Next Article in Journal
Attapulgite Nanorods Incorporated MXene Lamellar Membranes for Enhanced Decontamination of Dye Wastewater
Next Article in Special Issue
Improvement Strategies for Stability and Efficiency of Perovskite Solar Cells
Previous Article in Journal
High-Magnetic-Sensitivity Magnetoelectric Coupling Origins in a Combination of Anisotropy and Exchange Striction
Previous Article in Special Issue
High-Value Utilization of Silicon Cutting Waste and Excrementum Bombycis to Synthesize Silicon–Carbon Composites as Anode Materials for Li-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sb Nanoparticles Embedded in the N-Doped Carbon Fibers as Binder-Free Anode for Flexible Li-Ion Batteries

1
School of Physics and Electronic Engineering, Jiangsu Normal University, Xuzhou 221116, China
2
Research Institute of Interdisciplinary Science and School of Materials Science and Engineering, Dongguan University of Technology, Dongguan 523820, China
3
College of Materials Science and Engineering, Changsha University of Science and Technology, Changsha 410114, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(18), 3093; https://doi.org/10.3390/nano12183093
Submission received: 11 August 2022 / Revised: 1 September 2022 / Accepted: 3 September 2022 / Published: 6 September 2022

Abstract

:
Antimony (Sb) is considered a promising anode for Li-ion batteries (LIBs) because of its high theoretical specific capacity and safe Li-ion insertion potential; however, the LIBs suffer from dramatic volume variation. The volume expansion results in unstable electrode/electrolyte interphase and active material exfoliation during lithiation and delithiation processes. Designing flexible free-standing electrodes can effectively inhibit the exfoliation of the electrode materials from the current collector. However, the generally adopted methods for preparing flexible free-standing electrodes are complex and high cost. To address these issues, we report the synthesis of a unique Sb nanoparticle@N-doped porous carbon fiber structure as a free-standing electrode via an electrospinning method and surface passivation. Such a hierarchical structure possesses a robust framework with rich voids and a stable solid electrolyte interphase (SEI) film, which can well accommodate the mechanical strain and avoid electrode cracks and pulverization during lithiation/delithiation processes. When evaluated as an anode for LIBs, the as-prepared nanoarchitectures exhibited a high initial reversible capacity (675 mAh g−1) and good cyclability (480 mAh g−1 after 300 cycles at a current density of 400 mA g−1), along with a superior rate capability (420 mA h g−1 at 1 A g−1). This work could offer a simple, effective, and efficient approach to improve flexible and free-standing alloy-based anode materials for high performance Li-ion batteries.

1. Introduction

With the increasing aggravation of environmental pollution and demand for energy conversion and storage, it is critical to develop advanced energy storage devices nowadays [1]. Among the existing energy storage devices, Li-ion batteries (LIBs) have many advantages such as high energy density and high operating voltage; and are extensively used as power sources in various low power electronic devices such as mobile phones, laptops, electric bicycle, and other small power devices [2,3]. However, the commercial LIBs use graphite with low theoretical specific capacity (372 mA⋅h⋅g1) as the anode active materials, which greatly restrict their wide application in electric vehicles and the smart-grid where high power densities are needed. Therefore, it is important to search for alternative anode materials with high specific capacity and good rate capability. So far, various kinds of anode materials such as metal materials [4,5,6,7], semimetal materials [8], transition metal oxides [9,10,11,12,13], metal chalcogenides and transition metal chalcogenides have been studied extensively [14,15,16]. Among them, Sb as a metal material is recognized as a promising anode material for LIBs because of its high theoretical specific capacity (660 mAh g−1), fast charge carrier mobility, low electronegativity, decent working voltage, puckered-layer architecture and good thermodynamic properties [17]. Unfortunately, the bulk Sb electrode is primarily hampered by its gigantic volume change (~150%) during the alloying/dealloying reaction with Li, which would cause structural pulverization, and then result in a rapid capacity fade of the cycling performance for bulk Sb anode materials [18].
It is known that nanomaterials have better mechanical properties and a larger specific area compared with those of corresponding bulk materials [19,20,21]. As a result, when the size of materials reduces to nanoscale, not only can the volume expansion of the active materials in the process of lithium intercalation be inhibited, but also the transmission path of the lithium ion in the electrode material may be shortened [6]. Therefore, reducing the dimension of Sb to nanoscale is an effective strategy to alleviate volume change. For example, Meng He et al. reported that monodisperse Sb nanoparticles showed good capability as anode materials in rechargeable Li-ion after 100 cycles. Nevertheless, the longer-term cycling ability is still limited due to the aggregation of Sb nanoparticles and the unstable SEI film [18]. To address these problems, different kinds of Sb/carbon composites have been synthesized. For example, Sb/carbon nanocomposites were synthesized using high energy ball milling by Glushenkov et al. [22], which exhibited a good reversible capacity of 550 mAh g−1 at a current rate of 230 mA g−1 after 250 cycles. Qian et al. prepared rod-like Sb/C composites by a synchronous reduction and carbon deposition process, which exhibited a reversible capacity of 478.8 mAh g−1 at 100 mA⋅g−1 after 100 cycles [23]. Thus, it is recognized that constructing Sb/C nanoscale composites can successfully improve the cycling ability of the Sb material.
Compared with 3D and 2D materials, the ionic diffusion and electronic transport length can be effectively shortened along their geometry in 1D materials, leading to fast diffusion kinetics. Electrospinning is a feasible technique to prepare various 1D carbon-containing composites. The physical and chemical properties, such as surface-to-volume ratio, mechanical strength, crystallinity, particle size and distribution of metal active materials can be tuned by using different electrospinning precursor solutions and processing parameters [24,25,26]. A series of anode materials have been prepared by electrospinning such as Sn/NPCFs [7], Ge/CNFs [27], Si/C@CNFs and Fe2O3 @CNFs [9,28]. This method can also be adopted to prepare flexible and free-standing films [29,30] as working electrodes of rechargeable batteries without carbon black or binder additives [31], which could improve the volumetric energy and power density of batteries and are beneficial for the overall electrochemical performance.
Considerable efforts have been made to design highly flexible and binder-free electrodes by thermal oxidation [32], electrodeposition [33], vacuum filtration [34], self-assembly [35,36], hydrothermal method [37], and electrospinning [38,39]. However, it should be noted that forming intimate contacts between active materials and freestanding matrices via the construction of a flexible and free-standing electrode for LIBs is still a great challenge. For most freestanding electrodes, the active materials are only physically interconnected with the freestanding matrices without stable interfaces. Consequently, large stress will be generated during the repeated cycling test, resulting in poor cycling stability of the batteries. Therefore, constructing freestanding electrodes with a stable interface is important for the practical application of freestanding electrodes in flexible LIBs.
Herein, for the first time, combining the advantages of Sb nanoparticle and 1D carbon fibers, we report the synthesis of free-standing Sb nanoparticles embedded in 1D flexible N-doped carbon fibers (denoted as Sb@NCFs) using an electrospinning method for LIBs. Owing to the robust hierarchical structure and shortened Li-ion diffusion distance, the as-prepared Sb@NCFs exhibit excellent Li-ion storage performance including high specific capacity, superior rate performance and stable cycling stability.

2. Materials and Methods

2.1. Materials

Polyacrylonitrile (PAN, Mw = 1,500,000) and polyvinylpyrrolidone (PVP, Mw = 1,300,000) were purchased from Sigma-Aldrich (Shanghai) Trading Co., Ltd., Shanghai, China. Diisopropanolamine (DIPA) was purchased from Alfa Aesar (China) Chemicals Co., Ltd., Shanghai, China. N, N-Dimethylformamide (DMF) and antimony (Ⅲ) chloride (SbCl3) were from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. The reagents above were of analytical purity grade and used without further purification.

2.2. Synthesis of Sb@NCFs

The preparation process of Sb@NCFs is shown in Figure 1. The free-standing Sb@NCFs was synthesized by the electrospinning technique as follows: firstly, a certain amount of PAN (0.75 g) and PVP (0.25 g) were dissolved into 10 mL DMF, followed by adding 0.3 mL DIPA. The mixture above was continuously stirred for 12 h at room temperature to obtain a homogeneous solution. Subsequently, the solution was mixed with 0.7 g SbCl3 under continuous vigorous stirring for 3 h. Lastly, the obtained precursor solution was loaded into a 20 mL plastic syringe equipped with a 21# stainless steel needle for electrospinning. A sheet of copper foil on a stainless-steel drum was used as the collector and placed 15 cm away from the tip of the stainless-steel needle. The spinning voltage, the rotating speed of the drum, and the injection flow rate were 15 kV, 60 RPM, and 0.018 mm/min, respectively. Then, the as-spun fibers were placed between the ceramic sheets and annealed at 250 °C for 2 h at a heating rate of 3 °C/min in air. Finally, the as-spun fibers were annealed at 600 °C for 2 h at a heating rate of 2 °C under N2 atmosphere to obtain the Sb@NCFs (denoted as Sb3). To investigate the effect of organic carbon source precursor on the electrochemical properties of Sb@NCFs, the experiment without DIPA in precursor solution was carried out and the as-synthesized sample denoted as Sb2, the sample synthesized without PVP and DIPA in precursor solution (1.0 g PAN) was denoted as Sb1.

2.3. Material Characterizations

The phase and composition of Sb@NCFs were analyzed using X-ray diffraction (XRD, Bruker D8, Karlsruhe, Germany) techniques with Cu-Ka radiation (λ = 0.1542 nm, V = 40 kV, I = 40 mA) at scanning rate of 0.05° s−1 in the range of 10–80°. The morphologies of the Sb@NCFs were characterized using scanning electron microscope (SEM, JSM-6510, JEOL, Tokyo, Japan) and field emission transmission electron microscopy (FETEM, Tecnai G2 20, OR, USA). X-ray photoelectron spectroscopy (XPS, Thermo ESCALAB 250XI XPS, MA, USA) was employed to analyze compositions and chemical states of the elements. The specific surface areas and pore size distributions were recorded using the Brunauer–Emmett–Teller (BET) surface area analysis (Autosorb-IQ), FL, USA. Thermogravimetric analysis (TA) was performed in air from 25–800 °C with a heating rate of 5 °C min−1 by using an TA thermogravimetry analyzer (Q500), Delaware, FL, USA.

2.4. Electrochemical Characterizations

The electrochemical tests were performed on CR2032 coin cells. First, the Sb@NCFs were directly cut into circular electrodes (8 mm in diameter, 0.15~0.45 mg in weight) and used as working electrode. Then, coin cells were assembled in an argon-filled glove box (MBraun, Munich, Germany), of which the working electrode, separator, counter and reference electrode were Sb@NCFs, Celgard 3501 and lithium foil with an electrolyte of 40 μL LiPF6 in a 4:5:1 (v/v) EC/DEC/PC, respectively. After standing for 10–12 h with the filled electrolyte, the assembled cells were connected to the Arbin battery test system (BT2000, Arbin Instruments, College Station, TX, USA). The galvanostatic discharge–charge and rate performance were conducted in the voltage range from 0.01 to 2.8 V at a room temperature. Cyclic voltammetry (CV) curves were obtained on an electrochemical workstation (CHI 760E, Shanghai, China) in the potential range of 0.01–2.8 V (vs. Li+/Li) at a scan rate of 0.1 mV s−1. The electrochemical impedance spectroscopy (EIS) was measured in the frequency range of 10−1 to 106 Hz with an AC voltage amplitude of 5 mV.

3. Results and Discussion

3.1. Characterizations of Sb@NCFs Composite

The crystal structures of the Sb1, Sb2 and Sb3 were characterized using XRD and are shown in Figure 2a. All the diffraction peaks could be indexed to Sb phase (JCPDS 35-0732) [40,41]. The main diffraction peaks of Sb located at 28.69°, 40.077°, and 41.947° corresponded to the (012), (104), and (110) planes. No impurities could be detected, which implied Sb3+ was reduced completely to metallic Sb by carbothermal reduction reaction [42,43]. Moreover, one could see that by adding PVP and DIPA in the precursor, Sb3 exhibited better crystallinity than that of Sb1 and Sb2.
To investigate the morphologies of Sb1, Sb2 and Sb3, the SEM images are presented in Figure 2b and Figure S1a,b. As can be seen in Figure 2b, the Sb3 was composed of long and cross-linked fibers with smooth surfaces and with diameters ranging from 300 to 400 nm. The morphologies of Sb2 and Sb3 were similar to Sb3 (Figure S1a,b). This unique one-dimensional nanostructure can facilitate the rapid transfer of electrons and Li+. As shown in Figure 2c–e, the TEM of Sb3 was performed to check whether the metallic Sb had good contact with N doped carbon fibers. The TEM images displayed smooth surfaces without Sb particles, indicating that all Sb particles were encapsulated in the N-doped carbon fibers without agglomeration, which was consistent with the results of SEM. The conductive N-doped carbon coated on the outside of the metal particles not only served as a support but also suppressed the volume change during insertion/removal of lithium, further increasing the specific capacity of the Sb3. The HRTEM image in Figure 2e shows the lattice fringe of 0.21 nm, which was consistent with the XRD results and corresponded to the (110) plane of Sb with hexagonal phase (JCPDS: 35-0732). To further analyze the element distribution of the Sb, C and N in the Sb3, EDS was carried out. As shown in Figure 2g–j, all the energy-filtered SEM maps of the Sb, C and N matched well with the SEM image, indicating that the Sb, C and N were uniformly dispersed and coexist over the entire composite. These features are beneficial to relieve volume expansion and increase electrical conductivity when used as electrode materials of batteries. The as-spun Sb@NCFs showed excellent flexibility, as demonstrated by the optical image shown in Figure 2f, which could thus be used as binder-free electrode of LIBs directly. The XPS spectra were also employed to distinguish the chemical composition and valence state of Sb3 in Figure 3a. The peaks of Sb 3d, Sb 4d, C 1s and N 1s were observed in the survey scan spectra, indicating the coexistence of Sb, C, and N elements [44]. The C 1s spectrum is given in Figure 3b. The peaks at the binding energy of 284.6, 286.0, 287.0 and 288.2 eV were assigned to the binding energies of graphitized carbon, C-N group, C=C group, and O=C-N group bonds, respectively [45]. The peaks at the binding energy of 398.3 eV, 400.1 eV, 401.2 eV, and 402.7 eV were assigned to pyridine N, pyrrole N, quaternary N, and oxidized N (Figure 3c), respectively [46]. It is beneficial for the insertion of lithium ions because pyridine N and pyrrole N create more open channels and active sites [47]. The binding energies at 530.8 and 540.1eV were assigned to Sb 3d5/2 and Sb 3d3/2 (Figure 3d), respectively [48]. There was a peak at ca. 199 eV, which is the fingerprint peak of Cl [49], meaning that the Cl element from SbCl3 had been completely removed in the carbonization process.
The as-synthesized free-standing films were ground into powders to investigate the pore structures. The N2 adsorption–desorption isotherms of the Sb1, Sb2 and Sb3 were carried out and are presented in Figure S2. The adsorption–desorption isotherms exhibited the characteristics of type-IV isotherm with a H3 hysteresis loop (Figure S2a), indicating the mesoporous structure of the Sb1, Sb2 and Sb3 [41]. The specific surface areas of Sb1, Sb2, Sb3 were 0.632, 5.193, and 24.06 m2 g−1, respectively. The pore-size distributions were calculated according to density functional theory and are shown in Figure S2b, indicating Sb3 had a hierarchical porous structure with abundant mesopores and a small number of micropores centered at 1.4 nm. Correspondingly, the total pore volume of Sb3 was 0.099 cm3 g−1 while those of Sb1 and Sb2 were only 0.022 cm3 g−1 and 0.006 cm3 g−1. The large specific surface area and pore volume of Sb3 will contribute to the electrode/electrolyte contact area, providing sufficient active sites for charge transfer, shortening the diffusion length of the Li ions and finally enhancing the electrochemical performance.
The thermogravimetric (TG) curve of the Sb1, Sb2 and Sb3 (Figure S3, Supporting Information) was conducted to determine the content of Sb in the Sb@NCFs composite. The slight weight loss below 300 °C was mainly associated with the evaporation of water absorbed on the surface of samples. The weight loss between 300 °C and 600 °C can be attributed to the burnt out of carbon and oxidation of Sb particles. The residual part of Sb3 was around 22.9 wt.%, while Sb2 and Sb1 were around 15.5 wt.% and 10.7 wt.% after weight loss. The results showed that the Sb3 had the highest Sb content of all the samples.

3.2. Electrochemical Performance of Sb@NCFs Composite

The Sb@NCF electrode was mechanically flexible, as shown in Figure 1f. The selected first, second, and fifth discharge/charge curves of Sb1, Sb2 and Sb3 between 0.01 and 2.8 V at a current density of 0.1 A g−1 are shown in Figure 4. The first discharge curve of Sb3 (Figure 4c) showed that the voltage dropped rapidly to 0.8 V, which corresponded to the formation of the solid electrolyte interphase (SEI) on the electrode surface [42]. Then, a discernable plateau at about 0.8 V appeared which indicates the formation of a Li3Sb phase during lithium insertion. As the voltage decreased to 0.01 V, the battery exhibited an initial specific capacity of 675.2 mAh g−1, higher than the theoretical specific capacity of Sb (660 mAh g−1), which could be attributed to irreversible Li ion consumption owing to the decomposition of the electrolyte and the formation of SEI films [50]. The first charge curve of Sb3 exhibited a plateau of 1.0 V which was associated with the reversible dealloying process, with a specific capacity of 646.9 mAh g−1. For Sb2, the first discharge/charge curves (Figure 4b) were similar to the curves of Sb1 (Figure 4a). The initial charge and discharge specific capacities of Sb2 and Sb1 were 529.4, 568.7, 474.7 and 496.1 mAh g−1, respectively. Compared with Sb2 and Sb1, Sb3 delivered higher specific capacities which could be ascribed to its higher specific surface area, mesoporous structure, and content of Sb.
The initial three CV curves of Sb3 at a scan rate of 0.1 mV s−1 between 0 and 2.8 V (vs. Li+/Li) are presented in Figure 5a. In the first cathodic sweep, there was a strong cathode peak at about 0.3 V which was attributed to the formation of SEI film [6,42]. The peak did not appear in the second cycle, which reflects the high stability of the SEI film. The cathodic peak was located at 0.8 V in following cycles, which was ascribed to the reaction of metallic Sb to alloyed Li3Sb [6]. Apart from the first cycle, the CV curves overlapped well in the following cycles, demonstrating the good electrochemical reversibility of Sb@NCFs during cycling. Based on the analysis above, the electrochemical process for Sb@NCFs is as follows [51]:
S b + 3 L i + + 3 e L i 3 S b
To further investigate the electrochemical Li-ion storage performance at high current densities, the rate capabilities of Sb1, Sb2 and Sb3 are displayed in Figure 5b. As seen, the Sb3 sample exhibited the reversible capacities of ca. 620, 573.9, 492.5, 459.7, 441.8 and 421.5 mAh g−1 at the current densities of 0.1, 0.2, 0.4, 0.6, 0.8 and 1 A g−1, respectively. It was noted that the discharge capacity of Sb3 returned to 585.3 mAh g−1 when the current density recovered to 0.1 A g−1, suggesting that the structure of Sb3 electrode keeps integrity after high current charge–discharge process [22,52]. The specific capacities (ca. 530.4, 456, 386.4, 344.4, 333.3, and 323.6 mAh g−1) of Sb2 at the current densities of 0.1, 0.2, 0.4, 0.6, 0.8, and 1 A g−1 were lower than those of Sb3 without DIPA, owing to the smaller pore size and specific surface area. For Sb1 without DIPA and PVP, the specific capacities (ca. 496.1, 220.6, 180.9, 149.2, 125.5, and 104.3 mAh g−1) were even lower than those of Sb2 and Sb3. When the current density returned to 0.1 A g−1, the discharge capacity was only 460.7 mAh g−1 for Sb2 and 243.1 mAh g−1 for Sb1. The cycling stability of the Sb@NCFs electrode at a current density of 0.1 and 0.4 A g−1 is shown in Figure 5c. The discharge capacity decayed to 629 mAh g−1 after the first cycle and a high reversible discharge capacity of ca. 590 mAh g−1 could still be retained with a high coulombic efficiency after 300 cycles. Even at a current density of 0.4 A g−1, the Sb@NCFs electrode still maintained a high specific charge capacity of 480 mAh g−1 electrode after 300 cycles.
To further clarify the electrochemical performance of the Sb@NCFs, EIS measurements were performed after the 10 cycles (Figure 6). All Nyquist plots can be divided into three parts, the Z’-axis intercepts, the semicircle, and the oblique line in the high and low-frequency region, which represent the resistance of electrolyte (Re), the resistance of charge transfer (Rct), and the Warburg (Zw) impedance related to the lithium-ion diffusion (DLi), respectively [23,44,53,54]. The Nyquist plots were fitted with the equivalent electric circuit and the values of the Re, Rct and Zw are shown in the inset table of Figure 6. The diameter of the high-frequency semicircle revealed that the Sb3 had the lowest Rct value, which indicates the reduction of the unfavorable side reactions on the electrode/electrolyte interface for Sb3. The slope value of the low-frequency oblique line showed a trend of Sb3 > Sb2 > Sb1, which indicated that the Sb3 had the largest DLi and fastest lithium intercalation/deintercalation, since Sb3 had the highest conductivity and porosity. It could thus be concluded that Sb3 had the lowest impedance (including Re, Rct, and Zw) at the electrode/electrolyte interface, leading to high specific capacity retention and a superior rate performance [55,56,57,58].
Compared with the results reported in the literature, our Sb@NCF composite exhibited better electrochemical properties, as listed in Table 1 [6,22,23,42,44,52,59,60]. Considering all the above results, the Sb@NCF composite is more suitable for anode material than other developed materials for future LIBs. First, there is no need to add conductive agents, binders and collectors in the electrode preparation process. The flexible free-standing electrode is cost effective with excellent electrochemical performance. Secondly, the carbon matrix can prevent Sb nanoparticles from agglomerating into bulk particles and suppressing the volume change during repeated charging and discharging processes. Finally, the porous structure can provide rich electrolyte penetration channels and additional space to relieve structural decomposition/extraction caused by lithium-ion insertion/removal. Thus, the rate capability and cycling stability is significantly improved.

4. Conclusions

In summary, a highly flexible Sb@NCF composite with excellent electrochemical performance was prepared by the electrospinning method. The Sb nanoparticles were uniformly dispersed in the carbon matrix which could provide a stable porous structure for the rapid charge transfer and Li-ion intercalation/deintercalation. When used as a free-standing electrode of the LIB anode, the unique Sb@NCF electrode delivered a high reversible capacity of 675 mAh g−1 at the current density of 100 mA g−1 and excellent capacity retention rates of 88 and 87.6% at a current density rate of 100 and 400 mA g−1 after 300 cycles, respectively. The new flexible and binder-free electrode enables the practical applications of Sb-based anodes in high energy density LIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12183093/s1, Figure S1: SEM images of (a) Sb1 and (b) Sb2; Figure S2. (a) N2 adsorption-desorption isotherms and their as-calculated structural parameters in insert. (b) The pore size distribution of Sb1, Sb2, Sb3; Figure S3. TGA curves of Sb1, Sb2 and Sb3.

Author Contributions

Methodology, X.W., N.J. and J.L.; investigation and formal analysis, X.W., P.L., X.Z., Y.L. and W.Q.; writing—original draft preparation, X.W., N.J., J.L. and C.S.; writing—review and editing, N.J., J.L., W.Q. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Jiangsu Province Science and Technology Project, China (Grant No: BE2020759). Xuzhou Science and Technology Project, China (Grant No: KC19071), Natural Science Foundation of Hunan Province, China (Grant No: 2020JJ5580).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The authors claim that none of the material in our manuscript has been published or is under consideration for publication elsewhere. This manuscript is our original work.

References

  1. Qin, W.; Zhou, N.; Wu, C.; Xie, M.; Sun, H.; Guo, Y.; Pan, L. Mini-review on the redox additives in aqueous electrolyte for high performance supercapacitors. ACS Omega 2020, 5, 3801–3808. [Google Scholar] [CrossRef]
  2. Divakaran, A.M.; Hamilton, D.; Manjunatha, K.N.; Minakshi, M. Design, development and thermal analysis of reusable Li- ion battery module for future mobile and stationary applications. Energies 2020, 13, 1477. [Google Scholar] [CrossRef]
  3. Liu, J.; Wang, M.; Wang, Q.; Zhao, X.; Song, Y.; Zhao, T.; Sun, J. Sea urchin-like Si@MnO2@rGO as anodes for high-performance lithium-ion batteries. Nanomaterials 2022, 12, 285. [Google Scholar] [CrossRef] [PubMed]
  4. Xin, F.; Whittingham, M.S. Challenges and development of tin-based anode with high volumetric capacity for Li-ion batteries Electrochem. Energy R. 2020, 3, 643–655. [Google Scholar]
  5. Gulzar, U.; Li, T.; Bai, X.; Goriparti, S.; Brescia, R.; Capiglia, C.; Zaccaria, R.P. Nitrogen-doped single walled carbon nanohorns enabling effective utilization of Ge nanocrystals for next generation lithium ion batteries. Electrochim. Acta 2019, 298, 89–96. [Google Scholar] [CrossRef]
  6. Yi, Z.; Han, Q.G.; Zan, P.; Wu, Y.M.; Cheng, Y.; Wang, L.M. Sb nanoparticles encapsulated into porous carbon matrixes for high-performance lithium-ion battery anodes. J. Power Sources 2016, 331, 16–21. [Google Scholar] [CrossRef]
  7. Xu, Y.; Yuan, T.; Bian, Z.; Yang, J.; Zheng, S. Tuning particle and phase formation of Sn/carbon nanofibers composite towards stable lithium-ion storage. J. Power Sources 2020, 453, 227467. [Google Scholar] [CrossRef]
  8. Shi, Q.; Zhou, J.; Ullah, S.; Yang, X.; Tokarska, K.; Trzebicka, B.; Ta, H.Q.; Rümmeli, M.H. A review of recent developments in Si/C composite materials for Li-ion batteries. Energy Storage Mater. 2021, 34, 735–754. [Google Scholar] [CrossRef]
  9. Jayaraman, S.; Aravindan, V.; Ulaganathan, M.; Ling, W.C.; Ramakrishna, S.; Madhavi, S. Ultralong durability of porous α-Fe2O3 nanofibers in practical Li-ion configuration with LiMn2O4 cathode. Adv. Sci. 2015, 2, 1500050. [Google Scholar] [CrossRef] [PubMed]
  10. Kitta, M.; Kohyama, M. Nanoscale controlled Li-insertion reaction induced by scanning electron-beam irradiation in a Li4Ti5O12 electrode material for lithium-ion batteries. Phys. Chem. Chem. Phys. 2017, 19, 11581–11587. [Google Scholar] [CrossRef] [PubMed]
  11. Hu, X.; Zeng, G.; Chen, J.; Lu, C.; Wen, Z. 3D graphene network encapsulating SnO2 hollow spheres as a high-performance anode material for lithium-ion batteries. J. Mater. Chem. A 2017, 5, 4535–4542. [Google Scholar] [CrossRef]
  12. Divakaran, A.M.; Minakshi, M.; Bahri, P.A.; Paul, S.; Kumari, P.; Divakaran, A.M.; Manjunatha, K.N. Rational design on materials for developing next generation lithium-ion secondary battery. Prog. Solid State Chem. 2021, 62, 100298. [Google Scholar] [CrossRef]
  13. Solmaz, R.; Karahan, B.D. Effect of vinylene carbonate as electrolyte additive for Mn2O3/NiMnO3 anodes of lithium-ion batteries. Ionics 2021, 27, 2813–2824. [Google Scholar] [CrossRef]
  14. Miao, Y.E.; Huang, Y.P.; Zhang, L.; Fan, W.; Lai, F.L.; Liu, T.X. Electrospun porous carbon nanofiber@MoS2 core/sheath fiber membranes as highly flexible and binder-free anodes for lithium-ion batteries. Nanoscale 2015, 7, 11093–11101. [Google Scholar] [CrossRef] [PubMed]
  15. Lu, J.; Lian, F.; Guan, L.; Zhang, Y.; Ding, F. Adapting FeS2 micron particles as an electrode material for lithium-ion batteries via simultaneous construction of CNT internal networks and external cages. J. Mater. Chem. A 2019, 7, 991–997. [Google Scholar] [CrossRef]
  16. Xiao, X.; Ni, L.; Chen, G.; Ai, G.; Li, J.; Qiu, T.; Liu, X. Two-dimensional NiSe2 nanosheets on carbon fiber cloth for high-performance lithium-ion batteries. J. Alloy. Compd. 2020, 821, 153218. [Google Scholar] [CrossRef]
  17. Ge, X.; Liu, S.; Qiao, M.; Du, Y.; Li, Y.; Bao, J.; Zhou, X. Enabling superior electrochemical properties for highly efficient po tassium storage by impregnating ultrafine Sb nanocrystals within nanochannel-containing carbon nanofibers. Angew Chem. Int. Ed. Engl. 2019, 58, 14578–14583. [Google Scholar] [CrossRef] [PubMed]
  18. He, M.; Kravchyk, K.; Walter, M.; Kovalenko, M.V. Monodisperse antimony nanocrystals for high-rate Li-ion and Na-ion battery anodes: Nano versus bulk. Nano Lett. 2014, 14, 1255–1262. [Google Scholar] [CrossRef] [PubMed]
  19. Zhao, Q.; Zhao, M.; Qiu, J.; Lai, W.Y.; Pang, H.; Huang, W. One dimensional silver-based nanomaterials: Preparations and electrochemical applications. Small 2017, 13, 1701091. [Google Scholar] [CrossRef]
  20. David, M.E.; Ion, R.M.; Grigorescu, R.M.; Iancu, L.; Andrei, E.R. Nanomaterials used in conservation and restoration of cultural heritage: An up-to-date overview. Materials 2020, 13, 2064. [Google Scholar] [CrossRef]
  21. Li, J.; Zhao, X.; Li, D.; Zheng, X.; Yang, X.; Chou, S.; Zhu, Z. Effect of size and dimensionality on the band gap and conductivity of InAs, PbS, Ge, and Bi2S3 nanostructured semiconductors. Curr. Nanosci. 2016, 12, 324–329. [Google Scholar] [CrossRef]
  22. Ramireddy, T.; Rahman, M.M.; Xing, T.; Chen, Y.; Glushenkov, A.M. Stable anode performance of an Sb-carbon nanocomposite in lithium-ion batteries and the effect of ball milling mode in the course of its preparation. J. Mater. Chem. A 2014, 2, 4282–4291. [Google Scholar] [CrossRef]
  23. Fan, L.; Zhang, J.J.; Cui, J.H.; Zhu, Y.C.; Liang, J.W.; Wang, L.L.; Qian, Y.T. Electrochemical performance of rod-like Sb-C composite as anodes for Li-ion and Na-ion batteries. J. Mater. Chem. A 2015, 3, 3276–3280. [Google Scholar] [CrossRef]
  24. Zhang, L.H.; Qin, X.Y.; Zhao, S.Q.; Wang, A.; Luo, J.; Wang, Z.L.; Kang, F.Y.; Lin, Z.Q.; Li, B.H. Advanced matrixes for binder-free nanostructured electrodes in lithium-ion batteries. Adv. Mater. 2020, 32, 1908445. [Google Scholar] [CrossRef]
  25. Wang, J.; Wang, Z.; Ni, J.; Li, L. Electrospun materials for batteries moving beyond lithium-ion technologies. Electrochem. Energy R. 2021, 5, 211–241. [Google Scholar] [CrossRef]
  26. Li, Y.; Li, Q.; Tan, Z. A review of electrospun nanofiber-based separators for rechargeable lithium-ion batteries. J. Power Sources 2019, 443, 227262. [Google Scholar] [CrossRef]
  27. Li, W.H.; Yang, Z.Z.; Cheng, J.X.; Zhong, X.W.; Gu, L.; Yu, Y. Germanium nanoparticles encapsulated in flexible carbon nanofibers as self-supported electrodes for high performance lithium-ion batteries. Nanoscale 2014, 6, 4532–4537. [Google Scholar] [CrossRef]
  28. Xu, Y.; Yuan, T.; Bian, Z.H.; Sun, H.; Pang, Y.P.; Peng, C.X.; Yang, J.H.; Zheng, S.Y. Electrospun flexible Si/C@CNF nonwoven anode for high capacity and durable lithium-ion battery. Compos. Commun. 2019, 11, 1–5. [Google Scholar] [CrossRef]
  29. Yang, M.; Liu, L.; Yan, H.; Zhang, W.; Su, D.; Wen, J.; Liu, W.; Yuan, Y.; Liu, J.; Wang, X. Porous nitrogen-doped Sn/C film as free-standing anodes for lithium ion batteries. Appl. Surf. Sci. 2021, 551, 149246. [Google Scholar] [CrossRef]
  30. Zhang, T.; Qiu, D.; Hou, Y. Free-standing and consecutive ZnSe@carbon nanofibers architectures as ultra-long lifespan anode for flexible lithium-ion batteries. Nano Energy 2022, 94, 106909. [Google Scholar] [CrossRef]
  31. Fei, L.; Williams, B.P.; Yoo, S.H.; Carlin, J.M.; Joo, Y.L. A general approach to fabricate free-standing metal sulfide@carbon nanofiber networks as lithium ion battery anodes. Chem. Commun. 2016, 52, 1501–1504. [Google Scholar] [CrossRef]
  32. Xue, X.Y.; Xing, L.L.; Chen, Y.J.; Shi, S.L.; Wang, Y.G.; Wang, T.H. Synthesis and H2S sensing properties of CuO-SnO2 core/shell PN-junction nanorods. J. Phys.Chem. C. 2008, 112, 12157–12160. [Google Scholar] [CrossRef]
  33. Fu, J.Z.; Liu, H.; Liao, L.B.; Fan, P.; Wang, Z.; Wu, Y.Y.; Zhang, Z.W.; Hai, Y.; Lv, G.C.; Mei, L.F.; et al. Ultrathin Si/CNTspaper-like composite for flexible Li-ion battery anode with high volumetric capacity. Front. Chem. 2018, 6, 624. [Google Scholar] [CrossRef]
  34. Köse, H.; Aydin, A.O.; Dombaycioğlu, Ş. Graphene-based architectures of tin and zinc oxide nanocomposites for free-standing binder-free Li-ion anodes. Int. J. Energy Res. 2018, 42, 4710–4718. [Google Scholar] [CrossRef]
  35. Liao, H.-S.; Lin, J.; Liu, Y.; Huang, P.; Jin, A.; Chen, X.Y. Self-assembly mechanisms of nanofibers from peptide amphiphiles in solution and on substrate surfaces. Nanoscale 2016, 8, 14814–14820. [Google Scholar] [CrossRef]
  36. Rolandi, M.; Rolandi, R. Self-assembled chitin nanofibers and applications. Adv. Colloid Interface 2014, 207, 216–222. [Google Scholar] [CrossRef]
  37. Jessl, S.; Copic, D.; Engelke, S.; Ahmad, S.; De Volder, M. Hydrothermal coating of patterned carbon nanotube forest for structured lithium-ion battery electrodes. Small 2019, 15, e1901201. [Google Scholar] [CrossRef] [PubMed]
  38. Zhao, Y.; He, X.; Li, J.; Gao, X.; Jia, J. Porous CuO/SnO2 composite nanofibers fabricated by electrospinning and their H2S sensing properties. Sens. Actuators B-Chem. 2012, 165, 82–87. [Google Scholar] [CrossRef]
  39. Lu, X.F.; Wang, C.; Favier, F.; Pinna, N. Electrospun nanomaterials for supercapacitor electrodes: Designed architectures and electrochemical performance. Adv. Energy Mater. 2017, 7, 1601301. [Google Scholar] [CrossRef]
  40. Duan, J.; Zhang, W.; Wu, C.; Fan, Q.G.; Zhang, W.X.; Hu, X.L.; Huang, Y.H. Self-wrapped Sb/C nanocomposite as anode material for high-performance sodium-ion batteries. Nano Energy 2015, 16, 479–487. [Google Scholar] [CrossRef]
  41. Zhong, L.F.; Tang, A.D.; Wen, X.; Yan, P.; Wang, J.J.; Tan, L.; Chen, J. New finding on Sb (2–3 nm) nanoparticles and carbon simultaneous anchored on the porous palygorskite with enhanced catalytic activity. J. Alloys Compd. 2018, 743, 394–402. [Google Scholar] [CrossRef]
  42. Lv, H.L.; Qiu, S.; Lu, G.X.; Fu, Y.; Li, X.Y.; Hu, C.X.; Liu, J.R. Nanostructured antimony/carbon composite fibers as anode material for lithium-ion battery. Electrochim. Acta 2015, 151, 214–221. [Google Scholar] [CrossRef]
  43. Zheng, J.; Yang, Y.; Fan, X.L.; Ji, G.B.; Ji, X.; Wang, H.Y.; Hou, S.; Zachariah, M.R.; Wang, C.S. Extremely stable antimony-carbon composite anodes for potassium-ion batteries. Energ. Environ. Sci. 2019, 12, 615–623. [Google Scholar] [CrossRef]
  44. Yi, Z.; Han, Q.G.; Cheng, Y.; Wang, F.X.; Wu, Y.M.; Wang, L.M. A novel strategy to prepare Sb thin film sandwiched between the reduced graphene oxide and Ni foam as binder-free anode material for lithium-ion batteries. Electrochim. Acta 2016, 190, 804–810. [Google Scholar] [CrossRef]
  45. Wang, M.G.; Han, J.; Hu, Y.M.; Guo, R. Mesoporous C, N-codoped TiO2 hybrid shells with enhanced visible light photocatalytic performance. RSC Adv. 2017, 7, 15513–15520. [Google Scholar] [CrossRef]
  46. Yang, F.H.; Zhang, Z.A.; Du, K.; Zhao, X.X.; Chen, W.; Lai, Y.Q.; Li, J. Dopamine derived nitrogen-doped carbon sheets as anode materials for high-performance sodium ion batteries. Carbon 2015, 91, 88–95. [Google Scholar] [CrossRef]
  47. Liu, Y.C.; Zhang, N.; Jiao, L.F.; Chen, J. Tin Nanodots encapsulated in porous nitrogen-doped carbon nanofibers as a free-standing anode for avanced sodium-ion batteries. Adv Mater. 2015, 27, 6702–6707. [Google Scholar] [CrossRef]
  48. Gurgul, J.; Rinke, M.T.; Schellenberg, I.; Pöttgen, R. The antimonide oxides REZnSbO and REMnSbO (RE = Ce, Pr)—An XPS study. Solid State Sci. 2013, 17, 122–127. [Google Scholar] [CrossRef]
  49. Schaff, J.E.; Roberts, J.T. Adsorbed states of acetonitrile and chloroform on amorphous and crystalline ice studied with X-ray photoelectron spectroscopy. Surf. Sci. 1999, 426, 384–394. [Google Scholar] [CrossRef]
  50. Wan, F.; Guo, J.Z.; Zhang, X.H.; Zhang, J.P.; Sun, H.Z.; Yan, Q.; Han, D.X.; Niu, L.; Wu, X.L. In situ binding Sb nanospheres on graphene via oxygen bonds as superior anode for ultrafast sodium-ion batteries. ACS Appl. Mater. Interfaces 2016, 8, 7790–7799. [Google Scholar] [CrossRef] [PubMed]
  51. Yang, Y.C.; Yang, X.M.; Zhang, Y.; Hou, H.S.; Jing, M.J.; Zhu, Y.R.; Fang, L.B.; Chen, Q.Y.; Ji, X.B. Cathodically induced antimony for rechargeable Li-ion and Na-ion batteries: The influences of hexagonal and amorphous phase. J. Power Sources 2015, 282, 358–367. [Google Scholar] [CrossRef]
  52. He, X.M.; Pu, W.H.; Wang, L.; Ren, J.G.; Jiang, C.Y.; Wan, C.R. Synthesis of nano Sb-encapsulated pyrolytic polyacrylonitrile composite for anode material in lithium secondary batteries. Electrochim. Acta 2007, 52, 3651–3653. [Google Scholar] [CrossRef]
  53. Ghomi, J.S.; Akbarzadeh, Z. Ultrasonic accelerated Knoevenagel condensation by magnetically recoverable MgFe2O4 nanocatalyst: A rapid and green synthesis of coumarins under solvent-free conditions. Ultrason Sonochem 2018, 40, 78–83. [Google Scholar] [CrossRef] [PubMed]
  54. Huang, L.; Cheng, J.L.; Li, X.D.; Yuan, D.M.; Ni, W.; Qu, G.X.; Guan, Q.; Zhang, Y.; Wang, B. Sulfur quantum dots wrapped by conductive polymer shell with internal void spaces for high-performance lithium–sulfur batteries. J. Mater. Chem. A 2015, 3, 4049–4057. [Google Scholar] [CrossRef]
  55. Xu, Y.H.; Zhu, Y.J.; Liu, Y.H.; Wang, C.S. Electrochemical performance of porous carbon/tin composite anodes for sodium-ion and lithium-ion batteries. Adv. Energy Mater. 2013, 3, 128–133. [Google Scholar] [CrossRef]
  56. Wu, N.T.; Zhang, Y.; Guo, Y.; Liu, S.J.; Liu, H.; Wu, H. flakelike LiCoO2 with exposed {010} facets as a stable cathode material for highly reversible lithium storage. ACS Appl. Mater. Interfaces 2016, 8, 2723–2731. [Google Scholar] [CrossRef]
  57. Wu, L.; Pei, F.; Mao, R.J.; Wu, F.Y.; Wu, Y.; Qian, J.F.; Cao, Y.L.; Ai, X.P.; Yang, H.X. SiC-Sb-C nanocomposites as high-capacity and cycling-stable anode for sodium-ion batteries. Electrochim. Acta 2013, 87, 41–45. [Google Scholar] [CrossRef]
  58. Wang, H.Q.; Pan, Q.C.; Wu, Q.; Zhang, X.H.; Huang, Y.G.; Lushington, A.; Li, Q.Y.; Sun, X.L. Ultrasmall MoS2 embedded in carbon nanosheets coated Sn/SnOx as anode material for high-rate and long life Li-ion batteries. J. Mater. Chem. A 2017, 5, 4576–4582. [Google Scholar] [CrossRef]
  59. Chen, W.X.; Lee, J.Y.; Liu, Z.L. The nanocomposites of carbon nanotube with Sb and SnSb0.5 as Li-ion battery anodes. Carbon 2003, 41, 959–966. [Google Scholar] [CrossRef]
  60. NuLi, Y.N.; Yang, J.; Jiang, M.S. Synthesis and characterization of Sb/CNT and Bi/CNT composites as anode materials for lithium-ion batteries. Mater. Lett. 2008, 62, 2092–2095. [Google Scholar] [CrossRef]
Figure 1. Synthetic illustration of the synthetic process of Sb@NCFs.
Figure 1. Synthetic illustration of the synthetic process of Sb@NCFs.
Nanomaterials 12 03093 g001
Figure 2. (a) XRD patterns of Sb1, Sb2, Sb3, (b) SEM images of Sb3, (c,d) TEM image of Sb3, (e) HRTEM image of Sb3, (f) optical images of flexible Sb@NCFs, and (g) HAADF-STEM image of Sb3 and EDS mappings of (h) Sb, (i) C and (j) N.
Figure 2. (a) XRD patterns of Sb1, Sb2, Sb3, (b) SEM images of Sb3, (c,d) TEM image of Sb3, (e) HRTEM image of Sb3, (f) optical images of flexible Sb@NCFs, and (g) HAADF-STEM image of Sb3 and EDS mappings of (h) Sb, (i) C and (j) N.
Nanomaterials 12 03093 g002
Figure 3. XPS spectra of Sb3: (a) survey scan, (b) C1s, (c) N1s, (d) Sb3d.
Figure 3. XPS spectra of Sb3: (a) survey scan, (b) C1s, (c) N1s, (d) Sb3d.
Nanomaterials 12 03093 g003
Figure 4. Galvanostatic discharge/charge curves of the 1st, 2nd, 5th cycles for (a) Sb1, (b) Sb2 and (c) Sb3 at a current density of 100 mA g−1.
Figure 4. Galvanostatic discharge/charge curves of the 1st, 2nd, 5th cycles for (a) Sb1, (b) Sb2 and (c) Sb3 at a current density of 100 mA g−1.
Nanomaterials 12 03093 g004
Figure 5. (a) CV curves of Sb3, (b) rate performances of Sb1, Sb2, Sb3, galvanostatic discharge–charge cycles of Sb3 at a current density of (c) 0.1 Ag−1 and 0.4 Ag−1.
Figure 5. (a) CV curves of Sb3, (b) rate performances of Sb1, Sb2, Sb3, galvanostatic discharge–charge cycles of Sb3 at a current density of (c) 0.1 Ag−1 and 0.4 Ag−1.
Nanomaterials 12 03093 g005
Figure 6. Nyquist plots of the Sb1, Sb2 and Sb3.
Figure 6. Nyquist plots of the Sb1, Sb2 and Sb3.
Nanomaterials 12 03093 g006
Table 1. Comparison of the electrochemical performances of the prepared Sb@NCF composites with previously reported Sb-C-based composites.
Table 1. Comparison of the electrochemical performances of the prepared Sb@NCF composites with previously reported Sb-C-based composites.
MaterialsFree-Standing (Yes or No)Current Density (A g−1)Cycle NumbersCapacity (mAh g−1)References
Sb/CNTNo0.0530287[59]
Sb@CNo0.320408[52]
Sb/CNo0.23250550[22]
1.15-400
Sb/CNTNo0.2550277.4[60]
Sb/CNo0.1100315.9[42]
0.8-246.2
Sb@CNo0.1100478.8[23]
0.5-369.7
Sb/graphiteYes0.150424.1[44]
1-158
Sb/CNo0.2100565[6]
1500400.5
5-315.4
Sb@NCFsYes0.1300590This work
0.4300480
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, X.; Jia, N.; Li, J.; Liu, P.; Zhao, X.; Lin, Y.; Sun, C.; Qin, W. Sb Nanoparticles Embedded in the N-Doped Carbon Fibers as Binder-Free Anode for Flexible Li-Ion Batteries. Nanomaterials 2022, 12, 3093. https://doi.org/10.3390/nano12183093

AMA Style

Wang X, Jia N, Li J, Liu P, Zhao X, Lin Y, Sun C, Qin W. Sb Nanoparticles Embedded in the N-Doped Carbon Fibers as Binder-Free Anode for Flexible Li-Ion Batteries. Nanomaterials. 2022; 12(18):3093. https://doi.org/10.3390/nano12183093

Chicago/Turabian Style

Wang, Xin, Nanjun Jia, Jianwei Li, Pengbo Liu, Xinsheng Zhao, Yuxiao Lin, Changqing Sun, and Wei Qin. 2022. "Sb Nanoparticles Embedded in the N-Doped Carbon Fibers as Binder-Free Anode for Flexible Li-Ion Batteries" Nanomaterials 12, no. 18: 3093. https://doi.org/10.3390/nano12183093

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