Next Article in Journal
Concrete Made with Iron Ore Tailings as a Fine Aggregate: A Step towards Sustainable Concrete
Next Article in Special Issue
Environmental Impacts of Photovoltaic Energy Storage in a Nearly Zero Energy Building Life Cycle
Previous Article in Journal
Effect of Temperature on the Physical Salt Attack of Cement Mortars under Repeated Partial Immersion in Sodium Sulfate Solution
Previous Article in Special Issue
Analysing the Implications of Charging on Nanostructured Li2MnO3 Cathode Materials for Lithium-Ion Battery Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ga2Te3-Based Composite Anodes for High-Performance Sodium-Ion Batteries

Department of Chemical and Biological Engineering, Gachon University, Seongnam 13120, Gyeonggi, Korea
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(18), 6231; https://doi.org/10.3390/ma15186231
Submission received: 13 August 2022 / Revised: 2 September 2022 / Accepted: 6 September 2022 / Published: 8 September 2022

Abstract

:
Recently, metal chalcogenides have received considerable attention as prospective anode materials for sodium-ion batteries (SIBs) because of their high theoretical capacities based on their alloying or conversion reactions. Herein, we demonstrate a gallium(III) telluride (Ga2Te3)-based ternary composite (Ga2Te3–TiO2–C) synthesized via a simple high-energy ball mill as a great candidate SIB anode material for the first time. The electrochemical performance, as well as the phase transition mechanism of Ga2Te3 during sodiation/desodiation, is investigated. Furthermore, the effect of C content on the performance of Ga2Te3–TiO2–C is studied using various electrochemical analyses. As a result, Ga2Te3–TiO2–C with an optimum carbon content of 10% (Ga2Te3–TiO2–C(10%)) exhibited a specific capacity of 437 mAh·g−1 after 300 cycles at 100 mA·g−1 and a high-rate capability (capacity retention of 96% at 10 A·g−1 relative to 0.1 A·g−1). The good electrochemical properties of Ga2Te3–TiO2–C(10%) benefited from the presence of the TiO2–C hybrid buffering matrix, which improved the mechanical integrity and electrical conductivity of the electrode. This research opens a new direction for the improvement of high-performance advanced SIB anodes with a simple synthesis process.

Graphical Abstract

1. Introduction

In the last few decades, lithium-ion batteries (LIBs) have been utilized as an effective alternative to unsustainable fossil fuels in energy storage systems such as portable electronic devices and electric vehicles [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. Nevertheless, by virtue of the limited reserves and the high cost of Li, much attention has been drawn to developing alternative secondary batteries to overcome these issues [15]. Sodium-ion batteries (SIBs) are reputed as one of the most viable secondary batteries among many next-generation batteries because of their many similarities to LIBs and the abundance of Na on earth [16,17,18,19]. Because of the chemical similarity between Na and Li, the Na storage mechanism of anode materials is similar to that of LIB systems, including intercalation/deintercalation, conversion, and alloying/dealloying reactions. Therefore, considerable attempts have been dedicated to finding suitable anode materials for SIBs. However, the slow reaction kinetics owing to the large ionic radius (1.03 nm for Na+ relative to 0.75 nm for Li+) causes low cycling stability and rate capability or even complete electrochemical inactivity. Thus, the improvement of desirable anode materials for high-performance SIBs is urgently required for expanding their practical application scope. Recently, alloy-based materials have received considerable attention as SIB anode materials owing to their high theoretical capacities (Na–Si: 955 mAh·g−1, Na–Ge: 368 mAh·g−1, Na–Sn: 845 mAh·g−1, Na–Pb: 486 mAh·g−1) [20]. However, similar to LIBs, the large volume change of these materials significantly restricts the long-term cycling of SIBs.
At present, there are numerous potential anodes studied for SIBs, which can be classified in accordance with the reaction mechanism: intercalation (carbon-based materials) [21,22,23], conversion (sulfides and oxides) [24,25,26], and alloying reaction (Sb, Sn, and P) [27,28,29]. Among these, chalcogenide materials (particularly S- and Se-based materials) have received significant attention as they undergo sequential alloying and conversion reactions, resulting in high capacities [30,31,32]. Te, another chalcogen group element, has recently been recognized as an effective substitute for S and Se by virtue of the high theoretical Na storage capacity (420 mAh·g−1) based on alloying with two Na ions (Na2Te) [33]. Furthermore, the high density of Te (6.23 g·cm–3) results in a high theoretical volumetric capacity of 2621 mAh·cm–3, which is comparable to that of S (3468 mAh·cm–3) and Se (3252 mAh·cm–3). Furthermore, Te has superior electrical conductivity (ca. 1.9 × 10−2 S·cm– 1) when compared to other nonmetallic chalcogenides (S (ca. 5.1 × 10–16 S·cm–1) and Se (ca. 1.1 × 10–4 S·cm–1) [34]. In line with this, although a Ga2S3-based composite anode was previously used as the anode for SIB, the rate performance was not satisfactory due to the low electronic conductivity of S [35]. In contrast, Te-based composite electrodes have achieved good rate performance and high capacity retention thanks to the high electronic conductivity of Te, as shown in Table S1 [36,37,38,39,40,41,42,43,44]. Despite these great characteristics, the study of the Te-based SIB electrode mechanism has so far been rare [39,45,46,47,48].
Ga-based materials, such as Ga oxide/sulfide anodes, have a large theoretical capacity (682–1591 mAh·g−1), innate self-healing capability, and a high tolerance against volume change [49,50,51,52]. Ga-based materials have recently emerged as potential electrode materials because of their unique self-healing properties based on the low melting temperature of Ga (29.9 °C). The intermediate liquid Ga formed during sodiation increases the tolerance of volume expansion of active materials, which significantly contributes to the cycling stability [53]. For instance, a composite including reduced graphene oxide and gallium oxide nanosheets (Ga2O3 NSs/rGO) by Yang et al. provided a steady capacity of 555 mAh·g−1 at 0.1 A·g−1 [54], whereas a template-derived Ga2S3 nanorod anode could obtain a discharge capacity of 476 mAh·g−1 at 0.4 A·g−1 [55]. Using in situ microscopy, Wu et al. investigated the self-healing properties of a liquid metal Ga-alloy during the charge/discharge process [56]. Considering the aforementioned advantages of Te and Ga, gallium telluride alloys (i.e., GaxTey) are expected to be great candidate anode materials for SIBs.
In this work, a Ga2Te3-based composite electrode (Ga2Te3–TiO2–C) was successfully synthesized utilizing a simple solid-state high-energy ball milling (HEBM) method and studied as a potential SIB anode material. The feasibility of the Ga2Te3–TiO2–C anode for SIBs was investigated through galvanostatic measurements, differential capacity analysis, and electrochemical impedance spectroscopy (EIS). Furthermore, the reaction mechanism of Ga2Te3–TiO2–C anode during sodiation/desodiation was first investigated via ex situ X-ray diffraction (XRD) analysis. In addition, the best C content (10 wt.%) in the Ga2Te3–TiO2–C composite was derived through various electrochemical tests. The high cycling and rate performances of Ga2Te3–TiO2–C(10%) obtained in this study are superior or equivalent to those of the most recent chalcogenide-based electrodes in SIBs.

2. Experiment

2.1. Material Synthesis

The crystalline Ga2Te3 alloy was synthesized using a simple solid-state HEBM. First, commercial powder of Ti (325 mesh, 99.99%, Alfa Aesar, Haverhill, MA, USA), Ga2O3 (99.99%, Sigma Aldrich, St. Louis, MO, USA), Te (99.8%, Alfa Aesar) (where the ratio is 6:2:3), and ZrO2 balls were placed in the milling bowl (powder mixture: ball = 1:20, w/w). Under an Ar atmosphere, the powder compound was milled for 10 h at 300 rpm. Subsequently, the obtained powder (Ga2Te3–TiO2) was manually mixed with acetylene carbon black powder (C) (99.9+%, bulk density: 170–230 g·L−1, S.A.: 75 m2·g−1, Alfa Aesar) in mass ratios of 9:1 (Ga2Te3–TiO2–C(10%)), 8:2 (Ga2Te3–TiO2–C(20%)), and 7:3 (Ga2Te3–TiO2–C(30%)). The mixtures were ball-milled under the same condition as the first milling. The mechanochemical reaction route for preparing the Ga2Te3–TiO2–C composite is outlined below.
First step: 2Ga2O3 + 6Te + 3Ti → 2Ga2Te3 + 3TiO2 (Ga2Te3–TiO2),
Second step: Ga2Te3–TiO2 + C → Ga2Te3–TiO2–C.

2.2. Material Characterization

The morphology and crystallinity of Ga2Te3–TiO2 and Ga2Te3–TiO2–C were characterized by employing XRD (D/MAX–2200 Rigaku, Tokyo, Japan) with Cu Kα (λ = 1.54 A) radiation at a scan rate of 2°·min−1, as well as energy-dispersive X-ray spectroscopy (EDXS), scanning electron microscopy (SEM, Hitachi S4700, Tokyo, Japan), and high-resolution transmission electron microscopy (HRTEM, JEOL JEM-2100F, Tokyo, Japan). The mechanism of phase change state in the electrode during Na ion reaction was performed through ex situ XRD.

2.3. Electrochemical Measurements

To assess the electrochemical performance of the electrode, the coin-typed cells were assembled in an Ar-filled glovebox with Na foil acting as the counter electrode and polyethylene as the separation membrane. The SIB electrolyte was 1 M NaClO4 in propylene carbonate/ethylene carbonate (1:1 by v/v) with 5% fluoroethylene carbonate. An electrode was prepared using a 7.0:1.5:1.5 (w/w) combination of the active material, conductive carbon (Super-P, 99.9%, Alfa Aesar), and poly(acrylic acid) (PAA, Mw 450,000, Sigma Aldrich) binder. Then, electrodes were coated on Cu foil using a doctor blade and dried overnight in a vacuum oven at 70 °C (a typical mass loading of 1.0−1.5 mg on a Cu foil diameter of 12.5 mm). Using a battery-testing device (WBCS3000, WonATech, Seoul, Korea), the electrochemical performance of Ga2Te3–TiO2–C was assessed. For a voltage range of 0.01–2.5 V (vs. Na/Na+), the galvanostatic charge–discharge (GCD) profile was investigated. Using cyclic voltammetry (CV) measurements with a scanning rate of 0.1 mV·s−1, the electrochemical reactions of the electrodes with Na+ were evaluated. A battery cycler (WBCS3000, WonATech) was used to measure the rate capability at current densities of 0.1, 0.5, 1, 3, 5, and 10 A·g−1. The EIS was measured using a ZIVE MP1 (WonaTech) analyzer in the frequency range of 100 kHz–100 mHz at an AC amplitude of 10 mV.

3. Results and Discussion

The XRD results of the as-prepared Ga2Te3–TiO2 following the HEBM process are shown in Figure 1a. The peaks at 26.2°, 30.3°, 43.4°, 51.4°, 53.8°, 63.0°, 69.4°, and 79.5° corresponded to the (111), (200), (220), (311), (222), (400), (331), and (422) planes of Ga2Te3, respectively, indicating the formation of monoclinic Ga2Te3. The relatively small peaks observed at 33.3°, 48.5°, 67.9°, and 79.0° matched the (311), (020), (621), and (424) planes of TiO2, respectively. The presence of amorphous C decreased the crystallinity of Ga2Te3 and TiO2 in Ga2Te3–TiO2–C (Figure S1) [57]. The absence of impurity peaks for the precursor components (Ga2O3, Ti, and Te) in Ga2Te3–TiO2 indicated the complete conversion of the raw materials to the target product via a solid-state reaction. Nevertheless, the existence of small diffraction peaks near 2θ = 22.9°, 35.1°, and 37.5° were associated with the formation of GaTe (PDF#33-0571) and Ga2Te5 (PDF#45-0954), which are different forms of gallium telluride alloy (as shown in Figure S2). Given the advantages of gallium telluride alloy (i.e., GaxTey), they are also expected to be good candidate anode materials for SIBs; thus, the existence of different forms of minor GaxTey (namely, Ga2Te5 and GaTe) did not negatively affect the electrochemical performance of the Ga2Te3–TiO2–C composite (capacity still in the range of 682−1591 mAh·g−1). The EDXS showed that the four component elements of the composite in the electrode and the relative content of elements were reasonable with the stoichiometric ratio of Ga2Te3 and TiO2 (Figure 1b), indicating that the Ga2Te3-based composite was synthesized successfully.
Morphological and structural analyses of Ga2Te3–TiO2–C(10%) were conducted, including SEM, HRTEM, and EDXS, as indicated in Figure 2. According to SEM images (Figure 2a,b), the Ga2Te3–TiO2–C(10%) particle size ranged from sub-micrometers to a few micrometers. The HRTEM images (Figure 2c and Figures S3) revealed crystalline lattice spacings of 0.340, 0.294, and 0.208 nm, which corresponded to the (111), (200), and (220) crystal planes of Ga2Te3, respectively, and 0.311 nm, which corresponded to the (002) plane of TiO2. Additionally, amorphous C, which was anticipated to serve as a buffering network for the active material, formed surrounding Ga2Te3 and TiO2. The scanning transmission electron microscopy image with EDXS mapping analysis (Figure 2d and Figures S4) showed a uniform distribution of each element (Ga, Te, Ti, O, and C) in the Ga2Te3–TiO2–C(10%). Furthermore, the SEM–EDXS analysis results (Figures S5 and S6) of Ga2Te3–TiO2 with different content of C showed C concentrations almost identical to their theoretical values. Additionally, the stoichiometric ratio of the constituent elements was nearly identical to the theoretical values, according to a quantitative analysis of the EDXS result. In the EDS spectrum of G2Te3–TiO2, Te was determined to be 25%. However, in the presence of 10% carbon content, the distribution of Te gradually decreased (specifically to 18%, as shown in EDS spectrum of Ga2Te3–TiO2–C(10%) in Figure S5); thus, the distribution of Te content in the TEM elemental mapping was relatively less.
The Na ion storage characteristics of the Ga2Te3–TiO2–C electrode were studied using a half-cell form with Na metal as the counter electrode (Figure 3). The GCD voltage profiles of Ga2Te3–TiO2–C(10%), Ga2Te3–TiO2–C(20%), and Ga2Te3–TiO2–C(30%) for SIBs are shown in Figure 3a and Figures S7. The first discharge/charge capacities of Ga2Te3–TiO2–C(10%), Ga2Te3–TiO2–C(20%), and Ga2Te3–TiO2–C(30%) were 599/414, 550/357, and 462/279 mAh·g−1, respectively, which corresponded to initial coulombic efficiencies (ICEs) of 69.1%, 64.9%, and 60.4%, respectively. The poor reversibility between the first and second cycles was because of the SEI layer formation in the first cycle. This poor reversion was well documented in previous studies [58,59,60]. The large capacity difference between the first and second cycle indicated the large irreversible capacity contribution from the SEI layer. However, the reversibility of the electrode (Ga2Te3–TiO2–C (10%) was rapidly enhanced after the second cycle, which could be confirmed by the change in coulombic efficiency (Table S6). According to the EDXS results (Figure S5) and computed theoretical capacities of the separate components (Table S2), the capacity contributions of C and TiO2 to Ga2Te3–TiO2–C(10%) were estimated to be 13% and 22%, respectively. Furthermore, the roles of C and active material were examined (as shown in Figure S8). Ga2Te3–TiO2 achieved a high initial capacity (606 mAh·g−1), but its capacity gradually decreased due to the instability of the electrode structure without buffering C. Moreover, the electrode with only a buffering matrix (TiO2–C) showed very low electrochemical efficiency, close to the theoretical capacity (116 mAh·g−1) (Table S2). The low-capacity contribution of the TiO2–C (~35%) indicated its main role as a buffering matrix. Due to interfacial Na ion storage and electrolyte breakdown, the measured capacities of Ga2Te3–TiO2–C(10%) and Ga2Te3–TiO2 in the SIBs were higher than their theoretical capacities (336 and 333 mAh·g−1, respectively, as computed in Table S3). The change in the reversible capacity of Ga2Te3–TiO2–C for the SIBs was studied using the CE (Table S4) and DCP test of the first 300 cycles (Figure S9). The CE of Ga2Te3–TiO2–C(10%) reached ~99.82% after 150 cycles, slightly decreased, and then stabilized at 98.5% after 300 cycles. The DCP analysis revealed that, for 250 cycles, the main oxidation (at ~0.16, ~1.27, and ~1.42 V) and reduction (at ~0.79 and ~1.58V) peaks remained stable before becoming wider and shifting. However, this polarization had an almost negligible effect on sodiation/desodiation. The reversible capacity of Ga2Te3–TiO2–C(10%) was 436.6 mAh·g−1 (capacity retention (CR) of 97.7%) after 300 cycles at 100 mA·g−1, which was greater than those of Ga2Te3–TiO2–C(20%) (323.8 mAh·g−1) and Ga2Te3–TiO2–C(30%) (264.9 mAh·g−1) (Figure 3b). As shown in Figure S10, although some aggregated particles were observed, the Ga2Te3–TiO2–C(10%) electrode morphology was generally well maintained after 300 cycles. This is because of the presence of TiO2–C, which effieicntly stabilized the electrode structure and mitigated the significant volume variation. In addition, in EDS spectra after 300 cycles, the composition of the Ga2Te3 composite electrode was not significantly changed without impurities (Figure S11). This further proved the stability and good retention of the electrode after the electrochemical reaction. At 500 mA·g−1 (Figure 3c), the reversible capacity of Ga2Te3–TiO2–C(10%) slightly increased until 200 cycles, followed by a gradual decrease. The capacity variation depends on the variation of the redox peaks, in which the oxidation and reduction peaks gradually rise with the cycling, leading to a decrease in polarization and an increase in capacity. In contrast, the oxidation and reduction peaks gradually decrease with the increase in cycling, resulting in a reduced capacity due to the increase in polarization [10,61,62]. This trend was also shown in the DCP analysis (Figures S12 and S13) and CE variation (Table S5). The magnitudes of the reduction (at 0.59 and 1.48 V) and oxidation (at 0.16, 1.27, and 1.69 V) peaks gradually raised over 200 cycles, with a reduction in polarization (Figure S12), and then reduced after 200 cycles, with a rise in polarization (Figure S13). At 100 and 500 mA·g−1, the fluctuation of the DCP profile was examined as a function of the cycle number (Figure S14). The DCP curves of the Ga2Te3–TiO2–C(10%) electrode showed that the overall intensity of the redox peaks was generally stable as the cycle number increased to 300 at 100 mA·g−1. At 500 mA·g−1, the overall magnitudes of the redox peaks increased up to 250 cycles and then decreased with an increase in polarization. Despite the decrease in capacity after 250 cycles, the overall capacity of Ga2Te3–TiO2–C(10%) was still the highest over 500 cycles, reaching 204 mAh·g−1 after 500 cycles with a CR of 76.4%. Figure S15 shows a comparison of the CE variations in Ga2Te3–TiO2–C with varying C contents at 100 and 500 mA·g−1. Table S6 (at 100 mA·g−1) and Table S7 (at 500 mA·g−1) provide summaries of the detailed CE values for the electrodes throughout the first 10 cycles. As shown in Table S6, the ICE of the Ga2Te3–TiO2–C(10%) electrode was slightly higher (69.2%) than that of the Ga2Te3–TiO2–C(20%) (ICE = 64.8%) and Ga2Te3–TiO2–C(30%) electrodes (ICE = 60.5%). Then, after 10 cycles, the CE of the Ga2Te3–TiO2–C(10%) electrode marginally increased and reached the highest among the three various electrodes. This tendency was also discovered at 500 mA·g−1 (Table S7). After the first cycle, the high CE of the Ga2Te3–TiO2–C(10%) electrode showed a high degree of sodiation/desodiation reversibility. The CV curves of the Ga2Te3–TiO2–C(10%) electrode for the first five cycles in the voltage range of 0.005–2.5 V vs. Na/Na+ are shown in Figure 3d. A large reduction peak was observed at 1.37 V during the first discharge process, which denoted the intercalation of Na into Ga2Te3 to form Na2Te and Ga. The reaction between Ga and Na to generate NaGa4 was attributed to being responsible for the peak at 0.52 V. Thus, Na2Te and NaGa4 were the final products after the discharge step was complete. In the charge step, two oxidation peaks were noticed at 0.92 and 1.72 V. The first peak was the result of Na being completely excluded, turning NaGa4 into Ga. Then, Ga intruded into Na2Te to form Ga2Te3 when the anode was charged to 1.72 V. The ex situ investigations concern a thorough analysis of this phase transition. After the second cycle, the curves nearly overlapped, indicating the excellent stability and reversibility of Ga2Te3–TiO2–C(10%). The CV curves of Ga2Te3–TiO2–C(20%) and Ga2Te3–TiO2–C(30%) were almost identical to that of Ga2Te3–TiO2–C(10%), with a similar level of cyclic stability after the second cycle (Figure S16). In addition, the electrochemical performance of Ga2Te3–TiO2 was examined (Figure S17). The GCD profiles of Ga2Te3–TiO2 presented initial charge/discharge capacities of 606/426 mAh·g−1, corresponding to an ICE of 70.2%, which is higher than that of the Ga2Te3-based composite with different C contents. Despite this high ICE of Ga2Te3–TiO2, the capacity gradually decreased with the increase in cycle number, and reached 309 mAh·g−1 after 30 cycles, with a capacity retention of 67%. This is much lower than the Ga2Te3–TiO2 electrode with various carbon contents. In addition, the CV curves did not overlap in the first five cycles. Therefore, the presence of C clearly stabilized the electrode structure, leading to the enhanced electrochemical performance. The rate performances (Figure 3e) and normalized capacity retentions (Figure 3f) of the electrodes were determined. At 0.1, 0.5, 1.0, 3.0, 5.0, and 10.0 A·g1, the average specific capacities of Ga2Te3–TiO2–C(10%) were 455, 408, 374, 348, 321, and 318 mAh·g1, respectively (Figure 3e), which were considerably greater than those of Ga2Te3–TiO2–C(20%) and Ga2Te3–TiO2–C(30%). Surprisingly, even at 10 A·g−1, Ga2Te3–TiO2–C(10%) had a CR of up to 96% (Figure 3f). Additionally, Ga2Te3–TiO2–C(10%) demonstrated a high rate performance when the discharge rate was reduced from 10 A·g−1 to 0.1 A·g−1, resulting in high CR (99.3%).
The reaction mechanism during the first sodiation/desodiation process of the Ga2Te3–TiO2–C(10%) electrode was investigated using ex situ XRD (Figure 4a,b). Peaks corresponding to Na2Te and Ga were observed at a discharge voltage of 1.37 V (D: 1.37 V). When the electrode was fully discharged (D: 5 mV), NaGa4 peaks were observed and Na2Te peaks remained. The NaGa4 phase partly disappeared when the electrode was charged to 0.92 V (C: 0.92V). In a charging state of 1.72 V, the Na2Te phase partly disappeared, Ga was observed, and NaGa4 completely disappeared. Only the peaks corresponding to Ga2Te3 were observed again when the electrode was fully charged to 2.5 V (C: 2.5 V). Ga2Te3 undergoes the following structural changes during sodiation/desodiation:
1st discharge
  • Intercalation stage
    Ga2Te3 + xNa+ + xe → NaxGa2Te3 (2.5−1.37 V). (i)
  • Conversion stage
    NaxGa2Te3 + (6) Na+ + (6x)e → 3Na2Te + 2Ga (1.37−0.52 V). (ii)
  • Alloy stage
    4Ga + Na+ + e → NaGa4 (0.52−0.005 V). (iii)
1st charge
  • De-alloy stage
    NaGa4 → 4Ga + Na+ + e- (0.005−0.92 V). (iv)
  • De-conversion stage
    3Na2Te + 2Ga → LixGa2Te3 + (6x) Na+ + (6x)e (0.92−1.72 V). (v)
  • De-intercalation stage
    NaxGa2Te3 → Ga2Te3 + xNa+ +xe (1.72−2.5 V). (vi)
It is noteworthy that, after the first cycle, the Ga2Te3 phase (major peaks at 53.8°, 69.4°, and 71.5°) was completely recovered without any impurity peaks, showing a highly reversible interaction of Ga2Te3 with Na ions. The alloying/dealloying and conversion mechanism of the Ga2Te3 electrode during charge/discharge is shown by the ex situ XRD results, as schematically depicted in Figure 4c.
For the first, fifth, and 20th cycles, the EIS profiles of the Ga2Te3–TiO2–C(10%), Ga2Te3–TiO2–C(20%), and Ga2Te3–TiO2–C(30%) electrodes were obtained (Figure 5). The simplified equivalent circuit shown in Figure 5d includes the electrolyte resistance (Rb), charge-transfer resistance (Rct), SEI layer resistance (RSEI), interfacial double-layer capacitance (Cdl), Warburg impedance (Zw), and constant phase element (CPE). The Rct at the electrode–electrolyte interface is denoted by compressed semicircles in the mid-frequency region of the Nyquist plot. For all the electrodes, Rct gradually decreased as the cycle number increased from 1 to 20. Ga2Te3–TiO2–C(10%) exhibited the lowest value of Rct after 20 cycles (Table S8), demonstrating the most facile Na ion transportation, which led to the highest Na storage performance.
Currently, there are only a few reports on Ga-based or Te-based anodes for SIBs. However, chalcogenide materials (In2S3, Sb2Se3, etc.) have high specific capacities when they undergo sequential conversion and alloying reactions owing to their unique properties. A comparison of the performances of Ga2Te3–TiO2–C(10%) and other chalcogenide materials demonstrated the high potential of the Ga2Te3-based composite electrodes for future applications (Table 1).

4. Conclusions

We demonstrated a Ga2Te3-based composite as a prospective anode material for SIBs. The Ga2Te3–TiO2·C(10%) anode achieved a high reversible capacity of 437 mAh·g−1 after 300 cycles at 0.1 A·g−1, as well as a high rate capability (CR of 96% at 10 A·g−1 relative to 0.1 A·g−1). The nanoconfined Ga2Te3 crystallites embedded within an electrically conductive TiO2–C hybrid matrix effectively accommodated the Ga2Te3 particle volume variation and avoided the agglomeration of Ga during electrochemical reactions. In addition, Na ion diffusion kinetics and mechanical stability were enhanced by this beneficial morphology, thereby achieving high capacity and long-term cycling performance These findings offer a new direction toward the development of high-performance SIBs with long cycle lifetimes and expansion of the Ga- and Te-based materials in other electrochemical energy storage systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15186231/s1, Figure S1: XRD pattern of Ga2Te3–TiO2–C with different concentration of C. Figure S2: The existence of different forms of GaxTey (namely, Ga2Te5 and GaTe). Figure S3: HRTEM image of Ga2Te3–TiO2–C(10%). Figure S4: SEM image of distribution of the elements in Ga2Te3–TiO2–C(10%). Figure S5: EDX spectrum of as-synthesized Ga2Te3–TiO2–C(10%). Figure S6: EDX analysis of (a) Ga2Te3–TiO2–C(20%), and (b) Ga2Te3–TiO2–C(30%). Figure S7: Galvanostatic discharge-charge profiles of (a) Ga2Te3–TiO2–C(20%) and (b) Ga2Te3–TiO2–C(30%) for SIBs. Figure S8: Cyclic performance of various electrodes at 100 mA g−1. Figure S9: DCP profiles of Ga2Te3–TiO2–C(10%) during 300 cycles measured at 100 mA g−1: (a) 1−150 cycles, (b) 150−300 cycle. Enlarged view of (c) reduction and (d) oxidation peaks. Figure S10: SEM image of Ga2Te3–TiO2–C(10%) after 300 cycles. Figure S11: EDX analysis of Ga2Te3–TiO2–C(10%) after 300 cycles. Figure S12: (a) DCP profiles of Ga2Te3–TiO2–C(10%) during initial 200 cycles measured at 500 mA g−1. Enlarged view of (b) oxidation and (c) reduction peaks. Figure S13: (a) DCP of Ga2Te3–TiO2–C(10%) from 300 cycle to 500 cycles measured at 500 mA g−1. Enlarged view of (b) oxidation and (c) reduction peaks. Figure S14: DCP profiles of Ga2Te3–TiO2–C(10%) (a) at 100 mA g−1 during 300 cycles and (b) at 500 mA g−1 during 500 cycles. Figure S15: Coulombic efficieny of Ga2Te3-TiO2 with different C content at current densities of (a) 100 and (b) 500 mA g−1. Figure S16: (a) CV curves of Ga2Te3–TiO2–C (20%) and (b) Ga2Te3–TiO2–C (30%) for SIBs. Figure S17: (a) Galvanostatic discharge-charge profiles of Ga2Te3-TiO2 at a current density of 100 mA g−1, (b) CV curves of Ga2Te3-TiO2. Table S1: Rate performance of Te-based composite anode for SIB. Table S2: Calculation of capacity contribution of Ga2Te3, TiO2 and C in the Ga2Te3–TiO2–C composite for SIB. Table S3: Calculation of theoretical capacity of Ga2Te3–TiO2–C(10%) and Ga2Te3-TiO2 for SIB. Table S4: Coulombic efficiency variation of Ga2Te3–TiO2–C (10%) at various cycle numbers measured at 100 mA g−1 for SIB. Table S5: Coulombic efficiency variation of Ga2Te3–TiO2–C (10%) at various cycle numbers measured at 500 mA g−1 for SIB. Table S6: Coulombic efficiency of Ga2Te3–TiO2–C at current density of 100 mA g−1 during initial 10 cycles for SIB. Table S7: Coulombic efficiency of Ga2Te3–TiO2–C at current density of 500 mA g−1 during initial 10 cycles for SIB. Table S8: Charge-transfer resistance (Rct) of Ga2Te3–TiO2–C for SIB.

Author Contributions

Data curation, V.P.H.H.; Funding acquisition, J.H.; Investigation, V.P.H.H. and I.T.K.; Supervision, I.T.K. and J.H.; Validation, V.P.H.H. and J.H.; Writing—original draft, V.P.H.H.; Writing—review & editing, I.T.K. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Capacity Enhancement Project through a Korea Basic Science Institute (National Research Facilities and Equipment Center) grant funded by the Ministry of Education (2019R1A6C1010016) and a Korea Institute for Advancement of Technology (KIAT) grant funded by the Korea Government (MOTIE) (No. P0012453, The Competency Development Program for Industry Specialists).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef] [PubMed]
  2. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the development of advanced Li-ion batteries: A review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar] [CrossRef]
  3. Sung, G.-K.; Nam, K.-H.; Choi, J.-H.; Park, C.-M. Germanium telluride: Layered high-performance anode for sodium-ion batteries. Electrochim. Acta 2020, 331, 135393. [Google Scholar] [CrossRef]
  4. Nguyen, T.P.; Kim, I.T. Ag Nanoparticle-Decorated MoS2 Nanosheets for Enhancing Electrochemical Performance in Lithium Storage. Nanomaterials 2021, 11, 626. [Google Scholar] [CrossRef]
  5. Nguyen, T.P.; Kim, I.T. Self-Assembled Few-Layered MoS2 on SnO2 Anode for Enhancing Lithium-Ion Storage. Nanomaterials 2020, 10, 2558. [Google Scholar] [CrossRef]
  6. Preman, A.N.; Lee, H.; Yoo, J.; Kim, I.T.; Saito, T.; Ahn, S.-k. Progress of 3D network binders in silicon anodes for lithium ion batteries. J. Mater. Chem. A 2020, 8, 25548–25570. [Google Scholar] [CrossRef]
  7. Nguyen, T.P.; Kim, I.T. W2C/WS2 Alloy Nanoflowers as Anode Materials for Lithium-Ion Storage. Nanomaterials 2020, 10, 1336. [Google Scholar] [CrossRef]
  8. Kim, W.S.; Vo, T.N.; Kim, I.T. GeTe-TiC-C Composite Anodes for Li-Ion Storage. Materials 2020, 13, 4222. [Google Scholar] [CrossRef]
  9. Vo, T.N.; Kim, D.S.; Mun, Y.S.; Lee, H.J.; Ahn, S.-k.; Kim, I.T. Fast charging sodium-ion batteries based on Te-P-C composites and insights to low-frequency limits of four common equivalent impedance circuits. Chem. Eng. J. 2020, 398, 125703. [Google Scholar] [CrossRef]
  10. Hoang Huy, V.P.; Kim, I.T.; Hur, J. The Effects of the Binder and Buffering Matrix on InSb-Based Anodes for High-Performance Rechargeable Li-Ion Batteries. Nanomaterials 2021, 11, 3420. [Google Scholar] [CrossRef]
  11. Nguyen, Q.; Phung, V.; Kidanu, W.; Ahn, Y.; Nguyen, T.; Kim, I.T. Carbon-free Cu/SbxOy/Sb nanocomposites with yolk-shell and hollow structures as high-performance anodes for lithium-ion storage. J. Alloys Compd. 2021, 878, 160447. [Google Scholar] [CrossRef]
  12. Nguyen, T.P.; Kim, I.T. In Situ Growth of W2C/WS2 with Carbon-Nanotube Networks for Lithium-Ion Storage. Nanomaterials 2022, 12, 1003. [Google Scholar] [CrossRef] [PubMed]
  13. Nguyen, T.P.; Kim, I.T. Boron Oxide Enhancing Stability of MoS2 Anode Materials for Lithium-Ion Batteries. Materials 2022, 15, 2034. [Google Scholar] [CrossRef] [PubMed]
  14. Phan Nguyen, T.; Thi Giang, T.; Tae Kim, I. Restructuring NiO to LiNiO2: Ultrastable and reversible anodes for lithium-ion batteries. Chem. Eng. J. 2022, 437, 135292. [Google Scholar] [CrossRef]
  15. Tarascon, J.-M. Is lithium the new gold? Nat. Chem. 2010, 2, 510. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.; He, D.; Han, R.; Wei, G.; Qiao, Y. Nanostructured potassium and sodium ion incorporated Prussian blue frameworks as cathode materials for sodium-ion batteries. Chem. Commun. 2017, 53, 5569–5572. [Google Scholar] [CrossRef]
  17. Liu, Y.; Qiao, Y.; Wei, G.; Li, S.; Lu, Z.; Wang, X.; Lou, X. Sodium storage mechanism of N, S co-doped nanoporous carbon: Experimental design and theoretical evaluation. Energy Storage Mater. 2018, 11, 274–281. [Google Scholar] [CrossRef]
  18. Qiao, Y.; Ma, M.; Liu, Y.; Li, S.; Lu, Z.; Yue, H.; Dong, H.; Cao, Z.; Yin, Y.; Yang, S. First-principles and experimental study of nitrogen/sulfur co-doped carbon nanosheets as anodes for rechargeable sodium ion batteries. J. Mater. Chem. A 2016, 4, 15565–15574. [Google Scholar] [CrossRef]
  19. Zhang, S.; Yao, F.; Yang, L.; Zhang, F.; Xu, S. Sulfur-doped mesoporous carbon from surfactant-intercalated layered double hydroxide precursor as high-performance anode nanomaterials for both Li-ion and Na-ion batteries. Carbon 2015, 93, 143–150. [Google Scholar] [CrossRef]
  20. Chevrier, V.L.; Ceder, G. Challenges for Na-ion Negative Electrodes. J. Electrochem. Soc. 2011, 158, A1011. [Google Scholar] [CrossRef]
  21. Qin, D.; Liu, Z.; Zhao, Y.; Xu, G.; Zhang, F.; Zhang, X. A sustainable route from corn stalks to N, P-dual doping carbon sheets toward high performance sodium-ion batteries anode. Carbon 2018, 130, 664–671. [Google Scholar] [CrossRef]
  22. Gan, Q.; He, H.; Zhao, K.; He, Z.; Liu, S.; Yang, S. Plasma-Induced Oxygen Vacancies in Urchin-Like Anatase Titania Coated by Carbon for Excellent Sodium-Ion Battery Anodes. ACS Appl. Mater. Interface. 2018, 10, 7031–7042. [Google Scholar] [CrossRef] [PubMed]
  23. Lu, P.; Sun, Y.; Xiang, H.; Liang, X.; Yu, Y. 3D Amorphous Carbon with Controlled Porous and Disordered Structures as a High-Rate Anode Material for Sodium-Ion Batteries. Adv. Energy Mater. 2018, 8, 1702434. [Google Scholar] [CrossRef]
  24. Chen, T.; Ma, Y.; Guo, Q.; Yang, M.; Xia, H. A facile sol–gel route to prepare functional graphene nanosheets anchored with homogeneous cobalt sulfide nanoparticles as superb sodium-ion anodes. J. Mater. Chem. A 2017, 5, 3179–3185. [Google Scholar] [CrossRef]
  25. Xiao, Y.; Lee, S.H.; Sun, Y.-K. The Application of Metal Sulfides in Sodium Ion Batteries. Adv. Energy Mater. 2017, 7, 1601329. [Google Scholar] [CrossRef]
  26. Peng, S.; Han, X.; Li, L.; Zhu, Z.; Cheng, F.; Srinivansan, M.; Adams, S.; Ramakrishna, S. Unique Cobalt Sulfide/Reduced Graphene Oxide Composite as an Anode for Sodium-Ion Batteries with Superior Rate Capability and Long Cycling Stability. Small 2016, 12, 1359–1368. [Google Scholar] [CrossRef]
  27. Ying, H.; Han, W.-Q. Metallic Sn-Based Anode Materials: Application in High-Performance Lithium-Ion and Sodium-Ion Batteries. Adv. Sci. 2017, 4, 1700298. [Google Scholar] [CrossRef]
  28. Liu, Z.; Yu, X.-Y.; Lou, X.W.; Paik, U. Sb@C coaxial nanotubes as a superior long-life and high-rate anode for sodium ion batteries. Energy Environ. Sci. 2016, 9, 2314–2318. [Google Scholar] [CrossRef]
  29. Liu, Y.; Zhang, A.; Shen, C.; Liu, Q.; Cao, X.; Ma, Y.; Chen, L.; Lau, C.; Chen, T.-C.; Wei, F.; et al. Red Phosphorus Nanodots on Reduced Graphene Oxide as a Flexible and Ultra-Fast Anode for Sodium-Ion Batteries. ACS Nano 2017, 11, 5530–5537. [Google Scholar] [CrossRef]
  30. Wang, X.; Hwang, J.-Y.; Myung, S.-T.; Hassoun, J.; Sun, Y.-K. Graphene Decorated by Indium Sulfide Nanoparticles as High-Performance Anode for Sodium-Ion Batteries. ACS Appl. Mater. Interface. 2017, 9, 23723–23730. [Google Scholar] [CrossRef]
  31. Ge, P.; Cao, X.; Hou, H.; Li, S.; Ji, X. Rodlike Sb2Se3 Wrapped with Carbon: The Exploring of Electrochemical Properties in Sodium-Ion Batteries. ACS Appl. Mater. Interface. 2017, 9, 34979–34989. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, Y.; Kang, H.; Jiao, L.; Chen, C.; Cao, K.; Wang, Y.; Yuan, H. Exfoliated-SnS2 restacked on graphene as a high-capacity, high-rate, and long-cycle life anode for sodium ion batteries. Nanoscale 2015, 7, 1325–1332. [Google Scholar] [CrossRef] [PubMed]
  33. Deng, W.; Chen, J.; Yang, L.; Liang, X.; Yin, S.; Deng, X.; Zou, G.; Hou, H.; Ji, X. Solid Solution Metal Chalcogenides for Sodium-Ion Batteries: The Recent Advances as Anodes. Small 2021, 17, 2101058. [Google Scholar] [CrossRef]
  34. Zhang, J.; Yin, Y.-X.; Guo, Y.-G. High-Capacity Te Anode Confined in Microporous Carbon for Long-Life Na-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 27838–27844. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, K.; Ye, W.; Yin, W.; Chai, W.; Rui, Y.; Tang, B. A novel carbon-coated Ga2S3 anode material derived from post-synthesis modified MOF for high performance lithium ion and sodium ion batteries. Electrochim. Acta 2019, 322, 134790. [Google Scholar] [CrossRef]
  36. Pang, S.; Hu, Z.; Fan, C.; Zhang, W.; Cai, Y.; Han, S.; Liu, J.; Liu, J. Insights into the sodium storage mechanism of Bi2Te3 nanosheets as superior anodes for sodium-ion batteries. Nanoscale 2022, 14, 1755–1766. [Google Scholar] [CrossRef]
  37. Yin, H.; Shen, W.; Qu, H.-Q.; Li, C.; Zhu, M.-Q. Boosted charge transfer and Na-ion diffusion in cooling-fins-like Sb2Te3–Te nano-heterostructure for long cycle life and high rate capability anode. Nano Energy 2020, 70, 104468. [Google Scholar] [CrossRef]
  38. Han, X.; Jiang, Q.; Zhang, M.; Qin, Z.; Geng, H.; Sun, C.; Gu, H. Pseudocapacitance-boosted ultrafast and stable Na-storage in NiTe2 coupled with N-doped carbon nanosheets for advanced sodium-ion half/full batteries. Dalton Trans. 2021, 50, 17241–17248. [Google Scholar] [CrossRef]
  39. Cho, J.S.; Lee, S.Y.; Lee, J.-K.; Kang, Y.C. Iron Telluride-Decorated Reduced Graphene Oxide Hybrid Microspheres as Anode Materials with Improved Na-Ion Storage Properties. ACS Appl. Mater. Interfaces 2016, 8, 21343–21349. [Google Scholar] [CrossRef]
  40. Nagulapati, V.M.; Lee, J.H.; Kim, H.S.; Oh, J.; Kim, I.T.; Hur, J.; Lee, S.G. Novel hybrid binder mixture tailored to enhance the electrochemical performance of SbTe bi-metallic anode for sodium ion batteries. J. Electroanal. Chem. 2020, 865, 114160. [Google Scholar] [CrossRef]
  41. Yang, M.; Zhang, W.; Su, D.; Wen, J.; Liu, L.; Wang, X. Flexible SnTe/carbon nanofiber membrane as a free-standing anode for high-performance lithium-ion and sodium-ion batteries. J. Colloid Interface Sci. 2022, 605, 231–240. [Google Scholar] [CrossRef] [PubMed]
  42. Panda, M.R.; Raj, K.A.; Ghosh, A.; Kumar, A.; Muthuraj, D.; Sau, S.; Yu, W.; Zhang, Y.; Sinha, A.K.; Weyland, M.; et al. Blocks of molybdenum ditelluride: A high rate anode for sodium-ion battery and full cell prototype study. Nano Energy 2019, 64, 103951. [Google Scholar] [CrossRef]
  43. Liang, M.; Ma, L.; Chen, B.; Liu, E.; Shi, C.; He, C.; Zhao, N. Two Birds with One Stone: A NaCl-Assisted Strategy toward MoTe2 Nanosheets Nanoconfined in 3D Porous Carbon Network for Sodium-Ion Battery Anode. Energy Storage Mater. 2022, 47, 591–601. [Google Scholar] [CrossRef]
  44. Zhang, W.; Wang, X.; Wong, K.W.; Zhang, W.; Chen, T.; Zhao, W.; Huang, S. Rational Design of Embedded CoTe2 Nanoparticles in Freestanding N-Doped Multichannel Carbon Fibers for Sodium-Ion Batteries with Ultralong Cycle Lifespan. ACS Appl. Mater. Interfaces 2021, 13, 34134–34144. [Google Scholar] [CrossRef] [PubMed]
  45. Seo, J.-U.; Seong, G.-K.; Park, C.-M. Te/C nanocomposites for Li-Te Secondary Batteries. Sci. Rep. 2015, 5, 7969. [Google Scholar] [CrossRef]
  46. He, J.; Chen, Y.; Lv, W.; Wen, K.; Wang, Z.; Zhang, W.; Li, Y.; Qin, W.; He, W. Three-Dimensional Hierarchical Reduced Graphene Oxide/Tellurium Nanowires: A High-Performance Freestanding Cathode for Li–Te Batteries. ACS Nano 2016, 10, 8837–8842. [Google Scholar] [CrossRef]
  47. Liu, Y.; Wang, J.; Xu, Y.; Zhu, Y.; Bigio, D.; Wang, C. Lithium–tellurium batteries based on tellurium/porous carbon composite. J. Mater. Chem. A 2014, 2, 12201–12207. [Google Scholar] [CrossRef]
  48. Cho, J.S.; Ju, H.S.; Lee, J.-K.; Kang, Y.C. Carbon/two-dimensional MoTe2 core/shell-structured microspheres as an anode material for Na-ion batteries. Nanoscale 2017, 9, 1942–1950. [Google Scholar] [CrossRef]
  49. Yang, W.; Zhang, X.; Tan, H.; Yang, D.; Feng, Y.; Rui, X.; Yu, Y. Gallium-based anodes for alkali metal ion batteries. J. Energy Chem. 2021, 55, 557–571. [Google Scholar] [CrossRef]
  50. Yarema, M.; Wörle, M.; Rossell, M.D.; Erni, R.; Caputo, R.; Protesescu, L.; Kravchyk, K.V.; Dirin, D.N.; Lienau, K.; von Rohr, F.; et al. Monodisperse Colloidal Gallium Nanoparticles: Synthesis, Low Temperature Crystallization, Surface Plasmon Resonance and Li-Ion Storage. J. Am. Chem. Soc. 2014, 136, 12422–12430. [Google Scholar] [CrossRef]
  51. Jeong, J.-H.; Jung, D.-W.; Oh, E.-S. Lithium storage characteristics of a new promising gallium selenide anodic material. J. Alloys Compd. 2014, 613, 42–45. [Google Scholar] [CrossRef]
  52. Zhang, C.; Park, S.-H.; Ronan, O.; Harvey, A.; Seral-Ascaso, A.; Lin, Z.; McEvoy, N.; Boland, C.S.; Berner, N.C.; Duesberg, G.S.; et al. Enabling Flexible Heterostructures for Li-Ion Battery Anodes Based on Nanotube and Liquid-Phase Exfoliated 2D Gallium Chalcogenide Nanosheet Colloidal Solutions. Small 2017, 13, 1701677. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, W.; Chen, D.; Jiang, Y.; Feng, Y.; Rui, X.; Yu, Y. Synergetic enhancement of sodium storage in gallium-based heterostructures. Nano Energy 2021, 89, 106395. [Google Scholar] [CrossRef]
  54. Yang, M.; Sun, C.; Wang, T.; Chen, F.; Sun, M.; Zhang, L.; Shao, Y.; Wu, Y.; Hao, X. Graphene-Oxide-Assisted Synthesis of Ga2O3 Nanosheets/Reduced Graphene Oxide Nanocomposites Anodes for Advanced Alkali-Ion Batteries. ACS Appl. Energy Mater. 2018, 1, 4708–4715. [Google Scholar] [CrossRef]
  55. Wang, P.; Liu, M.; Mo, F.; Long, Z.; Fang, F.; Sun, D.; Zhou, Y.-n.; Song, Y. Exploring the sodium ion storage mechanism of gallium sulfide (Ga2S3): A combined experimental and theoretical approach. Nanoscale 2019, 11, 3208–3215. [Google Scholar] [CrossRef]
  56. Wu, Y.; Huang, L.; Huang, X.; Guo, X.; Liu, D.; Zheng, D.; Zhang, X.; Ren, R.; Qu, D.; Chen, J. A room-temperature liquid metal-based self-healing anode for lithium-ion batteries with an ultra-long cycle life. Energy Environ. Sci. 2017, 10, 1854–1861. [Google Scholar] [CrossRef]
  57. Voevodin, A.A.; Zabinski, J.S. Load-adaptive crystalline–amorphous nanocomposites. J. Mater. Sci. 1998, 33, 319–327. [Google Scholar] [CrossRef]
  58. Nguyen, Q.H.; Nguyen, Q.H.; So, S.; Hur, J. Efficient TiC-C hybrid conductive matrix for ZnTe anode in Lithium-ion storage. Appl. Surf. Sci. 2020, 534, 147679. [Google Scholar] [CrossRef]
  59. Hai Nguyen, Q.; So, S.; Hanh Nguyen, Q.; Kim, I.T.; Hur, J. Mechanochemical synthesis of InP nanoparticles embedded in hybrid conductive matrix for high-performance lithium-ion batteries. Chem. Eng. J. 2020, 399, 125826. [Google Scholar] [CrossRef]
  60. Nguyen, Q.H.; Nguyen, Q.H.; Hur, J. High-performance ZnTe-TiO2-C nanocomposite with half-cell and full-cell applications as promising anode material for Li-Ion batteries. Appl. Surf. Sci. 2020, 509, 144718. [Google Scholar] [CrossRef]
  61. Huy, V.P.H.; So, S.; Kim, I.T.; Hur, J. Self-healing gallium phosphide embedded in a hybrid matrix for high-performance Li-ion batteries. Energy Storage Mater. 2021, 34, 669–681. [Google Scholar] [CrossRef]
  62. Hieu, L.T.; So, S.; Kim, I.T.; Hur, J. Highly reversible lithiation/delithiation in indium antimonide with hybrid buffering matrix. Int. J. Energy Res. 2021, 45, 16145–16154. [Google Scholar] [CrossRef]
  63. Fan, T.-E.; Xie, H.-F. Sb2S3-rGO for high-performance sodium-ion battery anodes on Al and Cu foil current collector. J. Alloy. Compd. 2019, 775, 549–553. [Google Scholar] [CrossRef]
  64. Deng, M.; Li, S.; Hong, W.; Jiang, Y.; Xu, W.; Shuai, H.; Li, H.; Wang, W.; Hou, H.; Ji, X. Natural stibnite ore (Sb2S3) embedded in sulfur-doped carbon sheets: Enhanced electrochemical properties as anode for sodium ions storage. RSC Adv. 2019, 9, 15210–15216. [Google Scholar] [CrossRef]
  65. Lin, J.; Yao, L.; Zhang, C.; Ding, H.; Wu, Y.; Li, S.; Han, J.; Yue, G.; Peng, D. Construction of Sb2S3@SnS@C Tubular Heterostructures as High-Performance Anode Materials for Sodium-Ion Batteries. ACS Sustain. Chem. Eng. 2021, 9, 11280–11289. [Google Scholar] [CrossRef]
  66. Xie, J.; Liu, L.; Xia, J.; Zhang, Y.; Li, M.; Ouyang, Y.; Nie, S.; Wang, X. Template-Free Synthesis of Sb2S3 Hollow Microspheres as Anode Materials for Lithium-Ion and Sodium-Ion Batteries. Nano-Micro Lett. 2017, 10, 12. [Google Scholar] [CrossRef]
  67. Tang, X.; Huang, X.; Huang, Y.; Gou, Y.; Pastore, J.; Yang, Y.; Xiong, Y.; Qian, J.; Brock, J.D.; Lu, J.; et al. High-Performance Ga2O3 Anode for Lithium-Ion Batteries. ACS Appl. Mater. Interface. 2018, 10, 5519–5526. [Google Scholar] [CrossRef]
  68. Yuan, Y.; Yang, M.; Liu, L.; Xia, J.; Yan, H.; Liu, J.; Wen, J.; Zhang, Y.; Wang, X. The electrochemical storage mechanism of an In2S3/C nanofiber anode for high-performance Li-ion and Na-ion batteries. Nanoscale 2020, 12, 20337–20346. [Google Scholar] [CrossRef]
  69. Liu, B.; Cao, J.; Li, J.; Li, L.; Chen, D.; Zhang, S.; Cai, D.; Han, W. Highly conductive Co3Se4 embedded in N-doped 3D interconnected carbonaceous network for enhanced lithium and sodium storage. J. Colloid Interface. Sci. 2021, 586, 630–639. [Google Scholar] [CrossRef]
  70. Mahmood, A.; Ali, Z.; Tabassum, H.; Akram, A.; Aftab, W.; Ali, R.; Khan, M.W.; Loomba, S.; Alluqmani, A.; Adil Riaz, M.; et al. Carbon Fibers Embedded With Iron Selenide (Fe3Se4) as Anode for High-Performance Sodium and Potassium Ion Batteries. Front. Chem. 2020, 8, 408. [Google Scholar] [CrossRef]
  71. Nagulapati, V.M.; Kim, D.S.; Oh, J.; Lee, J.H.; Hur, J.; Kim, I.T.; Lee, S.G. Enhancing the Electrochemical Performance of SbTe Bimetallic Anodes for High-Performance Sodium-Ion Batteries: Roles of the Binder and Carbon Support Matrix. Nanomaterials 2019, 9, 1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) XRD pattern of Ga2Te3–TiO2; (b) EDX spectrum of Ga2Te3–TiO2.
Figure 1. (a) XRD pattern of Ga2Te3–TiO2; (b) EDX spectrum of Ga2Te3–TiO2.
Materials 15 06231 g001
Figure 2. (a) SEM image; (b) size distribution; (c) HRTEM image; (d) EDXS elemental mappings of Ga, Te, Ti, O, and C in Ga2Te3–TiO2–C(10%).
Figure 2. (a) SEM image; (b) size distribution; (c) HRTEM image; (d) EDXS elemental mappings of Ga, Te, Ti, O, and C in Ga2Te3–TiO2–C(10%).
Materials 15 06231 g002
Figure 3. Electrochemical performances of Ga2Te3–TiO2–C composites for SIBs: (a) GCD profiles of Ga2Te3–TiO2–C(10%) at 100 mA·g−1; (b) cycling performances of composite at 100 mA·g−1 and (c) 500 mA·g−1; (d) CV curves of Ga2Te3–TiO2–C(10%); (e) rate capabilities of Ga2Te3–TiO2–C composites; (f) capacity retention of Ga2Te3–TiO2–C composites from 0.1 to 10 A·g−1.
Figure 3. Electrochemical performances of Ga2Te3–TiO2–C composites for SIBs: (a) GCD profiles of Ga2Te3–TiO2–C(10%) at 100 mA·g−1; (b) cycling performances of composite at 100 mA·g−1 and (c) 500 mA·g−1; (d) CV curves of Ga2Te3–TiO2–C(10%); (e) rate capabilities of Ga2Te3–TiO2–C composites; (f) capacity retention of Ga2Te3–TiO2–C composites from 0.1 to 10 A·g−1.
Materials 15 06231 g003
Figure 4. (a,b) Ex situ XRD patterns obtained at selected cutoff potentials in initial sodiation/desodiation process; (c) schematic of electrochemical reaction mechanism of Ga2Te3–TiO2–C(10%) electrode during cycling.
Figure 4. (a,b) Ex situ XRD patterns obtained at selected cutoff potentials in initial sodiation/desodiation process; (c) schematic of electrochemical reaction mechanism of Ga2Te3–TiO2–C(10%) electrode during cycling.
Materials 15 06231 g004
Figure 5. EIS-based Nyquist plots for (a) Ga2Te3–TiO2–C(10%), (b) Ga2Te3–TiO2–C(20%), and (c) Ga2Te3–TiO2–C(30%) after one, five, and 20 cycles; (d) equivalent circuit.
Figure 5. EIS-based Nyquist plots for (a) Ga2Te3–TiO2–C(10%), (b) Ga2Te3–TiO2–C(20%), and (c) Ga2Te3–TiO2–C(30%) after one, five, and 20 cycles; (d) equivalent circuit.
Materials 15 06231 g005
Table 1. Comparison of performances of chalcogenide-based anodes for SIBs.
Table 1. Comparison of performances of chalcogenide-based anodes for SIBs.
AnodeCycling PerformanceRate CapabilitySynthesis MethodRef.
Ga2S3476 mAh·g−1 after 100 cycles at 0.4 A·g−1283 mAh·g−1 at 2 A·g−1Vapor thermal annealing[55]
Ga2S3–C385 mAh·g−1 after 200 cycles at 0.1 A·g−194 mAh·g−1 at 2.0 A·g−1Sulfuration process[35]
Sb2Se3/C485.2 mAh·g−1 after 100 cycles at 0.2 A·g−1237.9 mAh·g−1 at 2.0 A·g−1Hydrothermal process[31]
Sb2S3–rGO537 mAh·g−1 after 70 cycles at 0.1 A·g−1290 mAh·g−1 at 3.2 A·g−1Ultrasonication method[63]
Sb2S3–C455.8 mAh·g−1 after 100 cycles at 0.1 A·g−1263 mAh·g−1 at 1.0 A·g−1Modified natural stibnite ore[64]
Sb2S3@SnS@C437 mAh·g−1 after 100 cycles at 1 A·g−1448 mAh·g−1 at 5.0 A·g−1Hydrothermal method[65]
Sb2S3384 mAh·g−1 after 50 cycles at 0.2 A·g−1239 mAh·g−1 at 5.0 A·g−1Hydrothermal method[66]
In2S3/C372 mAh·g−1 after 200 cycles at 0.5 A·g−1236 mAh·g−1 at 2.0 A·g−1Electrospinning process[67,68]
Co3Se4@C449 mAh·g−1 after 20 cycles at 0.1 A·g−1328 mAh·g−1 at 5.0 A·g−1Annealing process[69]
Fe3Se4@C439 mAh·g−1 after 25 cycles at 0.05 A·g−1-Electrospinning process[70]
Bi2Te3364 mAh·g−1 after 1200 cycles at 5 A·g−1339 mAh·g−1 at 10 A·g−1Chemical reduction method[36]
SbTe–C421 mAh·g−1 after 200 cycles at 0.1 A·g−1413 mAh·g−1 at 1 A·g−1 Ball milling[71]
Ga2Te3–TiO2–C 437 mAh·g−1 after 300 cycles at 0.1 A·g−1318 mAh·g−1 at 10 A·g−1Ball millingThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huy, V.P.H.; Kim, I.T.; Hur, J. Ga2Te3-Based Composite Anodes for High-Performance Sodium-Ion Batteries. Materials 2022, 15, 6231. https://doi.org/10.3390/ma15186231

AMA Style

Huy VPH, Kim IT, Hur J. Ga2Te3-Based Composite Anodes for High-Performance Sodium-Ion Batteries. Materials. 2022; 15(18):6231. https://doi.org/10.3390/ma15186231

Chicago/Turabian Style

Huy, Vo Pham Hoang, Il Tae Kim, and Jaehyun Hur. 2022. "Ga2Te3-Based Composite Anodes for High-Performance Sodium-Ion Batteries" Materials 15, no. 18: 6231. https://doi.org/10.3390/ma15186231

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