Next Article in Journal
Geological Modeling of Shale Oil in Member 7 of the Yanchang Formation, Heshui South Area, Ordos Basin
Previous Article in Journal
Reducing the Cooling Energy Demand by Optimizing the Airflow Distribution in a Ventilated Roof: Numerical Study for an Existing Residential Building and Applicability Map
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of TiO2 on the Electrochemical Performance of Sb2O3 Anodes for Li-Ion Batteries

1
Mechanical Engineering Department, University of Texas Rio Grande Valley, Edinburg, TX 78539, USA
2
School of Earth, Environmental, and Marine Sciences, University of Texas, Rio Grande Valley, 1 W University Blvd, Brownsville, TX 78521, USA
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(15), 6598; https://doi.org/10.3390/app14156598 (registering DOI)
Submission received: 23 May 2024 / Revised: 12 July 2024 / Accepted: 17 July 2024 / Published: 28 July 2024

Abstract

:
Antimony (Sb) and its composites have been recognized as potentially good anode materials for lithium-ion batteries (LIBs) due to their relatively high theoretical capacity of 660 mAh g−1 and to their low cost. However, Sb-based anodes suffer from a high-volume change during the lithiation/delithiation process that results in capacity fading and anode degradation after prolonged charge/discharge cycles. To address this issue, Sb2O3/TiO2 nanocomposite electrodes can be synthesized and used as anodes for LIBs with high capacity and good electrochemical stability. In the present work, TiO2@Sb2O3 composites with different (TiO2:Sb2O3) ratios of 0:1, 1:1, 1:4 and 3:1 were synthesized and directly used as anode materials for LIBs. The electrochemical performance of the TiO2/Sb2O3 composite anode with different ratios of TiO2 to Sb2O3 was evaluated by galvanostatic charge/discharge, rate performance and cyclic voltammetry. The 3:1 (TiO2:Sb2O3) composite anode delivered the highest capacity compared to those of the TiO2, SbO3, 1:1 (TiO2:Sb2O3) and 1:4 (TiO2:Sb2O3) electrodes. The TiO2@Sb2O3 composite anode with a 3:1 ratio exhibited a stabilized capacity of 536 mAh g−1 after 100 cycles at 100 mA g−1 and showed excellent rate performance, with current densities between 50 and 500 mA g−1. The improved electrochemical performance was attributed to the synergistic effect of TiO2 (i.e., the coating of Sb2O3 with TiO2) on reducing the volume change of the Sb anode material after prolonged charge/discharge cycles and on maintaining a stable interface between the electrolyte and the composite electrode material.

1. Introduction

Current commercial rechargeable lithium-ion batteries (LIBs) use graphite as the anode material due to its high electrical conductivity, natural abundance and long cycle life [1,2]. However, the major challenges facing the graphite anode are its theoretical capacity of 372 mAh g−1, low practical energy density and restricted rate capability, especially at high current densities [3,4,5]. Methods to improve the electrochemical performance of existing LIBs are widely studied in order to fulfill the need for energy storage devices with high energy density and specific power capacity for electric vehicles. This can be accomplished by developing more stable electrolytes, improving manufacturing techniques and/or replacing/developing a better anode material than graphite [6,7]. Si, Sn, Sb and metal oxides have been widely used and examined as potential anodes for LIBs due to their high charge/discharge capacity and improved rate performance [8,9,10]. Metallic alloys, metal-oxide/carbon composites and mixed metal sulfides have also been used as anode materials for both LIBs and sodium ion batteries due to their high capacity and good rate performance [11,12,13,14,15,16,17]. For example, Murugesan et al. [18] synthesized copper-coated amorphous silicon particles via the polyol reduction method for use as an anode material in LIBs. The Si/C composite anode delivered a capacity of 600 mAh g−1 after 40 cycles at 100 mA g−1. Coating Sn electrodes with carbon has also demonstrated the desired higher specific capacities compared to graphite [19]. For example, Li et al. used an electrostatic spray deposition (ESD) to fabricate porous Sn@Carbon composites at 900 °C, which delivered a discharge capacity of 638 mAh g−1 after 315 cycles at 25 mA g−1 [20].
Nonetheless, Sb composite electrodes are highly attractive for use as anodes in LIBs due to their lower volume expansion of 135% than that of Si anodes (300–400%) [21]. Sb-based composite anode materials exhibit a high theoretical capacity of 660 mAh g−1 [22] and a good thermal stability similar to Sn and Si composites [23]. Novel methods have been used to prepare hollow and nano-sized structures of Sb as anodes for LIBs with improved electrochemical performance [24,25,26,27,28]. In fact, Sb has been a striking anode material for LIBs since it undergoes a simpler phase transformation compared to crystallogens (Sn, Ge, and Si) when alloying with Li [29]. In addition to metallic Sb and Sb-alloy materials, antimony trioxide (Sb2O3) has received great attention as a potential anode for LIBs due to its high theoretical specific capacity of 1109 mAh g−1 and its good structural stability (during and after cycling) compared with binary antimony oxides of Sb2O4 and Sb2O5 [30,31,32]. However, the Sb2O3 anode exhibits a high-volume change after prolonged charge/discharge cycles, thus leading to capacity degradation and poor electrochemical performance of the electrode. Many efforts have been made to address this issue using different synthesis methods. For example, Tan et al. successfully synthesized a bundle-shaped Sb2O3 as anode material for LIBs by preparing a Sb-MOF template through thermal annealing, which exhibited a specific capacity of 594.1 mAh g−1 after 40 cycles at 20 mA g−1 [33]. Liu et al. synthesized a polymer binder to prepare a CMC-FA@Sb2O3 electrode that delivered a capacity of 611.4 mAh g−1 after 200 cycles at 200 mA g−1 [34]. Chen et al., on the other hand, synthesized a mesoporous structure of Sb2O3 and rGO (reduced graphene oxide) via thermal decomposition [32]. The Sb2O3/rGO composite exhibited a reversible capacity of 513 mAh g−1 after 300 cycles at 500 mA g−1. The methods mentioned above showed great advancements towards supporting Sb2O3 as a potential anode material.
Nanostructured titanium dioxide (TiO2) materials have also been used as anodes for LIBs [35,36]. TiO2 is known for its synergistic effect on the electrochemical performance of TiO2-based composite electrodes of lowering the volume change during and after cycling, since it shows a low volume change of 5% after prolonged charge/discharge cycles [37]. Results reported in the literature showed that introducing the TiO2 phase into the TiO2-based composite electrode resulted in improved electrochemical performance of the electrode, which was attributed to the synergistic effect of TiO2 on reducing the diffusion resistance, enhancing the interfacial electron transfer and lowering the volume change of the electrode [38,39]. In fact, introducing TiO2 into the TiO2/Sb2O3 system can cause the development of an internal field and/or condense phase interface that initiates the synergy between TiO2 and Sb2O3 [38,39].
In addition, TiO2 is known to show a high charge/discharge plateau (operating voltage)—higher than 1.5 V—which can prevent the reaction with the electrolyte and of dendrite formation, thus resulting in good electrochemical stability of the TiO2-based composite electrode [40,41]. For instance, composites of TiO2 alloyed with transition metal oxides such as SnO2, MoO2 and Fe2O3 have been widely used as anode materials for LIBs due to their low volume change (<5%), safety, low cost and long cycle [42] life [42,43,44]. Yang et al. prepared biomass-derived carbon/MoO2@TiO2 anode material via hydrothermal processing [45]. The MoO2/TiO2/C composite electrode exhibited a near steady specific capacity of 600 mAh g−1 at 100 mA g−1 after 100 cycles. Gonzalez et al. synthesized SnO2/TiO2 micro belt-fibers for use as an anode material in LIBs [46]. The SnO2/TiO2 composite-fiber anode delivered a capacity of 279 mAh g−1 after 70 cycles at 100 mA g−1 and demonstrated better rate capability than pure SnO2. The addition of TiO2 to the SnO2 precursor solution resulted in an efficient network of SnO2/TiO2 micro-belt fibers with better lithium-ion transport and storage [46]. Off-stoichiometric TiO2−x nanoparticles were used to decorate the surface of graphite [47]. The TiO2-coated strategy led to the formation of different phases of TiO2, such as anatase and rutile, which resulted in improving the electrochemical performance of the graphite anode [47]. In fact, the electrochemical performance of TiO2 composite anodes can depend on the type of TiO2 phase in the composite anode [42]. There are many different reasons for the generation of different phases of TiO2, some as simple as the starting materials and reaction temperature. For example, using sodium sulfate and TiCl3, one can consistently form an anatase phase at 100 °C. However, by changing the sodium sulfate for sodium acetate one can consistently synthesis brookite [48].
Wang et al. reported results on the synthesis of a 1D Sb2O3@TiO2 hollow composite using the Kirkendall method for use as an anode in LIBs TiO2 [49]. The Sb2O3@TiO2 composite electrode delivered a high reversible capacity of 593 mAh g−1 after 100 cycles at 100 mA g−1. After 600 cycles, the composite electrode retained a discharge capacity of 439 mAh g−1 at 500 mA g−1. Han et al. effectively synthesized a Sb2O3/Sb@ TiO2 nanocomposite by coating TiO2 onto the surface of Sb2O3 nanorods [50]. The composite anode delivered a specific discharge capacity of 609 mAh g−1 after 100 cycles at 100 mA g−1. The improved electrochemical performance of the Sb2O3/Sb@ TiO2 nanocomposite anode was attributed to the synergistic effect of TiO2 on the volume expansion/contraction of the Sb2O3/Sb matrix after 100 charge/discharge cycles. In fact, the TiO2 phase, with low volume change and low capacity, can reduce the volume change and hence increase the capacity of the Sb2O3 electrode.
In the present work, TiO2/Sb2O3 composite material with different ratios of TiO2 to Sb2O3 was prepared by the hydrolysis of TiCl4 and subsequent hydrothermal treatment at 175 °C for use as anodes in Li-ion half cells. The hydrothermal synthesis method proposed in this work is very simple and novel and has never been used for similar systems that can lead to the mass production of the Sb2O3/TiO2 composite anodes. The morphology and characterization of the composite electrode was investigated via scanning electron microscope (SEM), energy dispersive X-ray spectrometer (EDS), X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). The electrochemical performance of the TiO2/Sb2O3 composite anode with different (TiO2:Sb2O3) ratios of 0:1, 1:1, 1:4 and 3:1 was evaluated by its galvanostatic charge/discharge, rate performance and cyclic voltammetry. The TiO2/Sb2O3 composite anode with a 3:1 ratio exhibited improved electrochemical performance and cycling stability compared to the pristine Sb electrode. The present work focuses on the hydrothermal synthesis of Sb2O3 microparticles coated with TiO2 nanoparticles at 175 °C. To the best of our knowledge, there are no available results reported in the literature on similar systems, specifically on those with similar structure and morphology to the Sb2O3/TiO2 nanocomposite anode. Additionally, high ratios of TiO2 to Sb2O3 were also used since the size of the Sb2O3 particles was much larger than the TiO2 nanoparticles. This is the first time such results have been reported on TiO2/Sb2O3 composite electrode systems.

2. Experimental

2.1. Materials and Characterization

Antimony (III) oxide (<250 nm), Poly(acrylonitrile) (PAN) (MW 150,000) and conductive carbon black were purchased from Sigma Aldrich USA (St. Louis, MO, USA). N, N-dimethyl Formamide (DMF) was obtained from Fisher Scientific Chemical USA (Waltham, MA, USA). Lithium salt (LiPF6) and commercial copper foil with 9µm of thickness were purchased from MTI (Richmond, CA, USA). Whatman glass microfibers were purchased from GE Healthcare and were punched into separators. The electrolyte used in this study was prepared in the lab by applying 1 M LiPF6 in a 1:1 v/v solution of ethylene carbonate (EC)/dimethyl carbonate (DMC), both of which were purchased from Sigma-Aldrich (USA). TiCl4 and NaCH3CH2OO were provided by Alfa Aesar (Haverhill, MA, USA).
The morphology and structural characterization of the TiO2/Sb2O3 nanocomposite anode were studied using a Sigma VP Carl Zeiss scanning electron microscope (SEM) (Jena, Germany) equipped with an energy dispersive spectrometer (EDS) from EDAX. XPS (X-ray photoelectron spectroscopy) and XRD (X-ray diffraction) were used to study the surface chemistry and the crystal structure of the samples. The XPS data were collected using a Thermo scientific Kα instrument that uses a single monochromatic Al Kα radiation. The XRD patterns were collected using a Bruker D2 X-ray diffractometer (Billerica, MA, USA) equipped with a Co source (Kα = 1.789 Å), an Fe filter, a step of 0.01°, a counting time of 5 s and a scintillation detector. The XRD patterns were fitted using the LeBail fitting procedure and crystallographic data from the literature [51,52].

2.2. Preparation of TiO2@Sb2O3 Nanocomposite Material

The TiO2 nanomaterial was prepared via the hydrolysis of TiCl4 followed by hydrothermal treatment at 175 °C. The water used was ultrapure 18 mega ohm Millipore water. The ratio of water to TiCl4 was 5:1, by volume of 50 mL of TiCl4 and 250 mL of water. The TiCl4 was added to chilled water under vigorous stirring in a 0.5 L three-neck round bottom. The water was cooled in an ice bath. The TiCl4 was added quickly to the water with the aid of an additional funnel. The solution was then placed into autoclaves, which were filled up to 80% of the flask, sealed and heated to 175 °C and held constant for 3 h. The TiO2@Sb2O3 composites with 1:1, 3:1 and 1:4 ratios were synthesized by dispersing 1.0 g of Sb2O3 in water and cooled to 4 °C, and 15 mL of previously prepared TiCl4 solution was then added and stirred. The TiCl4-Sb2O3 suspension was then added to an autoclave, heated to 175 °C and held constant for 3 h. Subsequent to reaction, the mixture was cooled to room temperature naturally upon opening the autoclave and a white precipitate was observed and then extracted from the Teflon bombs and washed with Millipore water. The washed white precipitate was then dried in a 50 °C oven overnight. In fact, the synthesis method used in this work is simple and novel and has never been used for similar systems.

2.3. Electrochemical Performance

2.3.1. Preparation of Electrode Solution

The TiO2 and TiO2/Sb2O3 nanocomposites with different TiO2:Sb2O3 ratios were homogenized using a mortar and pestle. A slurry was prepared at an 8:1:1 composition ratio of active material, conductive carbon black and Polyacrylonitrile (PAN) binder using DMF as the dispersant. To achieve homogeneity in the slurry, the solution was magnetically stirred for 24 h at room temperature. The slurry was then coated on a copper foil at a thickness of 10 microns and placed inside a Fisher Scientific Isotemp vacuum oven (model 285A) for 24 h at 60 °C. Once dried, the slurry was placed in a tube furnace (OTF-1200X, MTI Corporation) for a heat treatment at 350 °C (ramp rate was 2 °C/min) for 5 h under argon. Subsequent to heating, the copper coated with slurry was cut into 0.5”diameter discs and weighed to determine the amount of active material in the electrode.

2.3.2. Li-ion Half-Cell Assembly

Li-ion half cells were assembled inside a MBRAUN LAB star pro glovebox under an argon atmosphere with O2 and H2O concentrations < 0.5 ppm. The components used were CR 2032 (PHD Energy Inc.) coin cells, a Li+ chip (CR 2032 Ph.D. Energy Inc., Georgetown, TX, USA) as counter electrode and glass microfibers as the separator. To properly assemble the Li-ion half cells, an MSK-110 hydraulic crimping machine from MTI Corporation was operated at 1000 psi. Cyclic voltammetry was performed using Biologic Science Instruments (Grenoble, France) at a scan rate of 0.1 mVs−1 and with a voltage range between 0.01 and 3.0 V. Galvanostatic charge–discharge experiments were performed using the LANHE battery testing system (CT2001A) over 100 cycles while the rate performance was evaluated using the Arbin Instrument (Bt2000) (College Station, TX, USA).

3. Results and Discussion

3.1. Characterization

Figure 1 displays low and high magnification SEM images of the synthesized Sb2O3, TiO2 and TiO2/Sb2O3 with a (3:1) ratio after synthesis. The SEM images of the Sb2O3 show that the sample consisted of chunks of material of various sizes (Figure 1A,B) while the TiO2 (Figure 2C,D) exhibited a sponge-like texture (Figure 1C). However, upon closer examination at higher magnification, the particles exhibited a small-needles morphology (Figure 1D) while the (3:1) TiO2/Sb2O3 displayed more disordered and randomized morphology and appeared to consist of collections of small spherical particles (Figure 1E,F). The elemental composition of each material was analyzed by EDS and is shown in Figure 2, Figure 3 and Figure 4A. The EDS mappings in Figure 2, Figure 3 and Figure 4 confirm the presence of the Sb, Ti and O elements in the sample. The Cl element was observed in the EDS spectrum but is not mapped in the sample since the concentration of Cl was too low to generate a clear elemental map.
The elemental composition was determined using EDS and showed that the sample composition was 3.1% Cl, 61.6% O, 8.9% Sb and 26.4% Ti. The ratio between Sb and Ti was 2.97, indicating that the correct ratio was achieved from the synthesis. The data also indicated that the sample was not a pure mixture of TiO2 and Sb2O3. More importantly, the correlation between the elements in the maps shows a very homogenous sample, especially for the composite material. The EDS spectra for TiO2, Sb2O3 and (3:1) TiO2/Sb2O3 samples are included in the SI file of the revised manuscript (Figure S6).
Figure 5 shows the X-ray diffraction patterns for the (3:1) TiO2@Sb2O3 synthesized sample, the TiO2—as synthesized hydrothermally at 175 °C—and the Sb2O3 starting material. The LeBail fitting results of the samples are shown in Table 1. The TiO2@Sb2O3 samples showed the development of the Antimony trichloride (Sb4O5Cl2) phase of Sb mixed with the rutile phase of TiO2, as can be seen in Figure 5A and Table 1. It is important to note here that the Sb2O3 was purchased and added to TiCl4 in an aqueous solution and heated to 175 °C with the aim of coating TiO2 on the Sb2O3 particle surface. In fact, Sb2O3 can dissolve in aqueous solutions and, as a result, Sb2O3 and Sb4O5Cl2 coated with TiO2 were formed after the heat treatment at 175 °C. However, the amount of Cl in the Sb compound is small, which will not affect the electrochemical performance of the TiO2/Sb2O3 composite electrodes. The XRD results in this work (Figure S5) show that for the (1:4) TiO2/Sb2O3 sample, most of the Sb compound was Sb2O3, while for the 1:1 TiO2/Sb2O3 electrode, a mixture of Sb2O3 and Sb4O5Cl2 coated with TiO2 was formed after the hydrothermal process at 175 °C.
The XRD results of the Sb2O3:TiO2 with a 3:1 ratio showed the P21/C symmetry for the Sb4O5Cl2 phase and the rutile phase for the TiO2 with the P42/mnm space group, which is consistent with XRD data reported in the literature [53]. The overall reduced χ2 of the fitting was 1.28, which showed an excellent agreement between the data and the fitting. The TiO2 nanomaterial synthesized using the hydrothermal technique showed the presence of the rutile phase in the sample (Figure 5B), with a space group of P42/mnm where lattice parameters were consistent with the literature values (Table 1). The χ2 of the fitting was 2.41 [52]. The XRD results of Sb2O3 material showed the presence of two phases: the bulk of the Sb2O3 was present as the cubic phase with the space group Fd-3m, and the minor phase was the orthorhombic phase with the Pccn space group (Figure 5C). These two phases have lattice parameters matching those reported in the literature in which the overall fitting had a χ2 of 3.15 [54,55]. The χ2 values being below 5 showed good agreement between the literature and the collected data. More importantly, the structure and morphology of the TiO2/Sb2O3 composite are very different from what has been reported on similar systems [37,49]. The synthesis method used in this work is very simple, which can result in the implementation of this method by industry to mass-produce the Sb2O3/TiO2 composite anode. In addition, the simplicity of the methodology ensures that the studies are reproducible and thus that the results are reliable.
Figure 6A shows the XPS for the hydrothermally synthesized TiO2 nanoparticles at 175 °C. The XPS spectrum shows the presence of three peaks, located at 458.6, 464.3 and 472.0 eV, which corresponds to the Ti 2P3/2, Ti2P1/2 and the Ti 2P satellite peaks, respectively. The XPS peaks were determined to consist of individual peaks and no further deconvolution was necessary for the fitting. The observed peak energies were consistent with those observed on a rutile phase of TiO2 [56,57]. Figure 6B shows the Ti 2P for the synthesized TiO2@Sb2O3 (3:1) composite. The data show three peaks in the Ti 2P region, located at 458.6, 464.3 and 472.0 eV. The Ti peaks corresponded to the Ti 2P3/2, Ti2P1/2 and the Ti 2P satellite peaks, respectively. The energies for the Ti correspond to Ti in the 4+ oxidation state bound to oxygen as would be present in TiO2 [56,57].
Figure 7A shows the XPS spectrum of O1S for the hydrothermally synthesized TiO2. It was deconvolved into two peaks at 529.8 and 531.1 eV, representing the Ti-O bond and OH bound to the surface, respectively. Figure 7B shows the XPS of the Sb3d for the 3:1 TiO2@ Sb2O3 sample indicating the presence of four peaks, which were further deconvolved into five peaks. The fifth peak was determined to be from the O 1S spectrum, which overlapped with the Sb 3d. The binding energies of the individual peaks were located at 528.3, 530.7, 531.4, 537.7 and 540.3 eV, corresponding to the Sb (0) 3d5/2, Sb (III) 3d5/2, O 1S, Sb (0) 3d3/2 and the Sb(III)3d3/2, respectively [56,58,59,60,61,62]. The binding energies for the Ti2P, O1S and Sb 3d XPS fittings are summarized in Table 2.

3.2. Electrochemical Performance

Figure 8 shows the cyclic voltammetry (CV) curves of the initial five cycles for the TiO2, Sb2O3 and TiO2/Sb2O3 electrodes at a scanning rate of 0.1 mVs−1 between 0.01 and 3.0 V (vs. Li+/Li) The CV profiles of pure TiO2 (Figure 8A) show similar results to those observed in previous work [63,64,65]. As shown in Figure 8B, the CV profiles of the Sb2O3 electrode show that the first cathodic scan is different form the subsequent cathodic cycles, which can be attributed to the formation of the solid electrolyte interface (SEI) layer. The two cathodic peaks at 1 V and 0.6 V could be attributed to the conversion process from Sb2O3 to metallic Sb and to the formation of the Li3Sb during the alloying reaction process, respectively [31]. The peaks from the 1st to the 5th cycles in the anodic scan, located at 1.1 V and 1.3 V, correspond to the dealloying and delithiation reactions. These CV results are in agreement with those reported on the Sb2O3 electrodes [31,66]. The CV results of the TiO2/Sb2O3 electrode (Figure 8C) show that at the first cathodic scan, a distinct peak at 1.71 V indicated the transition of Sb2O3 to metallic Sb and Li2O (as shown in Equation (1). In the subsequent cathodic scans, a weak reduction peak was observed at 2.5 V, indicating the transformation of Sb2O3 to Sb metal. The reduction peaks at 0.44 and 1.07 V can be attributed to a series of electrochemical reactions, including the alloying of metallic Sb to LixSb and the formation of the SEI layer at the surface of the electrode that is caused by the electrolyte decomposition [34,49,67]. After the 1st cycle, these two peaks, at 1.71 and 0.44 V, were reversible and shifted slightly to the higher voltages, indicating the formation of a stable SEI film during the first cathodic scan [68]. This shift signified that the material was gradually activating as lithium was being deposited via insertion, which simultaneously increased the voltage. The CV profile of the (3:1) TiO2/Sb2O3 electrode shows complex redox reactions, indicating an activation process in the initial lithiation process that was also observed in the literature [67,68]. The CV results of different TiO2 to Sb2O3 ratios (1:1 and 1:4) are shown in the SI file Fig (Sx). The CV profiles clearly indicate that on increasing the Sb2O3 amount in the composite electrode, the reactions and alloying/dealloying of Sb take place at below 1 V. Also, the formation of Sb5O4Cl2 after heat treatment can also result in a complex CV profile of the composite electrode. More work is needed to understand the complexity of the Sb redox reactions discussed in this work.
The electrochemical reactions of the Sb2O3 anode are described in Equations (1) and (2).
Sb2O3 + 6Li+ + 6e → 2Sb + 3Li2O
Sb + XLi+ + Xe → LixSb, x ≤ 3
At the first anodic scan (Figure 8C), the oxidation peak at 1.16 V is credited towards the dealloying of LixSb to Sb [69], whereas the other peak at 2.03 V was attributed to TiO2, which is in agreement with results reported by Jeong et al. [70] of repeatedly having a peak along the range of 2–2.5 V. Lastly, the peak at 2.46 V was attributed to the pyrolysis of the PAN binder during the heat treatment of the TiO2/Sb2O3 slurry [71]. The results presented in Figure 8 indicate that the TiO2/Sb2O3 composite anode demonstrated good electrochemical stability and reversibility during charge/discharge cycles, which is indicated by the constant overlapping of CV scans.
Figures S1 and S2 show the CV scans at the first five cycles of the TiO2/Sb2O3 electrodes with 1:1 and 1:4 ratios of TiO2 to Sb2O3, as shown in the supplementary information (Figures S1 and S2).
Figure 9 shows the charge/discharge curves of the Sb2O3, TiO2 and TiO2/Sb2O3 electrodes after 100 cycles at a current density of 100 mA g−1 and with potentials ranging between 0.01 and 3.0 V (vs. Li+/Li). During the first discharge cycle (Figure 9a), a small plateau was present at around 1.70 V—which was related to the transition of Sb2O3 to metallic Sb and Li2O—and a visible plateau at 1.25 V was observed that was ascribed to the formation of the SEI layer. At the first charge cycle, three minimal voltage plateaus were observed at 1.16 V, 2.0 V and 2.46 V, which corresponded to the dealloying of LixSb and the forming of Sb2O3 and its interaction with TiO2 and PAN. The initial discharge and charge capacities were 1002 mAh g−1 and 632 mAh g−1, experiencing a capacity loss of 35% from the 1st discharge to the 2nd cycle. In subsequent cycles, the capacity was stable and experienced a 2.9% loss from the 50th to the 100th cycle, while the anode remarkably exhibited a specific capacity of 544 mAh g−1 after 100 cycles. Based on the CV scans and charge discharge profile for the TiO2/Sb2O3 composite electrode, the TiO2 phase tremendously assisted with the stability and the retained sufficient capacity of the Sb anode.
Figure 10 shows the cycling performance and coulombic efficiency of Sb2O3, TiO2 and (3:1) TiO2/Sb2O3 nanocomposite electrodes after 100 cycles at a current of 100 mA g−1. The (3:1) TiO2/Sb2O3 nanocomposite electrode displayed an initial charge capacity of 632 mAh g−1, corresponding to a coulombic efficiency of 63.1%, which is higher in comparison to the pristine Sb2O3 material and to commercial graphite. It is observed in Figure 10 that TiO2 had a specific capacity of 175 mAh g−1 at the first charge cycle and a low coulombic efficiency of 47.2%. The TiO2 electrode had a steady specific capacity yet maintained a low charge capacity. The Sb2O3 electrode exhibited a high charge specific capacity of 829 mAh g−1 in the first few cycles, corresponding to a coulombic efficiency of 83.2%, but it experienced a large capacity loss starting at the 15th cycle. After the 2nd cycle, the coulombic efficiency for all electrodes was over 99%, indicating good cycling stability and capacity retention.
Noticeably, the (3:1) TiO2/Sb2O3 composite electrode outperformed both materials and demonstrated a good cyclic performance, maintaining a stable specific capacity between the 30th and 70th cycles. And, remarkably, by the 77th cycle its specific capacity increased to 521 mAh g−1, eventually finishing at 536 mAh g−1 by its 100th cycle. It can be assumed that the thickness of the TiO2 coating along the surface of the Sb2O3 particles was significant enough to act as a buffer when undergoing high volume changes.
Figures S3 and S4 show the charge/discharge profile and cycling performance of the TiO2/Sb2O3 composite electrode with 1:1 and 1:4 ratios. The results in Figures S3 and S4 show a capacity degradation (capacity fading) caused by the high-volume change of the composite electrodes after 100 cycles at 100 mA g−1.
The results observed in Figures S3 and S4 on the electrochemical performance of TiO2/Sb2O3 composite electrodes with lower ratios (1:4) showed much higher amounts of the Sb material and a not-sufficient amount of TiO2 to observe a coating on Sb2O3 particles. The diffraction patterns of the three ratios studied in the hydrothermal synthesis are shown in the supplementary information (Figure S5). For the current work, it was desirable to perform a high level of TiO2 coating on the Sb2O3 particles to ensure that the volume expansion issue of the composite electrode, caused by a high amount of Sb, would be adequately addressed. The XRD results for the 1:1 ratio (Figure S5A) show some visible TiO2 coating in the diffraction pattern, as indicated by the peak at 29 in 2-theta, while for the (1:4) TiO2:Sb2O3 sample the TiO2 was not visible. To study the effect of TiO2 on the Sb2O3 material, one needs to have sufficient TiO2 present in the sample, as without enough amount of TiO2 in the composite there may only be a doping effect and not a complete encapsulation of the Sb2O3 particles. The diffraction pattern of the (1:4) TiO2:Sb2O3 sample indicates that the TiO2 phase was not visible. However, in the (1:1) TiO2:Sb2O3 sample, the Ti phase becomes visible in the diffraction pattern, as does the Sb5O4Cl2. In fact, Sb4O5Cl2 compounds were used as anode materials in LIBs, SIBs and potassium ion batteries [72]. The electrochemical performance of the Sb4O5Cl2 anode in LIBs was somehow similar to that of Sb2O3 [31].
To examine the rate performance of the nanocomposite electrode, Li-ion half cells were cycled at different current densities, starting at 50, 100, 200, 400, 500 and then back to 50 mA g−1. Figure 11 shows the rate performance in terms of charge capacity vs cycle numbers for TiO2, Sb2O3 and (3:1) TiO2/Sb2O3 electrodes. The Sb2O3 electrode exhibited a high capacity up to the 40th cycle, but was not steady and decreased rapidly over time, while TiO2 demonstrated a stable rate performance but showed low capacity. With the addition of TiO2 to Sb2O3, the rate performance was improved—by maintaining a high capacity over 300 mAh g−1—and there were excellent signs of cycling stability of the composite electrode. Particularly, when cycled back at a current density of 50 mA g−1, the (3:1) TiO2/Sb2O3 electrode recovered its capacity, demonstrating great reversibility. The nanocomposite electrode delivered a capacity of 354.8 mAh g−1 at the 10th cycle and at 50 mA g−1, while at the 60th cycle the capacity was 336.08 mAh g−1.
The electrochemical performance results of the proposed (3:1) TiO2/Sb2O3 composite electrode are similar to those reported by Han et al. [37,50] and Wang et al. [49] on similar composite electrodes in albeit different systems. However, the synthesis method and the systems reported in this work are different from those reported by Han et al. and Wang et al. Moreover, the morphology and structure of the TiO2/Sb2O3 composite materials discussed in the present work are different from those reported by Wang et al. and Han et al. [37,49,50].

4. Conclusions

In summary, TiO2/Sb2O3 nanocomposites and TiO2@Sb2O3 composites with different (TiO2:: Sb2O3) ratios of 0:1, 1:1, 1:4 and 3:1 were synthesized by the hydrolysis of TiCl4 and subsequent hydrothermal treatment at 175 °C and then used as anodes for LIBs. The (3:1) TiO2/Sb2O3 nanocomposite electrode outperformed all the electrodes synthesized in this work by delivering excellent cycling stability and rate capability. The addition of the synthesized TiO2 to Sb2O3 resulted in reduced volume change of the TiO2/Sb2O3 composite electrode and hence improved its electrochemical performance and rate capability. The (3:1) TiO2/Sb2O3 composite electrode delivered a capacity of 536 mAh g−1 after 100 cycles at 100 mA g−1 and exhibited great rate performance at different current densities. Halfway through its cycle-life, the (3:1) TiO2/Sb2O3 anode showed an overall loss of 2.9% compared to its final cycle. It was distinguished that as the ratio of TiO2 was increased in the nanocomposite system, the electrochemical performance of the TiO2/Sb2O3 electrode became more stable while still inhibiting antimony’s high capacity. This improved performance is attributed to the dispersion of TiO2, which acts as a buffer to mitigate the volume change of the electrode material while maintaining the integrity of its charge/discharge cycles. The XRD and XPS data showed that the TiO2 was in its respective rutile phase and resembled what is observed in the literature. Furthermore, the EDS mappings showed that the chemical makeup of the synthesized anode material was near its intended ratio of 3:1. The synthesized TiO2/Sb2O3 nanocomposite can be considered as a promising candidate for electrode materials of the next generation of rechargeable lithium-ion batteries due to its excellent capacity retention and cyclability, all through a simple approach.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app14156598/s1, Figure S1: Cycle voltammetry of TiO2/Sb2O3 composite with 1:1 ratio; Figure S2: Cycle voltammetry of TiO2/Sb2O3 composite with 1:4 ratio. Figure S3: Charge/discharge curves after100 cycles at 100 mAg−1 and cycling performance (B) of (1:1) TiO2/Sb2O3 composite electrode. Figure S4: Charge/discharge curves after 100 cycles at 100 mAg−1 and cycling performance (B) of (1:4) TiO2/Sb2O3 composite electrode. Figure S5: XRD patterns collected for the (A) TiO2/Sb2O3 1:4, (B) TiO2/Sb2O3 1:1 and (C) TiO2/Sb2O3 3:1. Figure S6: EDS spectrum for Sb2O3 (A), TiO2 (B) and (3:1) TiO2/Sb2O3 (C).

Author Contributions

Conceptualization, M.A.; Methodology, K.G., J.G.P. and M.A.; Formal Analysis, K.G., J.G.P., E.F. and M.A.; Investigation, K.G. and M.A.; Resources, J.G.P. and M.A.; Data Curation, J.G.P. and M.A.; Writing—Original Draft Preparation, K.G.; Writing—Review & Editing, J.G.P. and M.A.; Supervision, J.G.P. and M.A.; Project Administration, M.A.; Funding Acquisition, J.G.P. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

M. Alcoutlabi acknowledges funding from NSF PREM (DMR-2122178), Partnership for Fostering Innovation by Bridging Excellence in Research and Student Success. M. Alcoutlabi also acknowledges the funding from the Lloyd M. Bentsen, Jr. Endowed Chair in Engineering endowment at UTRGV. J.G. Parsons acknowledges the support provided by the UTRGV Chemistry Departmental Welch Foundation Grant (Grant No. BX-0048).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The authors acknowledge the support from the NSF PREM program under Award DMR-2122178: UTRGV-UMN Partnership for Fostering Innovation by Bridging Excellence in Research and Student Success. This work was partially supported by the Lloyd M. Bentsen, Jr. Endowed Chair in Engineering endowment at UTRGV. J.G. Parsons acknowledges the support provided by the UTRGV Chemistry Departmental Welch Foundation Grant (Grant No. BX-0048).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Asenbauer, J.; Eisenmann, T.; Kuenzel, M.; Kazzazi, A.; Chen, Z.; Bresser, D. The success story of graphite as a lithium-ion anode material—fundamentals, remaining challenges, and recent developments including silicon (oxide) composites. Sustain. Energy Fuels 2020, 4, 5387–5416. [Google Scholar] [CrossRef]
  2. Zhang, H.; Yang, Y.; Ren, D.S.; Wang, L.; He, X.M. Graphite as anode materials: Fundamental mechanism, recent progress and advances. Energy Storage Mater. 2021, 36, 147–170. [Google Scholar] [CrossRef]
  3. Aurbach, D.; Zinigrad, E.; Cohen, Y.; Teller, H. A short review of failure mechanisms of lithium metal and lithiated graphite anodes in liquid electrolyte solutions. Solid State Ion. 2002, 148, 405–416. [Google Scholar] [CrossRef]
  4. Agubra, V.A.; Zuniga, L.; Flores, D.; Campos, H.; Villarreal, J.; Alcoutlabi, M. A comparative study on the performance of binary SnO2/NiO/C and Sn/C composite nanofibers as alternative anode materials for lithium ion batteries. Electrochim. Acta 2017, 224, 608–621. [Google Scholar] [CrossRef]
  5. Ji, L.W.; Lin, Z.; Alcoutlabi, M.; Zhang, X.W. Recent developments in nanostructured anode materials for rechargeable lithium-ion batteries. Energy Env. Sci. 2011, 4, 2682–2699. [Google Scholar] [CrossRef]
  6. Yang, Z.G.; Zhang, J.L.; Kintner-Meyer, M.C.W.; Lu, X.C.; Choi, D.W.; Lemmon, J.P.; Liu, J. Electrochemical Energy Storage for Green Grid. Chem. Rev. 2011, 111, 3577–3613. [Google Scholar] [CrossRef]
  7. Xu, K. Electrolytes and Interphases in Li-Ion Batteries and Beyond. Chem. Rev. 2014, 114, 11503–11618. [Google Scholar] [CrossRef] [PubMed]
  8. Guo, H.; Zhao, H.; Yin, C.; Qiu, W. Si/SnSb alloy composite as high capacity anode materials for Li-ion batteries. J. Alloys Compd. 2006, 426, 277–280. [Google Scholar] [CrossRef]
  9. Nithyadharseni, P.; Reddy, M.; Nalini, B.; Ravindran, T.; Pillai, B.; Kalpana, M.; Chowdari, B. Electrochemical studies of CNT/Si–SnSb nanoparticles for lithium ion batteries. Mater. Res. Bull. 2015, 70, 478–485. [Google Scholar] [CrossRef]
  10. Park, M.-S.; Rajendran, S.; Kang, Y.-M.; Han, K.-S.; Han, Y.-S.; Lee, J.-Y. Si–Ni alloy–graphite composite synthesized by arc-melting and high-energy mechanical milling for use as an anode in lithium-ion batteries. J. Power Sources 2006, 158, 650–653. [Google Scholar] [CrossRef]
  11. Fang, Y.J.; Guan, B.Y.; Luan, D.Y.; Lou, X.W. Synthesis of CuS@CoS Double-Shelled Nanoboxes with Enhanced Sodium Storage Properties. Angew. Chem. Int. Ed. 2019, 58, 7739–7743. [Google Scholar] [CrossRef]
  12. Fang, Y.J.; Luan, D.Y.; Chen, Y.; Gao, S.Y.; Lou, X.W. Synthesis of Copper-Substituted CoS@CuS Double-Shelled Nanoboxes by Sequential Ion Exchange for Efficient Sodium Storage. Angew. Chem. Int. Ed. 2020, 59, 2644–2648. [Google Scholar] [CrossRef] [PubMed]
  13. Fang, Y.J.; Luan, D.Y.; Lou, X.W. Recent Advances on Mixed Metal Sulfides for Advanced Sodium-Ion Batteries. Adv. Mater. 2020, 32, 2002976. [Google Scholar] [CrossRef]
  14. Mishra, A.K.; Monika; Patial, B.S. A review on recent advances in anode materials in lithium ion batteries. Mater. Today Electron. 2024, 7, 100089. [Google Scholar] [CrossRef]
  15. Agubra, V.A.; Zuniga, L.; Flores, D.; Villareal, J.; Alcoutlabi, M. Composite Nanofibers as Advanced Materials for Li-ion, Li-O and Li-S Batteries. Electrochim. Acta 2016, 192, 529–550. [Google Scholar] [CrossRef]
  16. Agubra, V.A.; Zuniga, L.; De la Garza, D.; Gallegos, L.; Pokhrel, M.; Alcoutlabi, M. Forcespinning: A new method for the mass production of Sn/C composite nanofiber anodes for lithium ion batteries. Solid State Ion. 2016, 286, 72–82. [Google Scholar] [CrossRef]
  17. Zuniga, L.; Agubra, V.; Flores, D.; Campos, H.; Villareal, J.; Alcoutlabi, M. Multichannel hollow structure for improved electrochemical performance of TiO/Carbon composite nanofibers as anodes for lithium ion batteries. J. Alloys Compd. 2016, 686, 733–743. [Google Scholar] [CrossRef]
  18. Murugesan, S.; Harris, J.T.; Korgel, B.A.; Stevenson, K.J. Copper-coated amorphous silicon particles as an anode material for lithium-ion batteries. Chem. Mater. 2012, 24, 1306–1315. [Google Scholar] [CrossRef]
  19. Xue, L.; Xu, G.; Li, Y.; Li, S.; Fu, K.; Shi, Q.; Zhang, X. Carbon-coated Si nanoparticles dispersed in carbon nanotube networks as anode material for lithium-ion batteries. ACS Appl. Mater. Interfaces 2013, 5, 21–25. [Google Scholar] [CrossRef]
  20. Li, X.F.; Dhanabalan, A.; Gu, L.; Wang, C.L. Three-Dimensional Porous Core-Shell Sn@Carbon Composite Anodes for High-Performance Lithium-Ion Battery Applications. Adv. Energy Mater. 2012, 2, 238–244. [Google Scholar] [CrossRef]
  21. He, J.; Wei, Y.Q.; Zhai, T.Y.; Li, H.Q. Antimony-based materials as promising anodes for rechargeable lithium-ion and sodium-ion batteries. Mater. Chem. Front. 2018, 2, 437–455. [Google Scholar] [CrossRef]
  22. Tian, J.J.; Yang, H.; Fu, C.M.; Sun, M.L.; Wang, L.N.; Liu, T.X. In-situ synthesis of microspherical Sb@C composite anode with high tap density for lithium/sodium-ion batteries. Compos. Commun. 2020, 17, 177–181. [Google Scholar] [CrossRef]
  23. Moolayadukkam, S.; Bopaiah, K.A.; Parakkandy, P.K.; Thomas, S. Antimony (Sb)-Based Anodes for Lithium-Ion Batteries: Recent Advances. Condens. Matter 2022, 7, 27. [Google Scholar] [CrossRef]
  24. Hou, H.S.; Cao, X.Y.; Yang, Y.C.; Fang, L.B.; Pan, C.C.; Yang, X.M.; Song, W.X.; Ji, X.B. NiSb alloy hollow nanospheres as anode materials for rechargeable lithium ion batteries. Chem. Commun. 2014, 50, 8201–8203. [Google Scholar] [CrossRef] [PubMed]
  25. Ke, F.S.; Huang, L.; Solomon, B.C.; Wei, G.Z.; Xue, L.J.; Zhang, B.; Li, J.T.; Zhou, X.D.; Sun, S.G. Three-dimensional nanoarchitecture of Sn-Sb-Co alloy as an anode of lithium-ion batteries with excellent lithium storage performance. J. Mater. Chem. 2012, 22, 17511–17517. [Google Scholar] [CrossRef]
  26. Liu, J.; Yu, L.T.; Wu, C.; Wen, Y.R.; Yin, K.B.; Chiang, F.K.; Hu, R.Z.; Liu, J.W.; Sun, L.T.; Gu, L.; et al. New Nanoconfined Galvanic Replacement Synthesis of Hollow Sb@C Yolk-Shell Spheres Constituting a Stable Anode for High-Rate Li/Na-Ion Batteries. Nano Lett. 2017, 17, 2034–2042. [Google Scholar] [CrossRef] [PubMed]
  27. Bryngelsson, H.; Eskhult, J.; Nyholm, L.; Herranen, M.; Alm, O.; Edström, K. Electrodeposited Sb and Sb/SbO nanoparticle coatings as anode materials for Li-ion batteries. Chem. Mater. 2007, 19, 1170–1180. [Google Scholar] [CrossRef]
  28. 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]
  29. Liang, S.; Cheng, Y.J.; Zhu, J.; Xia, Y.; Müller-Buschbaum, P. A chronicle review of nonsilicon (Sn, Sb, Ge)-based lithium/sodium-ion battery alloying anodes. Small Methods 2020, 4, 2000218. [Google Scholar] [CrossRef]
  30. Reddy, M.V.; Subba Rao, G.; Chowdari, B. Metal oxides and oxysalts as anode materials for Li ion batteries. Chem. Rev. 2013, 113, 5364–5457. [Google Scholar] [CrossRef]
  31. Deng, M.; Li, S.; Hong, W.; Jiang, Y.; Xu, W.; Shuai, H.; Zou, G.; Hu, Y.; Hou, H.; Wang, W. Octahedral Sb2O3 as high-performance anode for lithium and sodium storage. Mater. Chem. Phys. 2019, 223, 46–52. [Google Scholar] [CrossRef]
  32. Chen, X.Y.; Yao, T.H.; Dong, H.; Ge, Q.J.; Chen, S.Q.; Ma, Z.H. Ultrafine SbO Nanoparticle-Decorated Reduced Graphene Oxide as an Anode Material for Lithium-Ion Batteries. Energy Fuel 2023, 37, 5586–5594. [Google Scholar] [CrossRef]
  33. Tan, Y.; Chen, L.; Chen, H.; Hou, Q.; Chen, X. Synthesis of a symmetric bundle-shaped Sb2O3 and its application for anode materials in lithium ion batteries. Mater. Lett. 2018, 212, 103–106. [Google Scholar] [CrossRef]
  34. Liu, Y.; Wang, H.; Yang, K.; Yang, Y.; Ma, J.; Pan, K.; Wang, G.; Ren, F.; Pang, H. Enhanced electrochemical performance of Sb2O3 as an anode for lithium-ion batteries by a stable cross-linked binder. Appl. Sci. 2019, 9, 2677. [Google Scholar] [CrossRef]
  35. Armstrong, A.R.; Armstrong, G.; Canales, J.; García, R.; Bruce, P.G. Lithium-ion intercalation into TiO2-B nanowires. Adv. Mater. 2005, 17, 862–865. [Google Scholar] [CrossRef]
  36. Batteries, L. New Materials, Developments and Perspectives. Ind. Chem. Libr. 1994, 5, 137–165. [Google Scholar]
  37. Han, Q.G.; Sun, Y.B.; Zhang, W.Q.; Li, X.; Li, Y.; Zhang, X.; Deng, Y.S. Synthesis of one-dimensional yolk-shell SbO/TiO composite as an anode material for enhanced lithium-storage properties. Ionics 2020, 26, 1221–1228. [Google Scholar] [CrossRef]
  38. Huang, M.; Chu, Y.T.; Xi, B.J.; Shi, N.X.; Duan, B.; Zhang, C.H.; Chen, W.H.; Feng, J.K.; Xiong, S.L. TiO-Based Heterostructures with Different Mechanism: A General Synergistic Effect toward High-Performance Sodium Storage. Small 2020, 16, 2004054. [Google Scholar] [CrossRef]
  39. Wang, S.Y.; Wu, N.N.; Liu, Y.Z.; Liu, W.; Liu, J.R. Improved electrochemical performances derived from synergistic titanium dioxide and iron titanate porous nanohybrids serving as lithium-ion battery anodes. J. Alloys Compd. 2017, 714, 583–592. [Google Scholar] [CrossRef]
  40. Wang, J.; Ran, R.; Tade, M.O.; Shao, Z.P. Self-assembled mesoporous TiO/carbon nanotube composite with a three-dimensional conducting nanonetwork as a high-rate anode material for lithium-ion battery. J. Power Sources 2014, 254, 18–28. [Google Scholar] [CrossRef]
  41. Wang, L.; Yuan, Y.F.; Zhang, T.; Chen, Q.; Guo, S.Y. Co O hollow nanospheres/carbon-assembled mesoporous polyhedron with internal bubbles encapsulating TiO nanosphere for high-performance lithium ion batteries. Nanotechnology 2019, 30, 355401. [Google Scholar] [CrossRef] [PubMed]
  42. Paul, S.; Rahman, M.A.; Sharif, S.B.; Kim, J.-H.; Siddiqui, S.-E.-T.; Hossain, M.A.M. TiO2 as an Anode of high-performance lithium-ion batteries: A Comprehensive Review towards Practical Application. Nanomaterials 2022, 12, 2034. [Google Scholar] [CrossRef] [PubMed]
  43. Pfanzelt, M.; Kubiak, P.; Fleischhammer, M.; Wohlfahrt-Mehrens, M. TiO2 rutile—An alternative anode material for safe lithium-ion batteries. J. Power Sources 2011, 196, 6815–6821. [Google Scholar] [CrossRef]
  44. Laskova, B.; Zukalova, M.; Zukal, A.; Bousa, M.; Kavan, L. Capacitive contribution to Li-storage in TiO2 (B) and TiO2 (anatase). J. Power Sources 2014, 246, 103–109. [Google Scholar] [CrossRef]
  45. Yang, L.R.; Niu, Z.R.; Wang, C.M.; Liu, Z.G.; Feng, X.X. Preparation and electrochemical performance of biomass-derived porous carbon/MoO2@TiO2 composite as anode materials for lithium-ion batteries. Solid State Ion. 2023, 389, 116110. [Google Scholar] [CrossRef]
  46. Gonzalez, G.; Hasan, M.T.; Ramirez, D.; Parsons, J.; Alcoutlabi, M. Synthesis of SnO/TiO micro belt fibers from polymer composite precursors and their applications in Li-ion batteries. Polym. Eng. Sci. 2022, 62, 360–372. [Google Scholar] [CrossRef]
  47. Rhee, D.Y.; Kim, J.; Moon, J.; Park, M.S. Off-stoichiometric TiO-decorated graphite anode for high-power lithium-ion batteries. J. Alloys Compd. 2020, 843, 156042. [Google Scholar] [CrossRef]
  48. Leal, J.H.; Cantu, Y.; Gonzalez, D.F.; Parsons, J.G. Brookite and anatase nanomaterial polymorphs of TiO synthesized from TiCl. Inorg. Chem. Commun. 2017, 84, 28–32. [Google Scholar] [CrossRef]
  49. Wang, Z.M.; Cheng, Y.; Li, Q.; Chang, L.M.; Wang, L.M. Facile synthesis of one-dimensional hollow SbO@TiO composites as anode materials for lithium ion batteries. J. Power Sources 2018, 389, 214–221. [Google Scholar] [CrossRef]
  50. Han, Q.G.; Sheng, Y.L.; Han, Z.W.; Li, X.; Zhang, W.Q.; Li, Y.; Zhang, X. Metallic Sb nanoparticles embedded into a yolk-shell SbO@TiO composite as anode materials for lithium ion batteries. New J. Chem. 2020, 44, 13430–13438. [Google Scholar] [CrossRef]
  51. Rodriguezcarvajal, J. Recent Advances in Magnetic-Structure Determination by Neutron Powder Diffraction. Physics B 1993, 192, 55–69. [Google Scholar] [CrossRef]
  52. Meagher, E.; Lager, G.A. Polyhedral thermal expansion in the TiO2 polymorphs; refinement of the crystal structures of rutile and brookite at high temperature. Can. Mineral. 1979, 17, 77–85. [Google Scholar]
  53. Sarnstrand, C. Crystal-Structure of Antimony(Iii) Chloride Oxide Sb4O5Cl2. Acta Crystallogr. B 1978, 34, 2402–2407. [Google Scholar] [CrossRef]
  54. Svensson, C. Refinement of Crystal-Structure of Cubic Antimony Trioxide, Sb2O3. Acta Crystallogr. B 1975, 31, 2016–2018. [Google Scholar] [CrossRef]
  55. Svensson, C. Crystal-Structure of Orthorhombic Antimony Trioxide, Sb2o3. Acta Crystallogr. B 1974, 30, 458–461. [Google Scholar] [CrossRef]
  56. Diebold, U.; Madey, T. TiO2 by XPS. Surf. Sci. Spectra 1996, 4, 227–231. [Google Scholar] [CrossRef]
  57. Song, D.; Hrbek, J.; Osgood, R. Formation of TiO nanoparticles by reactive-layer-assisted deposition and characterization by XPS and STM. Nano Lett. 2005, 5, 1327–1332. [Google Scholar] [CrossRef]
  58. Lee, M.K.; Park, Y.C. Contact Angle Relaxation and Long-Lasting Hydrophilicity of Sputtered Anatase TiO Thin Films by Novel Quantitative XPS Analysis. Langmuir 2019, 35, 2066–2077. [Google Scholar] [CrossRef]
  59. Matouk, Z.; Islam, M.; Gutiérrez, M.; Pireaux, J.J.; Achour, A. X-ray Photoelectron Spectroscopy (XPS) Analysis of Ultrafine Au Nanoparticles Supported over Reactively Sputtered TiO Films. Nanomaterials. 2022, 12, 3692. [Google Scholar] [CrossRef]
  60. Zeng, D.W.; Xie, C.S.; Zhu, B.L.; Song, W.L. Characteristics of SbO nanoparticles synthesized from antimony by vapor condensation method. Mater. Lett. 2004, 58, 312–315. [Google Scholar] [CrossRef]
  61. Nguyen, Q.H.; Salunkhe, T.T.; Nguyen, T.A.; Nguyen, T.L.; Kim, I. Optimal tailored preparation of Sb/Sb4O5Cl2 nanosheet composite anodes for efficient sodium-ion storage. Electrochim. Acta 2022, 436, 141429. [Google Scholar] [CrossRef]
  62. Chen, J.J.; Zhao, S.N.; Meng, W.J.; Guo, M.Q.; Wang, G.W.; Guo, C.L.; Bai, Z.C.; Li, Z.Q.; Ye, J.Y.; Song, H.; et al. Electrochemical Performance and Stress Distribution of Sb/SbO Nanoparticles as Anode Materials for Sodium-Ion Batteries. Batteries 2023, 9, 98. [Google Scholar] [CrossRef]
  63. Liu, G.R.; Zhang, S.C.; Wu, X.M.; Lin, R.X. Fabrication of rutile TiO nanorod arrays on a copper substrate for high-performance lithiumion batteries. Rsc. Adv. 2016, 6, 55671–55675. [Google Scholar] [CrossRef]
  64. Gan, Y.P.; Zhu, L.Y.; Qin, H.P.; Xia, Y.; Xiao, H.; Xu, L.S.; Ruan, L.Y.; Liang, C.; Tao, X.Y.; Huang, H.; et al. Hybrid nanoarchitecture of rutile TiO nanoneedle/graphene for advanced lithium-ion batteries. Solid State Ion. 2015, 269, 44–50. [Google Scholar] [CrossRef]
  65. McNulty, D.; Carroll, E.; O’Dwyer, C. Rutile TiO Inverse Opal Anodes for Li-Ion Batteries with Long Cycle Life, High-Rate Capability, and High Structural Stability. Adv. Energy Mater. 2017, 7, 1602291. [Google Scholar] [CrossRef]
  66. Han, Z.; Zheng, J.; Kong, F.; Tao, S.; Qian, B. One dimensional SbO2/Sb2O3@ NC microrod as anode for lithium-ion and sodium-ion batteries. Nano Sel. 2021, 2, 425–432. [Google Scholar] [CrossRef]
  67. Cheng, Y.; Yi, Z.; Wang, C.; Wang, L.; Wu, Y.; Wang, L. Nanostructured carbon/antimony composites as anode materials for lithium-ion batteries with long life. Chem. Asian J. 2016, 11, 2173–2180. [Google Scholar] [CrossRef] [PubMed]
  68. Zhou, X.Z.; Zhang, Z.F.; Lu, X.F.; Lv, X.Y.; Ma, G.F.; Wang, Q.T.; Lei, Z.Q. SbO Nanoparticles Anchored on Graphene Sheets via Alcohol Dissolution-Reprecipitation Method for Excellent Lithium-Storage Properties. Acs Appl. Mater. Inter. 2017, 9, 34927–34936. [Google Scholar] [CrossRef]
  69. Lv, H.; Qiu, S.; Lu, G.; Fu, Y.; Li, X.; Hu, C.; Liu, J. Nanostructured antimony/carbon composite fibers as anode material for lithium-ion battery. Electrochim. Acta 2015, 151, 214–221. [Google Scholar] [CrossRef]
  70. Jeong, J.-H.; Jung, D.-W.; Shin, E.W.; Oh, E.-S. Boron-doped TiO2 anode materials for high-rate lithium ion batteries. J. Alloys Compd. 2014, 604, 226–232. [Google Scholar] [CrossRef]
  71. Zhang, C.Q.; Yang, F.; Zhang, D.L.; Zhang, X.; Xue, C.L.; Zuo, Y.H.; Li, C.B.; Cheng, B.W.; Wang, Q.M. A binder-free Si-based anode for Li-ion batteries. Rsc. Adv. 2015, 5, 15940–15943. [Google Scholar] [CrossRef]
  72. Shi, Y.Q.; Wang, L.; Zhou, D.; Wu, T.L.; Xiao, Z.B. A Flower-Like SbOCl Cluster-based material as anode for potassium ion batteries. Appl. Surf. Sci. 2022, 583, 152509. [Google Scholar] [CrossRef]
Figure 1. SEM images of (A,B) Sb2O3 (C,D) TiO2 and (E,F) (3:1) TiO2Sb2O3 material.
Figure 1. SEM images of (A,B) Sb2O3 (C,D) TiO2 and (E,F) (3:1) TiO2Sb2O3 material.
Applsci 14 06598 g001
Figure 2. EDS mapping of the Sb2O3 sample showing SEM image of the mapped area (A), O (B) and Sb (C).
Figure 2. EDS mapping of the Sb2O3 sample showing SEM image of the mapped area (A), O (B) and Sb (C).
Applsci 14 06598 g002
Figure 3. EDS mapping of the TiO2 sample showing SEM image of the mapped area (A), O (B) and Ti (C).
Figure 3. EDS mapping of the TiO2 sample showing SEM image of the mapped area (A), O (B) and Ti (C).
Applsci 14 06598 g003
Figure 4. EDS mapping of the TiO2@Sb2O3 3:1 ratio sample showing SEM image (A), O (B), Ti (C), Sb (D) and Cl (E).
Figure 4. EDS mapping of the TiO2@Sb2O3 3:1 ratio sample showing SEM image (A), O (B), Ti (C), Sb (D) and Cl (E).
Applsci 14 06598 g004
Figure 5. Le Bail results for the fitting of the XRD patterns collected for the (A) (3:1) TiO2@Sb2O3, (B) TiO2 and (C) Sb2O3.
Figure 5. Le Bail results for the fitting of the XRD patterns collected for the (A) (3:1) TiO2@Sb2O3, (B) TiO2 and (C) Sb2O3.
Applsci 14 06598 g005
Figure 6. Ti 2P XPS spectrum for hydrothermally synthesized TiO2 nanoparticles at 175 °C (A) and the (3:1) TiO2@Sb2O3 (B). The bleu, green and cyan lines are the fitting curves of the XPS results of TiO2 and (3:1) TiO2@Sb2O3.
Figure 6. Ti 2P XPS spectrum for hydrothermally synthesized TiO2 nanoparticles at 175 °C (A) and the (3:1) TiO2@Sb2O3 (B). The bleu, green and cyan lines are the fitting curves of the XPS results of TiO2 and (3:1) TiO2@Sb2O3.
Applsci 14 06598 g006
Figure 7. XPS spectrum of O 1S of the TiO2 synthesized hydrothermally at 175 °C (A) and the Sb3d for 3:1 TiO2 /Sb2O3 (B). The bleu, green and red lines are the fitting curves of the XPS results shown in the Figure.
Figure 7. XPS spectrum of O 1S of the TiO2 synthesized hydrothermally at 175 °C (A) and the Sb3d for 3:1 TiO2 /Sb2O3 (B). The bleu, green and red lines are the fitting curves of the XPS results shown in the Figure.
Applsci 14 06598 g007
Figure 8. Cyclic voltammetry of (A) TiO2, (B) Sb2O3 and (C) TiO2/Sb2O3 nanocomposite with a 3:1 ratio.
Figure 8. Cyclic voltammetry of (A) TiO2, (B) Sb2O3 and (C) TiO2/Sb2O3 nanocomposite with a 3:1 ratio.
Applsci 14 06598 g008
Figure 9. Charge/discharge curves after 100 cycles of (a) Sb2O3, (b) TiO2 and (c) (3:1) TiO2/Sb2O3.
Figure 9. Charge/discharge curves after 100 cycles of (a) Sb2O3, (b) TiO2 and (c) (3:1) TiO2/Sb2O3.
Applsci 14 06598 g009
Figure 10. Cycling performance and coulombic efficiency of Sb2O3, TiO2 and (3:1) TiO2/Sb2O3 electrodes at a current density of 100 mA g−1.
Figure 10. Cycling performance and coulombic efficiency of Sb2O3, TiO2 and (3:1) TiO2/Sb2O3 electrodes at a current density of 100 mA g−1.
Applsci 14 06598 g010
Figure 11. Rate performance of (3:1) TiO2/Sb2O3, TiO2 and Sb2O3 electrodes at different current densities.
Figure 11. Rate performance of (3:1) TiO2/Sb2O3, TiO2 and Sb2O3 electrodes at different current densities.
Applsci 14 06598 g011
Table 1. Results of the LeBail fitting procedure for the TiO2 synthesized at 175 °C, the Sb2O3 starting material and the TiO2@Sb2O3 composites synthesized at different ratios.
Table 1. Results of the LeBail fitting procedure for the TiO2 synthesized at 175 °C, the Sb2O3 starting material and the TiO2@Sb2O3 composites synthesized at different ratios.
SamplePhaseSpace Groupa (Å)b (Å)c (Å)α (°)β (°)γ (°)χ2
TiO2 RutileP42/mnm4.59 (8)4.59 (8)2.95 (9)90.090.090.02.41
Sb2O3CubicFm-3m11.15 (3)11.15 (3)11.15 (3)90.090.090.03.15
OrthorhombicPccn4.91 (0)12.46 (5)5.40 (5)90.090.090.0
TiO2@Sb2O3 3:1Sb5O4Cl2P21/C6.21 (9)5.11 (6)13.49 (9)90.097.390.01.28
TiO2 P42/mnm4.59 (6)4.59 (6)2.95 (8)90.090.090.0
Table 2. Summary of the XPS fittings of the synthesized TiO2 nanoparticles at 175 °C (A) and the TiO2@Sb2O3 3:1 (B).
Table 2. Summary of the XPS fittings of the synthesized TiO2 nanoparticles at 175 °C (A) and the TiO2@Sb2O3 3:1 (B).
SampleEnergy
(eV)
Ti
2p3/2/2p½
Energy
(eV)
TiEnergy
(eV)
O
1 s
Energy
(eV)
Sb
3d5/2/3d3/2
(A) TiO2458.6/464.3Ti4+-O471.9Satellite530.4
529.8
O-Ti
O-H
N/A
(B) TiO2@Sb2O3 3:1458.6/464.4Ti4+-O471.9Satellite531.4O-M528.4/537.7
530.7/540.4
Sb(0)
Sb3+-O
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gomez, K.; Fletes, E.; Parsons, J.G.; Alcoutlabi, M. The Effect of TiO2 on the Electrochemical Performance of Sb2O3 Anodes for Li-Ion Batteries. Appl. Sci. 2024, 14, 6598. https://doi.org/10.3390/app14156598

AMA Style

Gomez K, Fletes E, Parsons JG, Alcoutlabi M. The Effect of TiO2 on the Electrochemical Performance of Sb2O3 Anodes for Li-Ion Batteries. Applied Sciences. 2024; 14(15):6598. https://doi.org/10.3390/app14156598

Chicago/Turabian Style

Gomez, Kithzia, Elizabeth Fletes, Jason G. Parsons, and Mataz Alcoutlabi. 2024. "The Effect of TiO2 on the Electrochemical Performance of Sb2O3 Anodes for Li-Ion Batteries" Applied Sciences 14, no. 15: 6598. https://doi.org/10.3390/app14156598

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