef]
  32. Pan, J.; Wang, N.; Zhou, Y.; Yang, X.; Zhou, W.; Qian, Y.; Yang, J. Simple synthesis of a porous Sb/Sb2O3 nanocomposite for a high-capacity anode material in Na-ion batteries. Nano Res. 2017, 10, 1794–1803. [Google Scholar] [CrossRef]
  33. Nam, D.-H.; Hong, K.-S.; Lim, S.-J.; Kim, M.-J.; Kwon, H.-S. High-Performance Sb/Sb2O3 Anode Materials Using a Polypyrrole Nanowire Network for Na-Ion Batteries. Small 2015, 11, 2885–2892. [Google Scholar] [CrossRef] [PubMed]
  34. Fei, J.; Cui, Y.; Li, J.; Xu, Z.; Yang, J.; Wang, R.; Cheng, Y.; Hang, J. A flexible Sb2O3/carbon cloth composite as a free-standing high performance anode for sodium ion batteries. Chem. Commun. 2017, 53, 13165–13167. [Google Scholar] [CrossRef] [PubMed]
  35. Li, D.; Yana, D.; Ma, J.; Qin, W.; Zhang, X.; Lu, T.; Pan, L. One-step microwave-assisted synthesis of Sb2O3/reduced graphene oxide composites as advanced anode materials for sodium-ion batteries. Ceram. Int. 2016, 42, 15634–15642. [Google Scholar] [CrossRef]
  36. Liu, Y.; Tai, Z.; Zhou, T.; Sencadas, V.; Zhang, J.; Zhang, L.; Konstantinov, K.; Guo, Z.; Liu, H.K. An All-Integrated Anode via Interlinked Chemical Bonding between Double-Shelled-Yolk-Structured Silicon and Binder for Lithium-Ion Batteries. Adv. Mater. 2017, 29, 1703028. [Google Scholar] [CrossRef] [PubMed]
  37. Koo, B.; Kim, H.; Cho, Y.; Lee, K.T.; Choi, N.S.; Cho, J. A Highly Cross-Linked Polymeric Binder for High-Performance Silicon Negative Electrodes in Lithium Ion Batteries. Angew. Chem. Int. Ed. 2012, 51, 8762–8767. [Google Scholar] [CrossRef] [PubMed]
  38. Wei, Y.; Wang, Z.; Ye, H.; Mou, J.; Lei, D.; Liu, Y.; Lv, W.; Li, B.; Kang, F.; He, Y.-B. A Stable Cross-Linked Binder Network for SnO2 Anode with Enhanced Sodium-Ion Storage Performance. ChemistrySelect 2017, 2, 11365–11369. [Google Scholar] [CrossRef]
  39. Song, J.X.; Zhou, M.J.; Yi, R.; Xu, T.; Gordin, M.L.; Tang, D.H.; Yu, Z.X.; Regula, M.; Wang, D.H. Interpenetrated Gel Polymer Binder for High-Performance Silicon Anodes in Lithium-ion Batteries. Adv. Funct. Mater. 2014, 24, 5904–5910. [Google Scholar] [CrossRef]
  40. Rohan, R.; Kuo, T.C.; Chiou, C.Y.; Chang, Y.L.; Li, C.C.; Lee, J.T. Low-cost and sustainable corn starch as a high-performance aqueous binder in silicon anodes via in situ cross-linking. J. Power Sources 2018, 396, 459–466. [Google Scholar] [CrossRef]
  41. Chevrier, V.L.; Ceder, G. Challenges for Na-ion Negative Electrodes. J. Electrochem. Soc. 2011, 158, A1011–A1014. [Google Scholar] [CrossRef]
  42. Vukovic, M.; Brankovic, Z.; Poleti, D.; Recnik, A.; Brankovic, G. Novel simple methods for the synthesis of single-phase valentinite Sb2O3. J. Sol-Gel Sci. Technol. 2014, 72, 527–533. [Google Scholar] [CrossRef]
  43. Shi, Y.T.; Lv, W.; Niu, S.Z.; He, Y.B.; Zhou, G.M.; Chen, G.H.; Li, B.H.; Yang, Q.H.; Kang, F.Y. A Carbon-Sulfur Hybrid with Pomegranate-like Structure for Lithium-Sulfur Batteries. Chem. Asian J. 2016, 11, 1343–1347. [Google Scholar] [CrossRef] [PubMed]
  44. Chou, Y.H.; Cundy, C.S.; Garforth, A.A.; Zholobenko, V.L. Mesoporous ZSM-5 catalysts: Preparation, characterisation and catalytic properties. Part I: Comparison of different synthesis routes. Microporous Mesoporous Mater. 2006, 89, 78–87. [Google Scholar] [CrossRef]
  45. Wang, L.; Wang, D.; Zhang, F.X.; Jin, J. Interface Chemistry Guided Long-Cycle-Life Li-S Battery. Nano Lett. 2013, 13, 4206–4211. [Google Scholar] [CrossRef] [PubMed]
  46. Xue, M.Z.; Fu, Z.W. Electrochemical reaction of lithium with nanostructured thin film of antimony trioxide. Electrochem. Commun. 2006, 8, 1250–1256. [Google Scholar] [CrossRef]
  47. Hou, H.S.; Jing, M.J.; Yang, Y.C.; Zhang, Y.; Zhu, Y.R.; Song, W.X.; Yang, X.M.; Ji, X.B. Sb porous hollow microspheres as advanced anode materials for sodium-ion batteries. J. Mater. Chem. A 2015, 3, 2971–2977. [Google Scholar] [CrossRef]
  48. Jin, J.; Huang, S.Z.; Shu, J.; Wang, H.E.; Li, Y.; Yu, Y.; Chen, L.H.; Wang, B.J.; Su, B.L. Highly porous TiO2 hollow microspheres constructed by radially oriented nanorods chains for high capacity, high rate and long cycle capability lithium battery. Nano Energy 2015, 16, 339–349. [Google Scholar] [CrossRef]
  49. Bao, W.Z.; Liu, L.; Wang, C.Y.; Choi, S.; Wang, D.; Wang, G.X. Facile Synthesis of Crumpled Nitrogen-Doped MXene Nanosheets as a New Sulfur Host for Lithium-Sulfur Batteries. Adv. Energy Mater. 2018, 8, 1702485. [Google Scholar] [CrossRef]
  50. Gao, P.; Zhang, C.F.; Wen, G.W. Equivalent circuit model analysis on electrochemical impedance spectroscopy of lithium metal batteries. J. Power Sources 2015, 294, 67–74. [Google Scholar] [CrossRef]
  51. Yang, X.; Ma, J.J.; Wang, H.J.; Chai, Y.Q.; Yuan, R. Partially reduced Sb/Sb2O3@C spheres with enhanced electrochemical performance for lithium ion storage. Mater. Chem. Phys. 2018, 213, 208–212. [Google Scholar] [CrossRef]
  52. Wang, Z.M.; Cheng, Y.; Li, Q.; Chang, L.M.; Wang, L.M. Facile synthesis of one-dimensional hollow Sb2O3@TiO2 composites as anode materials for lithium ion batteries. J. Power Sources 2018, 389, 214–221. [Google Scholar] [CrossRef]
  53. Zhou, J.; Zheng, C.; Wang, H.; Yang, J.; Hu, P.; Guo, L. 3D nest-shaped Sb2O3/RGO composite based high-performance lithium-ion batteries. Nanoscale 2016, 8, 17131–17135. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Characterization of as-synthesized Sb2O3 sample: (a) X-ray diffraction patterns; (b) scanning electron microscope image; (c) high-resolution transmission electron microscope image; (d) Nitrogen adsorption-desorption isothermal diagram and Barrett–Joyner–Halenda Plot (inset).
Figure 1. Characterization of as-synthesized Sb2O3 sample: (a) X-ray diffraction patterns; (b) scanning electron microscope image; (c) high-resolution transmission electron microscope image; (d) Nitrogen adsorption-desorption isothermal diagram and Barrett–Joyner–Halenda Plot (inset).
Applsci 09 02677 g001
Figure 2. (a) The schematic of esterification reaction between CMC and FA; (b) FTIR spectra of FA, CMC and CMA-FA polymer network.
Figure 2. (a) The schematic of esterification reaction between CMC and FA; (b) FTIR spectra of FA, CMC and CMA-FA polymer network.
Applsci 09 02677 g002
Figure 3. (ac) Cyclic voltammetry curves of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 at scan rate of 0.2 mV s−1; (df) galvanostatic charge-discharge profiles of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 in the initial five cycles at 0.2 A g−1.
Figure 3. (ac) Cyclic voltammetry curves of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 at scan rate of 0.2 mV s−1; (df) galvanostatic charge-discharge profiles of PVDF@Sb2O3, CMC@Sb2O3 and CMC-FA@Sb2O3 in the initial five cycles at 0.2 A g−1.
Applsci 09 02677 g003
Figure 4. (a) Cycling performances of CMC-FA@Sb2O3, PVDF@Sb2O3 and CMC@Sb2O3 at 0.2 A g−1; (b) Rate capabilities of CMC-FA@Sb2O3, PVDF@Sb2O3 and CMC@Sb2O3 at current density ranging from 0.05 A g−1 to 0.5 A g−1; Nyquist plots of different electrodes: (c) before cycling; (d) after 200 cycles at current density of 0.2 A g−1, inset shows electrochemical impedance spectroscopy (EIS) test circuitry.
Figure 4. (a) Cycling performances of CMC-FA@Sb2O3, PVDF@Sb2O3 and CMC@Sb2O3 at 0.2 A g−1; (b) Rate capabilities of CMC-FA@Sb2O3, PVDF@Sb2O3 and CMC@Sb2O3 at current density ranging from 0.05 A g−1 to 0.5 A g−1; Nyquist plots of different electrodes: (c) before cycling; (d) after 200 cycles at current density of 0.2 A g−1, inset shows electrochemical impedance spectroscopy (EIS) test circuitry.
Applsci 09 02677 g004
Figure 5. SEM images of electrodes before cycling: (a) PVDF@Sb2O3, (b) CMC@Sb2O3, (c) CMC-FA@Sb2O3; SEM images of electrodes after 200 cycles at current density of 200 mAh g−1: (d) PVDF@Sb2O3, (e) CMC@Sb2O3, (f) CMC-FA@Sb2O3.
Figure 5. SEM images of electrodes before cycling: (a) PVDF@Sb2O3, (b) CMC@Sb2O3, (c) CMC-FA@Sb2O3; SEM images of electrodes after 200 cycles at current density of 200 mAh g−1: (d) PVDF@Sb2O3, (e) CMC@Sb2O3, (f) CMC-FA@Sb2O3.
Applsci 09 02677 g005
Figure 6. Schematic illustration of the proposed mechanism of cross-linked CMC-FA binder and PVDF/CMC binder in Sb2O3 electrodes.
Figure 6. Schematic illustration of the proposed mechanism of cross-linked CMC-FA binder and PVDF/CMC binder in Sb2O3 electrodes.
Applsci 09 02677 g006
Table 1. Comparison of the electrochemical performances of CMC-FA@Sb2O3 with previously reported Sb2O3-based composites as anode materials for lithium ion batteries.
Table 1. Comparison of the electrochemical performances of CMC-FA@Sb2O3 with previously reported Sb2O3-based composites as anode materials for lithium ion batteries.
MaterialsCurrent Density (mA/g)Cycle NumberSpecific Capacity (mAh/g)Ref
octahedral Sb2O320050640.8[30]
carbon-coated Sb/Sb2O3100100686[51]
hollow Sb2O3@TiO2100100593[52]
bundle shaped Sb2O320100594.1[25]
3D nest-shaped Sb2O3/RGO composites50100562[53]
CMC-FA@ Sb2O3200200611.4This work

Share and Cite

MDPI and ACS Style

Liu, Y.; Wang, H.; Yang, K.; Yang, Y.; Ma, J.; Pan, K.; Wang, G.; Ren, F.; Pang, H. Enhanced Electrochemical Performance of Sb2O3 as an Anode for Lithium-Ion Batteries by a Stable Cross-Linked Binder. Appl. Sci. 2019, 9, 2677. https://doi.org/10.3390/app9132677

AMA Style

Liu Y, Wang H, Yang K, Yang Y, Ma J, Pan K, Wang G, Ren F, Pang H. Enhanced Electrochemical Performance of Sb2O3 as an Anode for Lithium-Ion Batteries by a Stable Cross-Linked Binder. Applied Sciences. 2019; 9(13):2677. https://doi.org/10.3390/app9132677

Chicago/Turabian Style

Liu, Yong, Haichao Wang, Keke Yang, Yingnan Yang, Junqing Ma, Kunming Pan, Guangxin Wang, Fengzhang Ren, and Huan Pang. 2019. "Enhanced Electrochemical Performance of Sb2O3 as an Anode for Lithium-Ion Batteries by a Stable Cross-Linked Binder" Applied Sciences 9, no. 13: 2677. https://doi.org/10.3390/app9132677

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop