留言板

尊敬的读者、作者、审稿人, 关于本刊的投稿、审稿、编辑和出版的任何问题, 您可以本页添加留言。我们将尽快给您答复。谢谢您的支持!

姓名
邮箱
手机号码
标题
留言内容
验证码
Qiong Jiang, Wen-qi Zhang, Jia-chang Zhao, Pin-hua Rao, and Jian-feng Mao, Superior sodium and lithium storage in strongly coupled amorphous Sb2S3 spheres and carbon nanotubes, Int. J. Miner. Metall. Mater., 28(2021), No. 7, pp. 1194-1203. https://doi.org/10.1007/s12613-021-2259-5
Cite this article as:
Qiong Jiang, Wen-qi Zhang, Jia-chang Zhao, Pin-hua Rao, and Jian-feng Mao, Superior sodium and lithium storage in strongly coupled amorphous Sb2S3 spheres and carbon nanotubes, Int. J. Miner. Metall. Mater., 28(2021), No. 7, pp. 1194-1203. https://doi.org/10.1007/s12613-021-2259-5
引用本文 PDF XML SpringerLink
研究论文

强耦合非晶态Sb2S3球与碳纳米管中的钠离子和锂离子存储

  • Research Article

    Superior sodium and lithium storage in strongly coupled amorphous Sb2S3 spheres and carbon nanotubes

    + Author Affiliations
    • A facile one-step strategy involving the reaction of antimony chloride with thioacetamide at room temperature is successfully developed for the synthesis of strongly coupled amorphous Sb2S3 spheres and carbon nanotubes (CNTs). Benefiting from the unique amorphous structure and its strongly coupled effect with the conductive network of CNTs, this hybrid electrode (Sb2S3@CNTs) exhibits remarkable sodium and lithium storage properties with high capacity, good cyclability, and prominent rate capability. For sodium storage, a high capacity of 814 mAh·g−1 at 50 mA·g−1 is delivered by the electrode, and a capacity of 732 mAh·g−1 can still be obtained after 110 cycles. Even up to 2000 mA·g−1, a specific capacity of 584 mAh·g1 can be achieved. For lithium storage, the electrode exhibits high capacities of 1136 and 704 mAh·g−1 at 100 and 2000 mA·g−1, respectively. Moreover, the cell holds a capacity of 1104 mAh·g−1 under 100 mA·g−1 over 110 cycles. Simple preparation and remarkable electrochemical properties make the Sb2S3@CNTs electrode a promising anode for both sodium-ion (SIBs) and lithium-ion batteries (LIBs).

    • The main area of application of lithium-ion batteries (LIBs) is the power supply of portable electronics. In this application, LIBs have been successfully commercialized. Recently, attention has been growing in the areas of LIB application in energy storage power stations, smart power grids, electric vehicles/hybrid electric vehicles, and other fields [14]. The lack of lithium, however, has caused the price of the battery materials to soar, limiting the wide application of LIBs. In contrast, sodium is widely available and relatively inexpensive; this has accelerated research into sodium-ion batteries (SIBs) [57]. More importantly, the chemical characteristics of sodium are similar to those of lithium; hence SIBs exhibit similar energy storage mechanisms to LIBs. Therefore, SIBs can be a more economical alternative to LIBs, especially in large-scale applications. However, an intrinsic drawback of SIBs is that the ionic radius of Na+ is larger than that of Li+ (0.098 nm vs. 0.069 nm), leading to slower diffusion kinetics and a greater tendency to cause volumetric strain in the host materials [8]. Therefore, it is challenging but desirable to develop suitably performing electrode materials capable of reversibly accommodating the relatively large Na+.

      As is known, graphite has become the preferred anode material for commercial LIBs, thanks to its abundant resources and low cost. However, compared with LIBs, sodium ions can only be minimally intercalated into the commercial graphite, and the reversible capacity of graphite in SIBs is very small [910]. The reversible capacity of other carbonaceous materials, such as hard carbon, is close to that of graphite in LIBs in their use as the SIB anode, at 300 mAh·g−1 [11]. However, the relatively low specific capacity of ~300 mAh·g−1 and security issues resulting from the formation of dendrites in the charge–discharge cycle has prompted researchers to search for higher capacity and safer anode materials [12].

      In the process of finding better anodes, owing to their better safety features and higher capacities, different materials such as alloys/metals [1317], metal sulfides [1821], and phosphorus/phosphides [2224], have attracted tremendous interest. One focus is on metal sulfides (MxS), which present high charge/discharge capacities, as some of them will have not only the conversion reaction between Li or Na and MxS, but also the alloying reaction of surplus metals with Li or Na. For example, owing to its unique laminar structure, Sb2S3 has been developed as a promising host for storing lithium and sodium [12,2531]. The mechanisms of conversion and alloying processes of Sb2S3 with Li and Na can be expressed using Eqs. (1) and (2):

      $$\begin{aligned}[b] &{\rm{Conversion}}: \;\frac{1}{2}{\rm{S}}{{\rm{b}}_2}{{\rm{S}}_3} + 3{{\rm{M}}^ + } + 3{{\rm{e}}^ - } \to\\ & \quad {\rm{Sb}} + \frac{3}{2}{{\rm{M}}_2}{\rm{S}}\;\;\;\;\;\;\;\left( {{\rm{M}} = {\rm{Li}},\;{\rm{Na}}} \right)\end{aligned} $$ (1)
      $$ {\rm{Alloy}}:\;{\rm{Sb}} + 3{{\rm{M}}^ + } + 3{{\rm{e}}^ - } \to {{\rm{M}}_3}{\rm{Sb}}\;\;\;\;\;\;\;\;\;\;\left( {{\rm{M}} = {\rm{Li}},\;{\rm{Na}}} \right) $$ (2)

      There are 3 mol of electron transfer per mol of Sb in Eq. (1), whose theoretical capacity is 473 mAh·g−1. The process of Eq. (2) generates an added capacity of 473 mAh·g−1 (3 mol electron transfer), leading to a total capacity of 946 mAh·g−1. However, during charge and discharge, capacity degradation is often observed in this material resulting from large volume expansion/contraction [2728].

      To address this problem, researchers attempted to engineer Sb2S3 by preparing and utilizing the nano-sizes, nanomorphologies, and carbon-based complex materials to restrain volume expansion [12,2531]. In particular, amorphous materials are expected to become a promising alternative solution for high property anodic materials. They offer high capacity owing to their strength, affording a higher interfacial area, high rate performance through the acceleration of diffusion of ions, and stable cycling reaction by adapting to the strain produced in the cyclic process. For example, Park et al. [26] synthesized amorphous Sb2S3/C composites by high-energy mechanical milling, in which amorphous Sb2S3 was evenly dispersed in the amorphous carbon matrix. Sb2S3/C exhibits remarkable electrochemical performance when applied to the LIB anode, including high energy density (first charge: 757 mAh·g−1), long cycling life (about 600 mAh·g−1 over 100 cycles), and high first-cycle coulombic efficiency (about 85%). More recently, Zhao and Manthiram [31] designed an electrode structure with amorphous Sb2S3 embedded in the conductive graphite matrix through a high-energy mechanical milling process. The Sb2S3–graphite electrode shows prominent properties in SIBs, including high charge capacity, outstanding rate capability, and a remarkable, stable cycle property. The existence of the active phase with an amorphous structure and the conductive carbon matrix makes it possible to have excellent properties. In spite of these advantages, there are still some problems in the complex synthesis of amorphous materials, such as long synthesis time, difficult synthesis, and high cost [26]. For example, using the methods of high energy ball milling, long ball milling times (typically >10 h) are necessary to form an amorphous structure. This is energy consuming, especially for batch milling of large quantities of materials.

      To alleviate the problem of Sb2S3, we have designed a hybrid material (denoted as Sb2S3@CNTs) with a strong coupling of Sb2S3 and carbon nanotubes (CNTs), consisting of amorphous Sb2S3 spheres that are tightly connected to carbon nanotubes. Consequently, the synthesized Sb2S3@CNTs hybrid reveals superb Na and Li-storage capabilities, reserving 732 mAh·g−1 at 50 mA·g−1 and 1104 mAh·g−1 at 100 mA·g−1 over 100 cycles for SIBs and LIBs, respectively, thus showing its potential application in advanced rechargeable batteries.

      The chemicals antimony chloride (SbCl3, 99.99wt%), thioacetamide (TAA, 99wt%), and ethylene glycol (EG, 99wt%) were bought from Sigma-Aldrich, USA and applied directly without any treatment. The carbon nanotubes (CNTs, 95wt%) were purchased from Nanostructured & Amorphous Materials, Inc., USA, and ultrasonicated in concentrated H2SO4/HNO3 for 2 h before use. The synthesis of the Sb2S3@CNTs hybrid was accomplished through a novel method at room temperature. First, 45 mg SbCl3 was dissolved in 5 mL EG to generate a colorless solution. Another 5 mL EG containing 25 mg of TAA along with a suspension of 8.5 mg CNTs was then synchronously added into the transparent solution. The solution generated black with some brick-red precipitate after stirring for 30 min. The solution was left for 24 h at room temperature, then centrifuged, followed by washing with anhydrous ethanol several times before vacuum-drying at 70°C for 10 h. The weight of collected Sb2S3@CNTs samples was around 40 mg. CNT-free Sb2S3 was prepared without the addition of dispersion of CNTs (denoted as pure Sb2S3). In the same way, the annealing treatment of the pure Sb2S3 and Sb2S3@CNTs samples was performed at 400°C for 6 h under Ar in a tube furnace (denoted as pure Sb2S3-400°C and Sb2S3@CNTs-400°C, respectively).

      The crystallographic information of the materials was analyzed by powder X-ray diffraction (XRD) on a Bruker D8 Advanced XRD, whose radiation source was Cu Kα. The morphology of the tested samples was observed by scanning electron microscope (SEM, Hitachi SU-70) and transmission electron microscope (TEM, JEOL 2100F). A Horiba Jobin Yvon LabRam Aramis Raman spectrometer equipped with a 532 nm diode-pumped solid-state laser was applied to investigate the chemical bonding properties of the materials. X-ray photoelectron spectroscopy (XPS, Kratos AXIS 165, Mg Kα radiation source) was used to examine the elemental state of Sb2S3. All binding energy values were referenced to the C 1s peak of carbon at 284.6 eV. Before photographing, the electrodes after cycling were rinsed with dimethyl carbonate (DMC).

      The electrodes, which were made of active material, conductive agent, and binder, were fabricated through the slurry coating. In this experiment, the active substances were the samples (commercial Sb2S3, pure Sb2S3, and Sb2S3@CNTs hybrid), the conductive agent was carbon black, and the binder was sodium alginate. The three were mixed at the weight ratio of 7:2:1. The resulting slurry was then pasted on the copper foil and desiccated overnight at 100°C under vacuum. Coin cells were fabricated in an argon-filled glove box. For LIBs, the counter electrode was lithium foil, and the electrolyte was 1.0 M LiPF6 solution, in which the solvent was fluoroethylene carbonate (FEC) and DMC at a volume ratio of 1:1. For SIBs, Na metal acted as the counter electrode, and the electrolyte was 1.0 M NaClO4 solution, whose solvent was the same as that of the lithium-ion battery electrolyte. The working and counter electrodes were separated by a Celgard 3501 separator. The electrochemical Na-storage and Li-storage properties were evaluated using an Arbin battery testing system (BT2000, Arbin Instruments, USA) in voltage windows of 0.01–2.0 V vs. Na/Na+ and 0.01–2.7 V vs. Li/ Li+ under various current densities. All of the current densities used in this study were calculated based on the active Sb2S3 material, and the loading mass of the electrode was around 1.5 mg·cm−2. The cyclic voltammetry (CV) measurements for analyzing electrochemical reactions of the electrode materials were recorded utilizing a Gamry Potentiostats (Gamry, USA) under a scan rate of 0.1 mV·s−1. Moreover, the electrochemical impedance spectroscopy (EIS) tests were performed on the same equipment within a frequency range of 0.01 Hz–1000 kHz.

      The morphologies of the obtained samples were observed by SEM and TEM. Morphologies for the as-synthesized samples of pure Sb2S3 and Sb2S3@CNTs are observed in Fig. 1. The as-synthesized sample showed a typical spherical shape, in which the Sb2S3 spheres were tightly attached with the CNT backbone (Figs. 1(c) and 1(d)). The size of the Sb2S3@CNTs hybrid ranged from 100 to 600 nm, clustered at approximately 200 nm. In contrast, the CNT-free Sb2S3 sample was composed of large, agglomerated particles between 300 and 800 nm in size, whose morphology was irregular (Figs. 1(a) and 1(b)). This suggests that the growth of Sb2S3 spheres is affected by the addition of CNTs, which not only modify the growth but also prevent particle agglomeration. The TEM and its high-resolution images confirmed that the Sb2S3 spheres were firmly anchored on the lateral wall of the carbon nanotubes (Figs. 1(e) and 1(f)). The absence of the lattice stripe of Sb2S3 was due to the amorphous feature of the as-synthesized product.

      Fig. 1.  SEM images of (a, b) pure Sb2S3 and (c, d) Sb2S3@CNTs hybrid; (e) TEM and (f) high-resolution TEM images of Sb2S3@CNTs hybrid. Inset in (f) shows CNTs are coupled to Sb2S3 spheres.

      Fig. 2(a) shows that the as-synthesized pure Sb2S3 and Sb2S3@CNTs hybrid were in an amorphous state and crystallized well after heat treatment annealing at 400°C in argon. The diffraction in XRD patterns of pure Sb2S3-400°C and Sb2S3@CNTs-400°C presented in Fig. 2(a) matched well with the orthorhombic Sb2S3 phase (PDF#06-0474) with lattice constants of a = 1.1226 nm, b = 1.1307 nm, and c = 0.3835 nm, confirming the formation of Sb2S3 in the experiment, agreeing well with the previous literature [12,2531]. The structure of the Sb2S3@CNTs was further validated by Raman spectroscopy (Fig. 2(b)). As indicated in Fig. 2(b), the Sb2S3@CNTs hybrid revealed a broad band of amorphous materials near 285 cm−1 [3233]. The additional peaks at 1351 and 1587 cm−1 were ascribed to CNTs. In order to study the chemical bonding state of the prepared Sb2S3@CNTs, XPS measurements were carried out. Fig. 2(c) reveals the two evident peaks at 538.7 and 529.4 eV that are designated as Sb 3d3/2 and Sb 3d5/2, respectively, confirming the presence of Sb (III). The S 2p spectrum of Sb2S3@CNTs in Fig. 2(d) could be divided into two peaks at 162.7 eV for S 2p1/2 and 161.5 eV for S 2p3/2, which is a favorable proof of the formation of sulfide [26,31,33]. According to XPS analyses, the successful synthesis of Sb2S3@CNTs is also proved.

      Fig. 2.  (a) XRD patterns and (b) Raman spectrum of pure Sb2S3, Sb2S3@CNTs, pure Sb2S3-400°C, and Sb2S3@CNTs-400°C; XPS spectrum of (c) Sb 3d and (d) S 2p for Sb2S3@CNTs hybrid.

      The Na-storage properties of the Sb2S3@CNTs hybrid as explored by CV and galvanostatic tests are presented in Fig. 3. Fig. 3(a) exhibits the CV curves of Sb2S3@CNTs at a scan rate of 0.1 mV·s−1 over the potential window of 0.01–2.0 V vs. Na+/Na. The initial CV profile of the Sb2S3@CNTs hybrid is significantly different from the succeeding CV profiles, which indicates that the activation process occurred during the first discharge [20]. It can be seen that there are three cathode peaks in the first negative scan, located at 1.03, 0.70, and 0.27 V. The peak at 1.03 V is formed by the combination of electrolyte decomposition, solid electrolyte interphase (SEI) formation, and the layered structure of Sb2S3 as inserted by Na+ [31,34]. The peaks at 0.7 and 0.27 V are assigned to the reductive transformation (Eq. (1)) and the alloying reaction (Eq. (2)), respectively [20,31]. In the subsequent cathode process, the intercalation reaction peaks and reductive transformation peaks appear at 1.4 and 0.92 V, respectively; the peaks at 0.62 and 0.27 V are both attributed to the alloying reaction [21,31]. All the anodic peaks are well-overlapped. In the anodic scan, the peaks centered at 0.79 and 1.3 V correspond to the dealloying reactions of Na3Sb forming Sb and desodiation reactions of Sb and Na2S to Sb2S3, respectively [28]. At approximately 1.85 V, the broad peaks can be observed, which are associated with the release of Na+ from the layer structure [31]. The above results suggest that the Sb2S3@CNTs hybrid has good reversibility.

      Fig. 3.  Electrochemical properties of the samples for SIBs within a potential range of 0.01–2.0 V vs. Na+/Na: (a) CV curves of Sb2S3@CNTs hybrid anode at a scan rate of 0.1 mV·s−1; (b) galvanostatic charge–discharge profiles of Sb2S3@CNTs hybrid anode during three initial cycles at a current density of 50 mA·g−1; (c) cycling stabilities for the Sb2S3@CNTs, pure Sb2S3, and commercial Sb2S3 at 50 mA·g−1; (d) rate performance for the Sb2S3@CNTs and pure Sb2S3.

      Fig. 3(b) exhibits galvanostatic charge–discharge curves of Sb2S3@CNTs anode in the range of 0.01–2.0 V vs. Na+/Na under 50 mA·g−1. The first discharge curve shows three voltage platforms, corresponding to the insertion and SEI formation, conversion (Eq. (1)), and alloying processes (Eq. (2)) in order, which is in good agreement with the CV findings. The first charge and discharge capacities of Sb2S3@CNTs for SIBs are 821 mAh·g−1 and 1236 mAh·g−1, respectively, exhibiting a coulombic efficiency (CE) of 66.4%. In contrast, the first charge/discharge capacities of commercial Sb2S3 and pure Sb2S3 are 379/941 mAh·g−1 and 524/1057 mAh·g−1, respectively, corresponding to lower coulombic efficiencies of 40.3% and 59.6%, respectively. The loss of the first-lap capacity is primarily owing to SEI generation and electrolyte decomposition [35]. The first discharge capacities of Sb2S3@CNTs and pure Sb2S3 are higher than the theoretical capacity of Sb2S3 (946 mAh·g−1), because the first discharge capacities measured by the samples include not only the discharge capacity generated by electrode reaction but also the discharge capacity generated by SEI formation and electrolyte decomposition [36]. With the generation and stability of SEI film, the decomposition of electrolyte is inhibited, so the coulombic efficiencies of the second and third cycle, which are more than 98%, are much higher than that of the first cycle (66.4%).

      The cyclic stability of Sb2S3@CNTs hybrid materials at 50 mA·g−1 is shown in Fig. 3(c), and the commercial Sb2S3 and pure Sb2S3 are also investigated for comparison. Among them, the hybrid electrode displays the best ability regarding capacity and stability. It can be seen that the hybrid material displays a stable cycling performance after the first cycle, though the second discharge capacity decays to 814 mAh·g−1. A reversible discharge capacity of the Sb2S3@CNTs electrode reaches 732 mAh·g−1 at a current density of 50 mA·g−1 after 110 cycles with high CE close to 100%. This is higher than the 493 and 170 mAh·g−1 specific capacities for commercial Sb2S3 and pure Sb2S3, respectively, reflecting its stable cycling capability. On the other hand, the cyclic performance of the annealed Sb2S3 (pure Sb2S3-400°C) electrode was tested under the same conditions and compared with the unannealed pure Sb2S3. As shown in Fig. 4, the unannealed amorphous Sb2S3 (pure Sb2S3) showed better cyclic performance than the annealed Sb2S3 (pure Sb2S3-400°C), which was attributed to its amorphous structure, having very high durability against the sodiation/desodiation-caused volume change [30]. The retention capacity of the Sb2S3@CNTs hybrid is also much higher than that of the Sb2S3 nanostructures reported in other articles [12,2830], suggesting that it is a highly effective strategy for designing a hybrid material (Sb2S3@CNTs) with an amorphous structure and strong coupling.

      Fig. 4.  Cycling stability for pure Sb2S3 and pure Sb2S3-400°C within a potential range of 0.01–2.0 V vs. Na+/Na.

      In addition to the prominent cycling stability, the Sb2S3@CNTs anode also offers significant rate capabilities, providing greater potential for their application in SIBs. As can be seen from Fig. 3(d), the Sb2S3@CNTs electrode delivers a capacity of around 810 mAh·g−1 at a current density of 50 mA·g−1, and 786 mAh·g−1 at 100 mA·g−1. They still exhibit capacities of 764 and 729 mAh·g−1, respectively, at higher rates of 300 and 500 mA·g−1, or 668 and 584 mAh·g−1, respectively, at 1000 and 2000 mA·g−1. When the current rate is turned back to 50 mA·g−1, a reversible capacity of 795 mAh·g−1 is restored accordingly. The result shows that there is little effect on the Na-storage activity of Sb2S3@CNTs even at high current charge and discharge. Notably, the rate capability of the Sb2S3@CNTs is also much better than the pure Sb2S3 and other reported electrodes for sodium storage through the conversion reaction [12,2830]. The remarkable rate capability can be imputed to the unique morphology of amorphous structure and strong coupling strategy, which enables faster charge transfer and reduces electrode reaction resistance during charging and discharging.

      The Li-ion storage capacity of the Sb2S3@CNTs hybrid electrode was also probed with coin-type half cells. Fig. 5(a) reveals the room temperature CV curve of the Sb2S3@CNTs anode measured at a scan rate of 0.1 mV·s−1 in the potential range of 0.01–2.7 V vs. Li+/Li. During the first cathodic and anodic scan, the broad cathodic peak from 1.0 to 2.0 V can be imputed to the SEI film generation and the reductive transformation between Li+ and sulfur atoms (Eq. (1)) [25]. At approximately 0.55 V, a reduction peak is assigned to the reaction of Li + with Sb to form a Li3Sb alloy (Eq. (2)) [37]. The anodic peaks at 1.95 and 1.0 V correspond to the inverse reactions of Eqs. (1) and (2), respectively [30]. In the subsequent scanning process, the cathode peaks are slightly shifted to higher potentials and the region between the CV profiles gets smaller, which is consistent with the formation of SEI during the first discharge [3738].

      Fig. 5.  Electrochemical properties of the samples for LIBs within a potential range of 0.01–2.7 V vs. Li+/Li: (a) CV curves of Sb2S3@CNTs hybrid anode at a scan rate of 0.1 mV·s−1; (b) galvanostatic charge–discharge profiles of Sb2S3@CNTs hybrid anode during three initial cycles at a current density of 100 mA·g−1; (c) cycling stabilities for the Sb2S3@CNTs, pure Sb2S3, and commercial Sb2S3 at 100 mA·g−1; (d) rate performance for the Sb2S3@CNTs and pure Sb2S3.

      The galvanostatic charge–discharge profiles of LIBs with Sb2S3@CNTs electrodes between 0.01 and 2.7 V at 100 mA·g−1 are indicated in Fig. 5(b). The capacities of initial charging and discharging for Sb2S3@CNTs are 1136 mAh·g−1 and 1659 mAh·g−1, corresponding to an initial coulombic efficiency (ICE) of 68.5%. In subsequent cycles, the curves of the charging and discharging are very similar, indicating that the electrochemical performance gradually stabilizes. In contrast, the commercial Sb2S3 and pure Sb2S3 have lower ICEs of about 51.8% and 60.5%, respectively. Fig. 5(c) shows the cycling performances of Sb2S3@CNTs, pure Sb2S3, and commercial Sb2S3 in LIBs at a low current density (100 mAh·g−1). It can be seen that, except for the first two cycles, the capacity of Sb2S3@CNTs is kept at about 1104 mAh·g−1 after 110 cycles, with a CE close to 100%. Instead, the pure Sb2S3 and commercial Sb2S3 maintained capacities of 682 and 527 mAh·g−1, respectively, after 110 cycles, and the capacities continued to decline during the cycles. Therefore, the Sb2S3@CNTs exhibited much better performance compared to pure Sb2S3 and commercial Sb2S3 in both cycling stability and capacity. Moreover, the cycling performance, compared with previously reported batteries with Sb2S3 electrodes, has been significantly improved [12,2527].

      The rate capability of the Sb2S3@CNTs for LIBs is revealed in Fig. 5(d). As can be seen, the Sb2S3@CNTs anode exhibited significantly improved rate capability. The discharge capacities of the hybrid anode reached 1136, 1019, 943, 905, 800, and 704 mAh·g−1 at current densities of 100, 200, 300, 500, 1000, and 2000 mA·g−1, respectively. When the current rate was turned back to 100 mA·g−1, a reversible capacity of 1043 mAh·g−1 was restored accordingly. Notably, the rate performance of Sb2S3@CNTs was relatively enhanced compared to pure Sb2S3 and many reported Sb2S3-based electrodes [12,2527].

      To further verify the superiority of the Sb2S3@CNTs hybrid material, EIS measurements of commercial Sb2S3, pure Sb2S3, and Sb2S3@CNTs electrodes for SIBs and LIBs were performed under an open circuit potential from 0.01 Hz to 1000 kHz; results are displayed in Figs. 6(a) and 6(b). The Nyquist plots of electrodes are composed of a semicircle in the high-to-medium-frequency regions related to the interface impedance (Rint), including the SEI film resistance (RSEI) and charge-transfer resistance (Rct), and an inclined line in the low frequency region corresponding to Warburg diffusion resistance (ZW) caused by the diffusion of Li+ or Na+ into the electrode material [28]. Clearly, among these electrodes, the semicircle diameter of the Sb2S3@CNTs electrode is the smallest, suggesting the lowest interface impedance (${R_{{\rm{int}}}} = {R_{{\rm{SEI}}}} + {R_{{\rm{ct}}}}$). For SIBs, the commercial Sb2S3 and pure Sb2S3 electrodes displayed interface impedances (Rint) of 545.3 and 498 Ω, respectively. Instead, the Sb2S3@CNTs revealed only 122.5 Ω, less than the commercial Sb2S3 and pure Sb2S3 electrodes. For LIBs, the Sb2S3@CNTs electrode exhibited an interface impedance (Rint) of 90.26 Ω, far less than the interface resistances of commercial Sb2S3 and pure Sb2S3, which were 623.5 and 523 Ω, respectively. The improvement kinetics of Sb2S3@CNTs is attributed to the introduction of conductive CNTs and the unique design of materials, where amorphous Sb2S3 spheres are strongly coupled with CNTs, thus improving the electrochemical properties of Sb2S3@CNTs.

      Fig. 6.  Nyquist plots of commercial Sb2S3, pure Sb2S3, and Sb2S3@CNTs electrodes for (a) SIBs and (b) LIBs (Zim—Imaginary part of the impedance; Zre—Real part of the impedance).

      The excellent Na and Li-storage abilities may stem from the unique design of the materials, where amorphous Sb2S3 spheres are strongly coupled with CNTs. Through this coupling, which is not physical mixing, the rapid charge transfer between the CNTs network and the insulated Sb2S3 sphere can be realized, and the resistance of the electrode reaction can be reduced [39]. Meanwhile, the CNTs could be used as a flexible backbone to anchor and stabilize Sb2S3 particles, preventing their agglomeration during charging and discharging [39]. Moreover, the unique structure of the amorphous spheres should play a key role in the high capacity, excellent cyclability, and rate property. The amorphous structure provides a higher interfacial area, which allows most of the material particles to react adequately with the ions, facilitating the rapid diffusion of ions, hence offering high capacity and rate capability. It is also conducive to the adjustment of mechanical stress/strain, thereby warranting maximum electrode stability [26,40]. The morphology of the Sb2S3 spheres after 110 cycles for Na and Li storage was investigated with SEM (Fig. 7). The SEM images of the cycled electrodes for both SIBs and LIBs show no obvious morphological changes, demonstrating a robust structure of the amorphous spheres, which can relieve the structural strain availably and fit the dramatic volume variation during reduplicative sodiation/desodiation and lithiation/delithiation. Thus, all of the above factors are conducive to a robust and stable Sb2S3@CNTs product to achieve high-performance Na and Li storage.

      Fig. 7.  SEM images after 110th cycled Sb2S3@CNTs electrodes for (a, b) SIBs and (c, d) LIBs.

      In summary, a strongly coupled Sb2S3@CNTs hybrid was designed and prepared through a facile one-step method at room temperature. The obtained Sb2S3@CNTs hybrid, which is made up of amorphous Sb2S3 spheres tightly anchored with CNTs, can serve as an ideal structure for storing Na and Li. The hybrid electrodes deliver a high sodium storage capacity of 814 mAh·g−1 at 50 mA·g−1, retaining 732 mAh·g−1 during the 110th cycle at 50 mA·g−1. Moreover, the electrode achieves a high lithium storage capacity of 1136 mAh·g−1 at 100 mA·g−1 and retains 1104 mAh·g−1 at the 110th cycle at 100 mA·g−1. Such excellent properties indicate the potential of the strongly coupled Sb2S3@CNTs hybrid, which can be regarded as new alternative electrode material in LIBs and SIBs. The strong-coupling strategy has enormous application potential and is easy to operate, providing feasibility for the design of other high-performance amorphous hybrid electrodes.

      The authors would like to acknowledge the financial support from the Shanghai University of Engineering Science.

    • [1]
      B. Dunn, H. Kamath, and J.M. Tarascon, Electrical energy storage for the grid: A battery of choices, Science, 334(2011), No. 6058, p. 928. doi: 10.1126/science.1212741
      [2]
      J. Liu, J.G. Zhang, Z.G. Yang, J.P. Lemmon, C. Imhoff, G.L. Graff, L.Y. Li, J.Z. Hu, C.M. Wang, J. Xiao, G. Xia, V.V. Viswanathan, S. Baskaran, V. Sprenkle, X.L. Li, Y.Y. Shao, and B. Schwenzer, Materials science and materials chemistry for large scale electrochemical energy storage: From transportation to electrical grid, Adv. Funct. Mater., 23(2013), No. 8, p. 929. doi: 10.1002/adfm.201200690
      [3]
      S.G. He, S.F. Wang, H.D. Chen, X.H. Hou, and Z.P. Shao, A new dual-ion hybrid energy storage system with energy density comparable to that of ternary lithium ion batteries, J. Mater. Chem. A, 8(2020), No. 5, p. 2571. doi: 10.1039/C9TA12660K
      [4]
      H.D. Chen, X.H. Hou, F.M. Chen, S.F. Wang, B. Wu, Q. Ru, H.Q. Qin, and Y.C. Xia, Milled flake graphite/plasma nano-silicon@carbon composite with void sandwich structure for high performance as lithium ion battery anode at high temperature, Carbon, 130(2018), p. 433. doi: 10.1016/j.carbon.2018.01.021
      [5]
      H.L. Pan, Y.S. Hu, and L.Q. Chen, Room-temperature stationary sodium-ion batteries for large-scale electric energy storage, Energy Environ. Sci., 6(2013), No. 8, p. 2338. doi: 10.1039/c3ee40847g
      [6]
      X.Y. Wang, S.F. Wang, K.X. Shen, S.G. He, X.H. Hou, and F.M. Chen, Phosphorus-doped porous hollow carbon nanorods for high-performance sodium-based dual-ion batteries, J. Mater. Chem. A, 8(2020), No. 7, p. 4007. doi: 10.1039/C9TA11246D
      [7]
      S.W. Kim, D.H. Seo, X.H. Ma, G. Ceder, and K. Kang, Electrode materials for rechargeable sodium-ion batteries: Potential alternatives to current lithium-ion batteries, Adv. Energy Mater., 2(2012), No. 7, p. 710. doi: 10.1002/aenm.201200026
      [8]
      J.F. Mao, T.F. Zhou, Y. Zheng, H. Gao, H.K. Liu, and Z.P. Guo, Two-dimensional nanostructures for sodium-ion battery anodes, J. Mater. Chem. A, 6(2018), No. 8, p. 3284. doi: 10.1039/C7TA10500B
      [9]
      P. Ge and M. Fouletier, Electrochemical intercalation of sodium in graphite, Solid State Ionics, 28-30(1988), p. 1172. doi: 10.1016/0167-2738(88)90351-7
      [10]
      Y. Han, S.Y. Liu, L. Cui, L. Xu, J. Xie, X.K. Xia, W.K. Hao, B. Wang, H. Li, and J. Gao, Graphene-immobilized flower-like Ni3S2 nanoflakes as a stable binder-free anode material for sodium-ion batteries, Int. J. Miner. Metall. Mater., 25(2018), No. 1, p. 88. doi: 10.1007/s12613-018-1550-6
      [11]
      D.A. Stevens and J.R. Dahn, High capacity anode materials for rechargeable sodium-ion batteries, J. Electrochem. Soc., 147(2000), No. 4, p. 1271. doi: 10.1149/1.1393348
      [12]
      A.S. Hameed, M.V. Reddy, J.L.T. Chen, B.V.R. Chowdari, and J.J. Vittal, RGO/stibnite nanocomposite as a dual anode for lithium and sodium ion batteries, ACS Sustainable Chem. Eng., 4(2016), No. 5, p. 2479. doi: 10.1021/acssuschemeng.5b01211
      [13]
      J.F. Qian, Y. Chen, L. Wu, Y.L. Cao, X.P. Ai, and H.X. Yang, High capacity Na-storage and superior cyclability of nanocomposite Sb/C anode for Na-ion batteries, Chem. Commun., 48(2012), No. 56, p. 7070. doi: 10.1039/c2cc32730a
      [14]
      A. Darwiche, C. Marino, M.T. Sougrati, B. Fraisse, L. Stievano, and L. Monconduit, Better cycling performances of bulk Sb in Na-ion batteries compared to Li-ion systems: An unexpected electrochemical mechanism, J. Am. Chem. Soc., 134(2012), No. 51, p. 20805. doi: 10.1021/ja310347x
      [15]
      Y.H. Liu, Y.H. Xu, Y.J. Zhu, J.N. Culver, C.A. Lundgren, K. Xu, and C.S. Wang, Tin-coated viral nanoforests as sodium-ion battery anodes, ACS Nano, 7(2013), No. 4, p. 3627. doi: 10.1021/nn400601y
      [16]
      H.L. Zhu, Z. Jia, Y.C. Chen, N. Weadock, J.Y. Wan, O. Vaaland, X.G. Han, T. Li, and L.B. Hu, Tin anode for sodium-ion batteries using natural wood fiber as a mechanical buffer and electrolyte reservoir, Nano Lett., 13(2013), No. 7, p. 3093. doi: 10.1021/nl400998t
      [17]
      B. Luo, T.F. Qiu, D.L. Ye, L.Z. Wang, and L.J. Zhi, Tin nanoparticles encapsulated in graphene backboned carbonaceous foams as high-performance anodes for lithium-ion and sodium-ion storage, Nano Energy, 22(2016), p. 232. doi: 10.1016/j.nanoen.2016.02.024
      [18]
      Z. Hu, L.X. Wang, K. Zhang, J.B. Wang, F.Y. Cheng, Z.L. Tao, and J. Chen, MoS2 nanoflowers with expanded interlayers as high-performance anodes for sodium-ion batteries, Angew. Chem. Int. Edit., 53(2014), No. 47, p. 12794. doi: 10.1002/anie.201407898
      [19]
      J.J. Wang, C. Luo, J.F. Mao, Y.J. Zhu, X.L. Fan, T. Gao, A.C. Mignerey, and C.S. Wang, Solid-state fabrication of SnS2/C nanospheres for high-performance sodium ion battery anode, ACS Appl. Mater. Interfaces, 7(2015), No. 21, p. 11476. doi: 10.1021/acsami.5b02413
      [20]
      X.H. Xiong, G.H. Wang, Y.W. Lin, Y. Wang, X. Ou, F.H. Zheng, C.H. Yang, J.H. Wang, and M.L. Liu, Enhancing sodium ion battery performance by strongly binding nanostructured Sb2S3 on sulfur-doped graphene sheets, ACS Nano, 10(2016), No. 12, p. 10953. doi: 10.1021/acsnano.6b05653
      [21]
      D.Y.W. Yu, P.V. Prikhodchenko, C.W. Mason, S.K. Batabyal, J. Gun, S. Sladkevich, A.G. Medvedev, and O. Lev, High-capacity antimony sulphide nanoparticle-decorated graphene composite as anode for sodium-ion batteries, Nat. Commun., 4(2013), art. No. 2922. doi: 10.1038/ncomms3922
      [22]
      Y.J. Zhu, Y. Wen, X.L. Fan, T. Gao, F.D. Han, C. Luo, S.C. Liou, and C.S. Wang, Red phosphorus–single-walled carbon nanotube composite as a superior anode for sodium ion batteries, ACS Nano, 9(2015), No. 3, p. 3254. doi: 10.1021/acsnano.5b00376
      [23]
      J.F. Mao, X.L. Fan, C. Luo, and C.S. Wang, Building self-healing alloy architecture for stable sodium-ion battery anodes: A case study of tin anode materials, ACS Appl. Mater. Interfaces, 8(2016), No. 11, p. 7147. doi: 10.1021/acsami.6b00641
      [24]
      X.L. Fan, J.F. Mao, Y.J. Zhu, C. Luo, L.M. Suo, T. Gao, F.D. Han, S.C. Liou, and C.S. Wang, Superior stable self-healing SnP3 anode for sodium-ion batteries, Adv. Energy Mater., 5(2015), No. 18, art. No. 1500174. doi: 10.1002/aenm.201500174
      [25]
      P.V. Prikhodchenko, J. Gun, S. Sladkevich, A.A. Mikhaylov, O. Lev, Y.Y. Tay, S.K. Batabyal, and D.Y.W. Yu, Conversion of hydroperoxoantimonate coated graphenes to Sb2S3@Graphene for a superior lithium battery anode, Chem. Mater., 24(2012), No. 24, p. 4750. doi: 10.1021/cm3031818
      [26]
      C.M. Park, Y. Hwa, N.E. Sung, and H.J. Sohn, Stibnite (Sb2S3) and its amorphous composite as dual electrodes for rechargeable lithium batteries, J. Mater. Chem., 20(2010), No. 6, p. 1097. doi: 10.1039/B918220A
      [27]
      D.Y.W. Yu, H.E. Hoster, and S.K. Batabyal, Bulk antimony sulfide with excellent cycle stability as next-generation anode for lithium-ion batteries, Sci. Rep., 4(2014), art. No. 4562. doi: 10.1038/srep04562
      [28]
      H.S. Hou, M.J. Jing, Z.D. Huang, Y.C. Yang, Y. Zhang, J. Chen, Z.B. Wu, and X.B. Ji, One-dimensional rod-like Sb2S3-based anode for high-performance sodium-ion batteries, ACS Appl. Mater. Interfaces, 7(2015), No. 34, p. 19362. doi: 10.1021/acsami.5b05509
      [29]
      Y.Y. Zhu, P. Nie, L.F. Shen, S.Y. Dong, Q. Sheng, H.S. Li, H.F. Luo, and X.G. Zhang, High rate capability and superior cycle stability of a flower-like Sb2S3 anode for high-capacity sodium ion batteries, Nanoscale, 7(2015), No. 7, p. 3309. doi: 10.1039/C4NR05242K
      [30]
      Z. Yi, Q.G. Han, Y. Cheng, Y.M. Wu, and L.M. Wang, Facile synthesis of symmetric bundle-like Sb2S3 micron-structures and their application in lithium-ion battery anodes, Chem. Commun., 52(2016), No. 49, p. 7691. doi: 10.1039/C6CC03176E
      [31]
      Y.B. Zhao and A. Manthiram, Amorphous Sb2S3 embedded in graphite: A high-rate, long-life anode material for sodium-ion batteries, Chem. Commun., 51(2015), No. 67, p. 13205. doi: 10.1039/C5CC03825A
      [32]
      B.H. Juárez, S. Rubio, J. Sánchez-Dehesa, and C. López, Antimony trisulfide inverted opals: Growth, characterization, and photonic properties, Adv. Mater., 14(2002), No. 20, p. 1486. doi: 10.1002/1521-4095(20021016)14:20<1486::AID-ADMA1486>3.0.CO;2-P
      [33]
      J.F. Qian, X.Y. Wu, Y.L. Cao, X.P. Ai, and H.X. Yang, High capacity and rate capability of amorphous phosphorus for sodium ion batteries, Angew. Chem. Int. Edit., 52(2013), No. 17, p. 4633. doi: 10.1002/anie.201209689
      [34]
      S.H. Dong, C.X. Li, X.L. Ge, Z.Q. Li, X.G. Miao, and L.W. Yin, ZnS–Sb2S3@C core-double shell polyhedron structure derived from metal-organic framework as anodes for high performance sodium ion batteries, ACS Nano, 11(2017), No. 6, p. 6474. doi: 10.1021/acsnano.7b03321
      [35]
      F.X. Xie, L. Zhang, Q.F. Gu, D.L. Chao, M. Jaroniec, and S.Z. Qiao, Multi-shell hollow structured Sb2S3 for sodium-ion batteries with enhanced energy density, Nano Energy, 60(2019), p. 591. doi: 10.1016/j.nanoen.2019.04.008
      [36]
      W.P. Sun, X.H. Rui, J.X. Zhu, L.H. Yu, Y. Zhang, Z.C. Xu, S. Madhavi, and Q.Y. Yan, Ultrathin nickel oxide nanosheets for enhanced sodium and lithium storage, J. Power Sources, 274(2015), p. 755. doi: 10.1016/j.jpowsour.2014.10.105
      [37]
      W.H. Yin, W.W. Chai, K. Wang, W.K. Ye, Y.C. Rui, and B.H.J. Tang, A highly Meso@Microporous carbon-supported Antimony sulfide nanoparticles coated by conductive polymer for high-performance lithium and sodium ion batteries, Electrochim. Acta, 321(2019), art. No. 134699. doi: 10.1016/j.electacta.2019.134699
      [38]
      S. Wang, S. Yuan, Y.B. Yin, Y.H. Zhu, X.B. Zhang, and J.M. Yan, Green and facile fabrication of MWNTs@Sb2S3@PPy coaxial nanocables for high-performance Na-ion batteries, Part. Part. Syst. Charact., 33(2016), No. 8, p. 493. doi: 10.1002/ppsc.201500227
      [39]
      J.F. Ni, Y. Zhao, T.T. Liu, H.H. Zheng, L.J. Gao, C.L. Yan, and L. Li, Strongly coupled Bi2S3@CNT hybrids for robust lithium storage, Adv. Energy Mater., 4(2014), No. 16, art. No. 1400798. doi: 10.1002/aenm.201400798
      [40]
      S.M. Hwang, J. Kim, Y. Kim, and Y. Kim, Na-ion storage performance of amorphous Sb2S3 nanoparticles: Anode for Na-ion batteries and seawater flow batteries, J. Mater. Chem. A, 4(2016), No. 46, p. 17946. doi: 10.1039/C6TA07838A

    Catalog



    • Hua Liao, Huizhe Liu, Qingyi Gou, et al. Advancement of Sb–based anodes for rechargeable lithium–ion and sodium–ion batteries. Journal of Power Sources, 2025, 625: 235739.doi: 10.1016/j.jpowsour.2024.235739
      Yathavan Subramanian, Anitha Dhanasekaran, Lukman Ahmed Omeiza, et al. A review on applications of carbon nanotubes-based metal-sulfide composite anode materials (CNTs/MS) for sodium (Na)-ion batteries. Emergent Materials, 2024, 7(2): 357.doi: 10.1007/s42247-023-00501-3
      Jie Lu, Huiling Du, Yupu Shi, et al. In-situ growth of Mn2V2O7-NaV8O20 heterogeneous nanoribbons for superior Zn-ion storage. Journal of Energy Storage, 2024, 101: 113778.doi: 10.1016/j.est.2024.113778
    • /

      返回文章
      返回