Next Article in Journal
Synthesis of 3D Porous Cu Nanostructures on Ag Thin Film Using Dynamic Hydrogen Bubble Template for Electrochemical Conversion of CO2 to Ethanol
Previous Article in Journal
Structural Construction of WO3 Nanorods as Anode Materials for Lithium-Ion Batteries to Improve Their Electrochemical Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Porous Nb2O5 Formed by Anodic Oxidation as the Sulfur Host for Enhanced Performance Lithium-Sulfur Batteries

School of Materials Science and Engineering, Hebei University of Technology, Tianjin 300130, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(4), 777; https://doi.org/10.3390/nano13040777
Submission received: 14 January 2023 / Revised: 14 February 2023 / Accepted: 15 February 2023 / Published: 20 February 2023

Abstract

:
Lithium-sulfur batteries (LSBs), with their high theoretical specific capacity and energy density, have great potential to be a candidate for secondary batteries in the future. However, Li-S batteries suffer from multiple issues and challenges, for example, uneven growth of lithium dendrites, low utilization of the active material (sulfur), and low specific capacity. This paper reports a low-cost and anodic oxidation method to produce niobium pentoxide with a porous structure (P-Nb2O5). A simple one-step process was used to synthesize P-Nb2O5 with porous structures by anodizing niobium at 40 V in fluorinated glycerol. The porous Nb2O5 showed excellent rate capability and good capacity retention by maintaining its structural integrity, allowing us to determine the advantages of its porous structure. As a result of the highly porous structure, the sulfur was not only provided with adequate storage space and abundant adsorption points, but it was also utilized more effectively. The initial discharge capacity with the P-Nb2O5 cathode rose to 1106.8 mAh·g−1 and dropped to 810.7 mAh·g−1 after 100 cycles, which demonstrated the good cycling performance of the battery. This work demonstrated that the P-Nb2O5 prepared by the oxidation method has strong adsorption properties and good chemical affinity.

1. Introduction

With the increasing use of electrical energy, LSBs have already been presented in all aspects of everyday life [1]. Compared with traditional insertion lithium-ion batteries, LSBs are promising energy storage systems for the next generation due to their high theoretical capacity (1675 mAh·g−1) and high energy density (2600 Wh·kg−1) [2,3]. However, the practical implementation of Li-S continues to face formidable challenges, such as the poor electrical conductivity of sulfur, the series of polysulfide intermediates (Li2Sn, 6 ≤ n ≤ 8) that can dissolve in electrolytes, and the intermediate polysulfides that diffuse to form shuttle effects, which cause pronounced capacity fading [4]. Further, during charging and discharging, the conversion between sulfur and Li2S/Li2S2 can damage the electrode structure and lead to a number of problems such as rapid capacity loss [5]. The performance and the practical application of lithium-sulfur batteries continues to be hindered by these essential problems. A large number of researchers have spent a significant amount of time researching and exploring lithium sulfur battery cathode materials to solve the above problems.
Qiao et al. developed a multistage porous Co-NC@Sn (NH3) structure that was constructed as a sulfur host material for Li-S batteries, and it was a simple method for improving Li-S electrochemistry [6]. Manthiram employed a series of materials, i.e., designing and synthesizing a composite using carbon and polar materials, to modify the anode material. For example, mesoporous carbon materials that are well-designed and porous carbon composite materials can provide a large surface area for restraining the polysulfides through physical adsorption [7]. Lu et al. provided three-dimensional (3D) porous graphene foams constructed of reduced graphene oxide (r-GO) for the sulfur cathodes of LSBs. Recently, materials with various pore structure have attracted great attention. The presence of a pore structure can increase the surface energy of a system, which contributes to the improvement of the abundant active sites and ensures the effective wetting and penetration of electrolytes to the electrode surface, facilitating the transfer of interfacial charges [8,9]. A pore structure also provides rapid Li+ ion transport highways, resulting in significant rate capabilities [10].
Meanwhile, researchers have spent a significant amount of time and effort studying the use of polar metal oxides in LSBs, and polar metal oxides with strong chemical adsorption, such as TiO2, MnO2, V2O5, and Nb2O5, were investigated to alleviate intermediate polysulfide diffusion (LIPSs, Li2Sn, n = 4–8). Zhao et al. developed a V2O3–VN@NC sulfur host that showed excellent S storage performances, and it had a high reversible capacity, good rate capability, and stable cyclic properties of the reaction process. Using density functional theory computations, they also verified that V2O3 played a key role in the adsorption of active S during the polysulfide reaction [11]. Song studied a novel reduced graphene oxide (rGO) wrapped a yolk–shell vanadium dioxide (VO2) sphere hybrid host (rGO/VO2) and determined that the polar VO2 could effectively trap polysulfide products by means of strong chemical bonding with oxygen bonds such that the polysulfides largely remained inside the anode electrode material. This designed rGO/VO2/S cathode delivered outstanding cycle stability with a low fading rate. Moreover, the porous nature of the conductive 3D porous rGO network facilitated the penetration of the electrolytes and the transportation of the ions, providing efficient transport routes for the lithium ions and electrons and buffering the volume expansion during the lithiation process. Further, this interconnected carbon network could promote chemical interactions with polysulfides. The chemical adsorption of the polysulfide species onto the carbon network hampered the shuttling effect [12].
NbOx-based materials (Nb2O5 and NbO2) and their composites have been studied and used for electrode materials for a variety of energy storage systems. They enable a fast recharging, high energy density, and long-term storage [13,14]. Zhan al. synthesized a novel urchin-like Nb2O5/CNT composite as a modified layer on a polypropylene separator. They prepared precursor materials using hydrothermal synthesis, and then they created the final materials using high-temperature sintering. The synthesized compound material not only provided physical trapping sites for polysulfides, but it also effectively slowed down the diffusion of the polysulfides due to its strong chemical interactions with the polysulfides [15]. Liu al. developed a Pt-Nb2O5 hybrid catalysis for use in lithium-sulfur battery cathode materials. The synthesized material facilitated the reduction in the higher order polysulfides and the capture of the polysulfides onto the Nb2O5 [16]. In order to obtain a stronger adsorption capacity of polar materials, the physical morphology factors of polar metal compounds can be studied. An ideal polar compound requires a light weight, a suitable surface for fast Li+ transportation, and an adsorptive capacity for S. The reaction equation of a lithium-sulfur battery is as follows: 16Li + S8 → 8Li2S. The refinement of the reaction process can be divided into the following different stages: (i) S8 → Li2S8, (ii) Li2S8 → Li2S6 → Li2S4, and (iii) Li2S4 → Li2S2 → Li2S. Preventing the loss of intermediates during a reaction is an effective means of inhibiting the formation of lithium dendrites. It was discovered that Nb2O5 has high ionic conductivities for Li+ ions at the surface, and polar Nb2O5 also has a strong redox ability for Li2S6. The redox kinetics of the materials efficiently ameliorated the long-chain shuttle effect of the polysulfides to the Li2S2/Li2S. At the same time, it was found that Nb2O5 has two forms: crystalline and amorphous. It was also found that the amorphous materials strongly adsorbed lithium polysulfides (Li2Sx, x < 6), and we found that although the amorphous materials showed that a remarkable increase in their application had been made possible by constant innovations in the technology of their preparation and their advantages in electrochemistry had increased significantly, the vast majority of catalytic materials that have been developed for sulfur cathodes are crystalline. Therefore, Nb2O5 material with an amorphous form can be prepared as the cathode mixing material for LSBs, taking into account the advantages of niobium oxide and its amorphous form to improve the overall performance of the battery. However, it was found that Nb2O5 transformed from an amorphous to a crystalline state after being calcinated at high temperatures. Therefore, a new porous amorphous Nb2O5 method of preparation was determined to be necessary. Porous amorphous Nb2O5 was identified as a potential material for lithium-sulfur batteries.
The electrochemical anodization method is a manufacturing process for the preparation of multistage porous oxide films on substrates [17]. The anodization of niobium has been widely studied at the basic research level [18]. In the present work, we introduced P-Nb2O5 with an amorphous form by means of anodic oxidation to obtain a porous structure [19]. We prepared the material in one step and did not use methods such as hydrothermal synthesis or high temperature calcination. The method used in the experiments was simple and safe, and at the same time, it allowed for the direct preparation of metal oxides with amorphous porous structures. The electrolyte diffusion and penetration were enhanced by the 3D interconnected mesoporous space, and electron transfer kinetics could also be improved by the structures, mitigating the cathode volume change through physical confinement. These porous materials mitigated the shuttle effect through physical trapping within their large surface areas, leading to the high-performance achieved by the LSBs [20].

2. Experimental Section

2.1. Synthesis of the P-Nb2O5 Samples

The metal niobium sheet (99.9%, Crown Thailand, Tianjin, China) with a thickness of 0.05 mm was cut to a size of 10 cm × 10 cm. Then, the niobium pieces were sonicated in ethanol solution at room temperature for 30 min and dried in an oven. The dried samples were anodized at 40 V for 2.5 h in 50 mL mixtures of deionized water and glycerol (1:20) containing 0.741 g NH4F. During electrochemical oxidation, a graphite electrode was attached to the anode of a Sorensen DLM 300-2 and a niobium metal electrode was attached to the cathode, and the output voltage of the power supply was 40 V. Potentials of 40 V were employed to stimulate oxide development, and anodization was carried out for 150 min. The prepared P-Nb2O5 was stripped from the matrix with an ultrasonic machine in ethanol, and the oxide layer was collected through centrifugation at 1000 RPM for 10 min. Finally, it was dried at 60 °C in an oven.

2.2. Characterization of the P-Nb2O5 Samples

The structures and shapes of the samples were examined using scanning electron microscopes (SEM, Sigma 500, Oberkochen, Germany), which were equipped with energy dispersive Xray spectroscopy (EDS) systems, and transmission electron microscopes (TEM, JEM-2011) with EDS. Characterization of the morphological and dimensional structures of the samples was completed using atomic force electron microscopy (AFM, Bruker Dimension ICON). The powder XRD patterns were identified by powder X-ray diffraction (Bruker D8 Advance X-ray diffractometer, Cu KR radiation) [21]. Thermal gravity analysis (TGA, Perkin Elmer, Series7, Waltham, MA, USA) was performed to test for sulfur content. A photoelectron spectrometer (ESCALab MKII) was used to perform the X-ray photoelectron spectroscopy (XPS) [22]. Raman spectroscopy was conducted to study the structural and molecular interactions (633 nm, Horiba Jobin Yvon, Stow, MA, USA).

2.3. Electrode Prepazration and Battery Fabrication

The batteries with the P-Nb2O5 mixed positive electrode for Li-S were tested with CR2032-type coin batteries, which were assembled in an Ar-filled glove box (<5 ppm of H2O and O2). The prepared P-Nb2O5 material and sublimated sulfur were ground and mixed well in a ratio of 1:3, and then the melt diffusion method was used to further mix them. The mixed material was placed in a reactor under heating conditions of 155 °C for 12 h, and after natural cooling, the final mixed S@P-Nb2O5 material was obtained. A conventional slurry-coating process was used to fabricate the P-Nb2O5 electrodes. Subsequently, we prepared N-methyl-2-pyrrolidone (NMP) slurries with 70 wt% active materials, carbon black (20 wt%), and a polyvinylidene fluoride (PVDF) (10 wt%) binder. The amount of sulfur mass in each electrode was approximately 1 mg·cm−2. The batteries were assembled in a glove box filled with argon, and the electrochemical tests were conducted on standard CR2032-type coin batteries. S@Nb2O5 was used as the cathode, Li foil was used as the anode, and 1 M lithium bis (trifluoromethane sulfonel) imide (LiTFSI, 2 wt% LiNO3, 1,3-dioxolane DOL, and 1,2-dimethoxyethane DME (v/v, 1:1)) solution was used as the electrolyte solution. We also tested the purchased granular Nb2O5 (G-Nb2O5, 99.9%, Aladdin, Shanghai, China) as a control for the experiment. An electrochemical evaluation was performed using an electrochemical workstation (CHI 660E, Shanghai Chenhua, Shanghai, China) in the potential range of 1.7–2.8 V. Cyclic voltammetry (CV, with a scan rate of 0.05 mV·s−1) was used to measure the potential range of 1.7–2.8 V and the impedance spectrum was measured over a frequency range of 100 kHz to 0.01 Hz. The sulfur load and electrolyte doses were 0.5–1.2 mg·cm−2 and 40–60 μL, respectively. Battery performance was measured using a Neware BTS4000 battery testing system (Shenzhen, China) at room temperature. In this study, our current density and specific capacity were calculated using the mass of elemental sulfur.

3. Results

The preparation of the porous mixed cathode material was realized using the anodic oxidation technique and a subsequent grinding process. The anodic oxidation technique had the advantage of being a more convenient operation method and oxidation process without high-voltage currents, which made it a with low-risk option. The porous structure of P-Nb2O5 offered a high immobilization ability towards the intermediate polysulfide in the anode material during the oxidation process in the inorganic environment. At the same time, the P-Nb2O5 microstructure exhibited the presence of abundant channels for rapid ion transport, and it had enough surface area for the reaction energy storage.
The morphology and microstructure of the P-Nb2O5 sheet were characterized by scanning electron microscopy (SEM) and an atomic force microscope (AFM). The SEM (Figure 1a–c) and AFM (Figure 1d–f) observations showed uniform nanopores of approximately 50 nm in size [23]. The magnified AFM image (Figure 1e) revealed that the aperture depth of P-Nb2O5 could reach 10–20 μm, and the nanopores were approximately 25–50 nm in diameter. The sulfur could be immobilized by these nanopores during the constant current charge/discharge cycle of the battery [24]. A comparison of the SEM (Figure S1a,b) and AFM (Figure S1c,d) of the unoxidized pure niobium flakes further demonstrated the porous structures formed by the anodic oxidation. As seen in Figure 1f, many pores were connected one another, resulting in the accumulation of abundant pores. The porous structures may have endowed the as-synthesized P-Nb2O5 material with a high surface area, and this also increased the surface area of the catalyst, which, in turn, could facilitate ion transfer and improve its rate performance. We used G-Nb2O5 in granular form (Figure S5a,b) the main control for comparison. Further, the energy dispersive spectrometer (EDS) mapping results demonstrated the uniform distribution of the elements (Nb and O) in the oxide (Figure 1g–i), indicating that the Nb2O5 was successfully prepared. P-Nb2O5 was mixed with sulfur and then morphologically characterized (Figure S2a–c), and the porous structure was homogeneously mixed with sulfur. Figure S2d–g shows that the synthesized substance had a homogeneous sulfur and P-Nb2O5 mixture. The overall structure of the material was found to be relatively stable, as indicated by its morphology after cycling (Figure S3a–c).
The greater surface area of the P-Nb2O5 indicated that there was increased contact area with the active substances. The high-resolution transmission electron microscopy (HRTEM) image showed the P-Nb2O5 microstructure and a state of void distribution (Figure 2a–c). Moreover, an abundance of mesopores was observed in the oxidized structure of the P-Nb2O5, which resulted in an increase in the active interfacial energy and the strength of the physical adsorption on the sulfur (Figure 2b). A deeper probe into the amorphous structure was carried out by HRTEM observation. As shown in Figure 2d–f, the energy dispersive spectroscopy (EDS) mapping further demonstrated that that Nb and O were evenly distributed in the oxide (Figure 2e,f), which also indicated the successful synthesis of P-Nb2O5.
Figure 3a shows the XRD pattern of the P-Nb2O5, and no diffraction peaks could be detected in the Nb2O5 pattern, indicating that the synthesized P-Nb2O5 was amorphous in nature [25]. It also shows that the G-Nb2O5 was in a crystalline state. The presence of elemental sulfur in the mixture could be determined after mixing the sulfur, as seen in the analysis of the XRD pattern shown in Figure S4. The textural properties of the P-Nb2O5 with a thickness of approximately 20 μm were characterized by the N2 physical adsorption/desorption experiments (Figure 3b,c) [26]. Figure 3b shows the N2 adsorption-desorption isotherm of the P-Nb2O5. The results corresponded to the TEM plots, indicating the presence of many gaps within the P-Nb2O5. BET analysis determined the specific surface area of the P-Nb2O5 to be 19.38 m2·g−1. Pore distributions in the BJH ranged from 2–10 nm, as shown in Figure 3c. Since the P-Nb2O5 had a porous structure, at the electrolyte–electrode interface, the high specific surface area increased the contact area, providing plentiful active sites for the sulfur interfacial reactions [27]. In addition, the Raman spectra shown in Figure 3d also showed an obvious single peak and peak width, which also indicated the formation of an amorphous substance. We conducted X-ray photoelectron spectroscopy (XPS), shown in Figure 3e,f, to the study of the composition of the amorphous P-Nb2O5. Through a peak at 206.4 eV, which was related to the Nb3d5/2, and a peak at 209.3 eV, which was related to the Nb3d3/2, the predominant chemical state in the P-Nb2O5 amorphous material was found to be Nb(V). Additionally, the O1s XPS results could be fitted to correspond to the two peaks. As seen in Figure 3e, the two elements detected in the Nb2O5 were O and Nb. The two peaks of the O1s centered at 529.3 and 531.1 eV originated from the Nb-O and Nb-OH, respectively (see Figure 3f).
Thermogravimetric analysis (TGA) toward the composite was carried out under an Ar atmosphere (Figure 4) to confirm the content of the sulfur. The weightlessness of the samples was divided into several different stages, depending on the temperature. The weight loss stage was seen at a low temperature range (under 100 °C) and corresponded to the slight mass loss at approximately 100 °C, which was attributed to the evaporation of water from the rising temperature (which was caused by the adsorption of the pores in the samples and could be neglected). It was then evident, as seen in the figure, that these samples showed weight loss from 180 °C to 350 °C in an inert gas environment. At this stage, the TGA diagrams showed a very rapid weight loss, mainly due to the rapid sublimation of sulfur in the composite, with a weight loss of ~74%. The weight percentage of the S in the composite was analyzed to be ~74 wt% using thermogravimetric analysis. The experimental results showed that the P-Nb2O5 had a high S loading capacity and sulfur-fixing effect.
We created the Li-S batteries using the P-Nb2O5 as the mixed material in the positive electrode, and we investigated the electrochemical performance of the prepared active material. The specific capacities calculated in this work were all based on sulfur weight [28]. Figure 5a shows the CV curves from the first to third cycles in a voltage range of 1.7~2.8 V at a scan rate of 0.1 mV s−1. During the first cathodic reduction step, two reduction peaks at 2.054 and 2.336 V were observed. During the electrochemical reduction, the soluble long-chain polysulfide form at 2.336 V was reduced to an insoluble short-chain polysulfide, forming another cathodic peak at 2.054 V and indicating that the sulfur underwent a multi-step reduction during the whole reaction process. During the subsequent anodic scan process, a strong oxidation peak could be observed at 2.345 V. During the ensuing electrochemical cycle reaction, the curve maintained a strong oxidation peak and small shoulders of 2.345 V and 2.385 V, respectively. This was caused by the transition of the long-chain polysulfides to short-chain polysulfides and the permeation of the electrolytes during the reaction. Further, the peak position and area of the CV curve remained constant from the first to the third cycles, indicating that the P-Nb2O5 material had excellent capacity retention [29]. Figure 5b displays the initial charge/discharge profiles of the P-Nb2O5 at 0.2 C in the voltage window of 1.7–2.8 V. It was clear that the P-Nb2O5 exhibited a higher capacity of 1106.8 mAh·g−1, indicating an increased sulfur utilization due to the excellent porous structure of the P-Nb2O5 electrode.
The cycling performances of the P-Nb2O5, G-Nb2O5, and CNT electrodes at 0.2 c are shown in Figure 5c. The initial discharge capacity of the P-Nb2O5 was 1106.8 mAh·g−1, indicating the good sulfur utilization of the electrode, with a capacity storage rate of 73.14% after 100 cycles. Compared to the granular Nb2O5 and CNT as the cathode mixing material, the battery that used the P-Nb2O5 as the mixing material for the positive electrodes showed excellent rate capability (Figure 5d,e) and delivered specific capacities of 1108.6, 1016.4, 892.6, 826.4, 753.4, and 713.4 mAh·g−1 at 0.2, 0.5, 1, 2, 3, and 5 C, respectively. The promoted rate capability of the P-Nb2O5 could be attributed to the porous nature, fast charge transfer, and rapid transport capabilities. When the rate returned to 0.2 C, the specific capacities of the batteries with the P-Nb2O5, G-Nb2O5, and CNT forms were 958.85, 655.01, and 870.70 mAh·g−1, respectively [30]. Electrochemical impedance spectroscopy on the batteries with the G-Nb2O5 and P-Nb2O5 electrodes was carried out, and the differences in the internal resistance were investigated, as shown in Figure 5f. The two Nyquist diagrams of the batteries in Figure 5f have the same shape, including a semicircle involving the charge transfer resistance (Rct) and a sloping line associated with the Li diffusion. The P-Nb2O5 had a lower charge transfer resistance (Rct = 117.6 Ω) compared to the G-Nb2O5, indicating that the P-Nb2O5 had the fastest charge transfer kinetics. To further investigate the status of the potential applications of the P-Nb2O5 cathode in practice, we explored the long cycle of the P-Nb2O5 and G-Nb2O5 cathodes, as seen in Figure 5g. The P-Nb2O5 cathode maintained a reversible capacity of up to 602 mAh·g−2 after 400 cycles at 1 C, with very high stability and discharge capacity compared to the G-Nb2O5 and CNT, which was primarily due to the redox reaction facilitated by the presence of the porous Nb2O5 and the suppression of the shuttle effect of the LiPSs. This demonstrated that the porous Nb2O5 could improve the reaction kinetics during the polysulfide reactions. In order to demonstrate the potential of the P-Nb2O5 for applications in LSBs, cycle performance tests were carried out on the LSBs with high sulfur loadings. As shown in Figure 5h, the electrode exhibited a high area capacity of 4.8 mAh·cm−2 at a high sulfur load of 5.2 mg·cm−2.
Adsorption experiments were performed on the polysulfides to verify the strength of the P-Nb2O5 on the polysulfides. The adsorption capacity of the P-Nb2O5 on the polysulfide was assessed by the observation of the color of the sample in a photograph (Figure 6a, inset shown). After 12 h, the solution in the bottle containing the P-Nb2O5 changed to near-transparent, indicating the good adsorption of the material to the polysulfide. The results showed that the P-Nb2O5 had a stronger ability to anchor the polysulfide than the G-Nb2O5, thus effectively increasing the cycling capacity of the battery. In addition, the strong interaction between the polysulfide and the P-Nb2O5 was verified by the low UV-Vis spectral intensity (Figure 6a). The LSV curve for the cathode is shown in Figure 6b, in which the P-Nb2O5 was more porous than the G-Nb2O5, which had a higher current response, indicating that the redox kinetics of the LiPSs were significantly higher. We further characterized the catalytic effect of the LiPSs by assembling a P-Nb2O5 symmetric battery with two identical electrodes and by using cv curves. Several highly overlapping CV curves, as seen in Figure 6c, showed that the electrodes had great reversibility. The redox current response of the P-Nb2O5 electrode was quite high compared to the G-Nb2O5, as shown in Figure 6d, indicating that the P-Nb2O5 could greatly improve the conversion of the LiPSs.
To demonstrate more clearly the electrocatalytic performance of the P-Nb2O5 in the Li-S batteries, the diffusion rate of the lithium ions (Li+) was evaluated by CV at different sweep rates. Further, according to the Randle–Sevick equation (Ip = 2.69 × 105n1.5aD0.5Li+  D l I + 0.5 0.5v0.5CLi), the diffusion coefficient of the lithium ions could be calculated from the peak current (Ip) and the square root of the scan rate (ν0.5), which have a linear relationship [31]. In the equation, Ip, n, a, DLi+, CLi, and v denote the peak current, number of charge transfers, geometric electrode area, geometric electrode area, concentration of lithium ions in the electrolyte, and scan rate, respectively. Therefore, Ip/v0.5 can be expressed as the diffusion capacity of the Li+. The slope of the final fit is proportional to the square root of DLi+. The slope of the P-Nb2O5 cathode was found to be higher than that of the G-Nb2O5 cathode, as shown in Figure 6e–h. The values of DLi+ at the P-Nb2O5 cathode peaks A, B, and C were 0.167, 0.091, and 0.077, respectively, and at the G-Nb2O5 cathode peaks A, B, and C, they were 0.109, 0.040, and 0.038, respectively, as can be seen in the graph. This demonstrated that the P-Nb2O5 facilitated the conversion of the polysulfides during the insertion/extraction of the Li+.

4. Conclusions

In summary, we developed a porous metal oxides positive electrode mixed material for reversible Li-S batteries. Using anodizing methods, the Nb2O5 apertures of nanoscale size were prepared, and they had abundant polysulfide-retaining capabilities and catalytically active sites. Our experiments demonstrated that using P-Nb2O5 as a mixing material for lithium-sulfur batteries resulted in a capacity of 602 mAh·g−1 after 400 cycles. The research provides an effective and viable approach to the development of high-performance LSBs. The porous Nb2O5 electrodes exhibited superior cycling stability due to their excellent polar adsorption and retention ability for dissolving intermediate polysulfides, which could be attributed to the influence of the porous structure of the Nb2O5, which provided higher pore volumes and more active sites for sulfur adaptation. The porous structure also reduced the problem of the volume change during the battery cycling. At same time, Nb2O5 showed a strong entrapment capacity for Li2S6, and it observably enhanced the redox kinetics from the long-chain shuttle effect. The amorphous characteristics achieved a high conversion of Li2Sx, and high sulfur utilization accompanied the growth of the particulate Li2S. The results of this study suggest that the low-cost and anodic oxidation method for producing niobium oxide with uniform pores on its surface has promising potential applications for the creation of a porous positive electrode mixed material for higher capacity Li-S batteries. This research offers practical prospects for the study of porous polar metal oxides. The experimental results also offer broad prospects for the development of amorphous porous structures of Nb2O5 in LSBs.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano13040777/s1. Figure S1. (a,b) SEM image of unoxidized niobium flakes; (c,d) AFM diagram of unoxidized niobium flakes; Figure S2. (a–c) SEM image of porous niobium oxide after sulphury mixing; (d–g) SEM images and elemental mapping images of P-Nb2O5 after sulphury mixing. Nb, O and S are represented in green, red and yellow, respectively; Figure S3. (a–c) SEM image of P-Nb2O5 electrode after cycling; Figure S4. XRD patterns of P-Nb2O5 and G-Nb2O5 after sulphury mixing; Figure S5. (a,b) SEM image of G-Nb2O5.

Author Contributions

Methodology, J.W.; formal analysis, J.W. and L.C.; investigation, J.W. and B.Z.; writing—original draft preparation, J.W.; writing—review and editing, J.W., Y.Z. and C.L.; supervision, H.W., Y.Z. and C.L.; project administration, Y.Z. and H.W.; funding acquisition, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (project number U21A2055); the Natural Science Foundation of Hebei Province of China (project number E2020202007); and the Natural Science Foundation of Tianjin of China (project number 21JCYBJC01380).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, Y.; Chen, W.; Lei, T.; Jiao, Y.; Huang, J.; Hu, A.; Gong, C.; Yan, C.; Wang, X.; Xiong, J. Strategies toward High-Loading Lithium–Sulfur Battery. Adv. Energy Mater. 2020, 10, 2000082. [Google Scholar] [CrossRef]
  2. Zhou, S.; Hu, J.; Liu, S.; Lin, J.-X.; Cheng, J.; Mei, T.; Wang, X.; Liao, H.-G.; Huang, L.; Sun, S.-G. Biomimetic Micro Cell Cathode for High Performance Lithium–Sulfur Batteries. Nano Energy 2020, 72, 104680. [Google Scholar] [CrossRef]
  3. Li, Z.; Zhou, H.-Y.; Zhao, F.-L.; Wang, T.-X.; Ding, X.; Han, B.-H.; Feng, W. Three-Dimensional Covalent Organic Frameworks as Host Materials for Lithium-Sulfur Batteries. Chin. J. Polym. Sci. 2020, 38, 550–557. [Google Scholar] [CrossRef]
  4. Ghosh, A.; Kumar, A.; Roy, A.; Nguyen, C.; Ahuja, A.; Adil, M.D.; Chatti, M.; Kar, M.; MacFarlane, D.R.; Mitra, S. Ultrathin Lithium Aluminate Nanoflake-Inlaid Sulfur as a Cathode Material for Lithium–Sulfur Batteries with High Areal Capacity. ACS Appl. Energy Mater. 2020, 3, 5637–5645. [Google Scholar] [CrossRef]
  5. Qi, Q.; Lv, X.; Lv, W.; Yang, Q.-H. Multifunctional Binder Designs for Lithium-Sulfur Batteries. J. Energy Chem. 2019, 39, 88–100. [Google Scholar] [CrossRef] [Green Version]
  6. Li, L.; Du, X.; Liu, G.; Zhang, Y.; Zhang, Z.; Li, J. Micro-/Mesoporous Co-NC Embedded Three-Dimensional Ordered Macroporous Metal Framework as Li-S Battery Cathode towards Effective Polysulfide Catalysis and Retention. J. Alloy. Compd. 2022, 893, 162327. [Google Scholar] [CrossRef]
  7. Gui, Y.; Chen, P.; Liu, D.; Fan, Y.; Zhou, J.; Zhao, J.; Liu, H.; Guo, X.; Liu, W.; Cheng, Y. TiO2 Nanotube/RGO Modified Separator as an Effective Polysulfide-Barrier for High Electrochemical Performance Li-S Batteries. J. Alloy. Compd. 2022, 895, 162495. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Tao, L.; Xie, C.; Wang, D.; Zou, Y.; Chen, R.; Wang, Y.; Jia, C.; Wang, S. Defect Engineering on Electrode Materials for Rechargeable Batteries. Adv. Mater. 2020, 32, 1905923. [Google Scholar] [CrossRef]
  9. Peng, L.; Fang, Z.; Zhu, Y.; Yan, C.; Yu, G. Holey 2D Nanomaterials for Electrochemical Energy Storage. Adv. Energy Mater. 2018, 8, 1702179. [Google Scholar] [CrossRef]
  10. Nazarian-Samani, M.; Haghighat-Shishavan, S.; Nazarian-Samani, M.; Kashani-Bozorg, S.F.; Ramakrishna, S.; Kim, K.-B. Perforated Two-Dimensional Nanoarchitectures for next-Generation Batteries: Recent Advances and Extensible Perspectives. Prog. Mater. Sci. 2021, 116, 100716. [Google Scholar] [CrossRef]
  11. Zhao, M.; Lu, Y.; Yang, Y.; Zhang, M.; Yue, Z.; Zhang, N.; Peng, T.; Liu, X.; Luo, Y. A Vanadium-Based Oxide-Nitride Heterostructure as a Multifunctional Sulfur Host for Advanced Li–S Batteries. Nanoscale 2021, 13, 13085–13094. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, L.L.; Zhu, D.; Cao, L.L.; Stephan, D.W. N-Heterocyclic Carbene Stabilized Parent Sulfenyl, Selenenyl, and Tellurenyl Cations (XH+, X = S, Se, Te). Dalton Trans. 2017, 46, 3095–3099. [Google Scholar] [CrossRef] [PubMed]
  13. Kovács, K.; Kiss, G.; Stenzel, M.; Zillgen, H. Anodic Oxidation of Niobium Sheets and Porous Bodies. J. Electrochem. Soc. 2003, 150, B361. [Google Scholar] [CrossRef]
  14. Dai, Y.; Ma, L.; Hu, J.; Wang, J.; Yan, H.; Zhang, H.; Wang, H.-Q.; Lai, C.; Li, W.; Zheng, J.-C. Highly Porous Niobium Ox-ide/Carbon Matrix Materials with Distinct Pseudocapacitive Performances in Aqueous Electrolytes. Electrochim. Acta 2021, 371, 137792. [Google Scholar] [CrossRef]
  15. Zhan, L.; Zhou, X.; Luo, J.; Fan, X.; Ning, X. Urchin-like Nb2O5/CNT Modified Separator for Lithium-Sulfur Batteries. Int. J. Hydrog. Energy 2022, 47, 27671–27679. [Google Scholar] [CrossRef]
  16. Chen, H.; Wang, J.; Zhao, Y.; Zeng, Q.; Zhou, G.; Jin, M. Three-Dimensionally Ordered Macro/Mesoporous Nb2O5/Nb4N5 Heterostructure as Sulfur Host for High-Performance Lithium/Sulfur Batteries. Nanomaterials 2021, 11, 1531. [Google Scholar] [CrossRef]
  17. Wang, J.; Li, G.; Luo, D.; Zhao, Y.; Zhang, Y.; Zhou, G.; Shui, L.; Wang, X. Amorphous–Crystalline-Heterostructured Niobium Oxide as Two-in-One Host Matrix for High-Performance Lithium–Sulfur Batteries. J. Mater. Chem. 2021, 9, 11160–11167. [Google Scholar] [CrossRef]
  18. Ma, Q.; Hu, M.; Yuan, Y.; Pan, Y.; Chen, M.; Zhang, Y.; Long, D. Colloidal Dispersion of Nb2O5/Reduced Graphene Oxide Nanocomposites as Functional Coating Layer for Polysulfide Shuttle Suppression and Lithium Anode Protection of Li-S Battery. J. Colloid Interface Sci. 2020, 566, 11–20. [Google Scholar] [CrossRef]
  19. Zhao, L.; Liu, G.; Zhang, P.; Sun, L.; Cong, L.; Wu, T.; Zhang, B.; Lu, W.; Xie, H.; Wang, H. Nitrogen–Sulfur Dual-Doped Porous Carbon Spheres/Sulfur Composites for High-Performance Lithium–Sulfur Batteries. RSC Adv. 2019, 9, 16571–16577. [Google Scholar] [CrossRef] [Green Version]
  20. Shang, C.; Wei, B.; Zhang, X.; Shui, L.; Wang, X.; Zhou, G. Vanadium Nitride-Decorated Lotus Root-like NCNFs as 3D Current Collector for Li-S Batteries. Mater. Lett. 2019, 236, 240–243. [Google Scholar] [CrossRef]
  21. Chu, R.X.; Lin, J.; Wu, C.Q.; Zheng, J.; Chen, Y.L.; Zhang, J.; Han, R.H.; Zhang, Y.; Guo, H. Reduced Graphene Oxide Coated Porous Carbon–Sulfur Nanofiber as a Flexible Paper Electrode for Lithium–Sulfur Batteries. Nanoscale 2017, 9, 9129–9138. [Google Scholar] [CrossRef] [PubMed]
  22. Lim, E.; Kim, H.; Jo, C.; Chun, J.; Ku, K.; Kim, S.; Lee, H.I.; Nam, I.-S.; Yoon, S.; Kang, K.; et al. Advanced Hybrid Supercapacitor Based on a Mesoporous Niobium Pentoxide/Carbon as High-Performance Anode. ACS Nano 2014, 8, 8968–8978. [Google Scholar] [CrossRef]
  23. Wang, M.; Xia, X.; Zhong, Y.; Wu, J.; Xu, R.; Yao, Z.; Wang, D.; Tang, W.; Wang, X.; Tu, J. Porous Carbon Hosts for Lithium–Sulfur Batteries. Chem. Eur. J. 2019, 25, 3710–3725. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, C.; Sun, L.; Wang, X.; Wang, L. Spherical Hybrid Hierarchical Porous Structure: A Plastic Model with Tunable Inner Pores for Lithium-Sulfur Batteries. Carbon 2019, 153, 691–698. [Google Scholar] [CrossRef]
  25. Wang, Y.; Huang, J.; Lu, J.; Lu, B.; Ye, Z. Fabricating Efficient Polysulfide Barrier via Ultrathin Tantalum Pentoxide Grown on Separator for Lithium–Sulfur Batteries. J. Electroanal. Chem. 2019, 854, 113539. [Google Scholar] [CrossRef]
  26. Kong, L.-L.; Zhang, Z.; Zhang, Y.-Z.; Liu, S.; Li, G.-R.; Gao, X.-P. Porous Carbon Paper as Interlayer to Stabilize the Lithium Anode for Lithium–Sulfur Battery. ACS Appl. Mater. Interfaces 2016, 8, 31684–31694. [Google Scholar] [CrossRef]
  27. Liu, F.; Wang, E.; Wu, C.; Sun, D.; Li, J. Ultrathin CoNi-Layered Double Hydroxide Grown on Nickel Foam as High-Performance Current Collector for Lithium-Sulfur Batteries. J. Solid State Electrochem. 2021, 25, 2033–2039. [Google Scholar] [CrossRef]
  28. Wen, Z.; Fang, W.; Chen, L.; Guo, Z.; Zhang, N.; Liu, X.; Chen, G. Anticorrosive Copper Current Collector Passivated by Self-Assembled Porous Membrane for Highly Stable Lithium Metal Batteries. Adv. Funct. Mater. 2021, 31, 2104930. [Google Scholar] [CrossRef]
  29. Wang, J.; Liu, Y.; Cheng, M.; Zhao, H.; Wang, J.; Zhao, Z.; Duan, X.; Wang, C.; Wang, J. Hierarchical Porous Carbon-Graphene-Based Lithium–Sulfur Batteries. Electrochim. Acta 2019, 318, 161–168. [Google Scholar] [CrossRef]
  30. Luan, C.; Chen, L.; Li, B.; Zhu, L.; Li, W. Electrochemical Dealloying-Enabled 3D Hierarchical Porous Cu Current Collector of Lithium Metal Anodes for Dendrite Growth Inhibition. ACS Appl. Energy Mater. 2021, 4, 13903–13911. [Google Scholar] [CrossRef]
  31. Tian, Y.; Zhang, X.; Zhang, Y.; Li, J.; Jia, A.; Liu, G.; Bakenov, Z. Cobalt-Doped Oxygen-Deficient Titanium Dioxide Coated by Carbon Layer as High-Performance Sulfur Host for Li/S Batteries. J. Alloys Compd. 2021, 861, 157969. [Google Scholar] [CrossRef]
Figure 1. (ac) SEM images of the porous niobium oxide obtained by anodic oxidation and (df) AFM images of the porous niobium oxide obtained by anodic oxidation. (gi) SEM images and elemental mapping images of P-Nb2O5. The Nb and O are represented in green and red, respectively.
Figure 1. (ac) SEM images of the porous niobium oxide obtained by anodic oxidation and (df) AFM images of the porous niobium oxide obtained by anodic oxidation. (gi) SEM images and elemental mapping images of P-Nb2O5. The Nb and O are represented in green and red, respectively.
Nanomaterials 13 00777 g001
Figure 2. (a,b) TEM images. (c) HRTEM image from the selected area. Inset: fast Fourier transform (FFT) images for framed areas. (df) TEM images and elemental mapping images of the P-Nb2O5.
Figure 2. (a,b) TEM images. (c) HRTEM image from the selected area. Inset: fast Fourier transform (FFT) images for framed areas. (df) TEM images and elemental mapping images of the P-Nb2O5.
Nanomaterials 13 00777 g002
Figure 3. (a) XRD patterns; (b) adsorption-desorption isotherms of the P-Nb2O5 and particles of P-Nb2O5; (c) pore size distribution of the P-Nb2O5 and G-Nb2O5; (d) Raman spectra; (e,f) high-resolution XPS spectra of the C1s, O1s, and Nb3d peaks of the anodized niobium samples; (e) Nb 3d; and (f) O1s.
Figure 3. (a) XRD patterns; (b) adsorption-desorption isotherms of the P-Nb2O5 and particles of P-Nb2O5; (c) pore size distribution of the P-Nb2O5 and G-Nb2O5; (d) Raman spectra; (e,f) high-resolution XPS spectra of the C1s, O1s, and Nb3d peaks of the anodized niobium samples; (e) Nb 3d; and (f) O1s.
Nanomaterials 13 00777 g003
Figure 4. TGA curves of the P-Nb2O5 and G-Nb2O5.
Figure 4. TGA curves of the P-Nb2O5 and G-Nb2O5.
Nanomaterials 13 00777 g004
Figure 5. (a) CV curves of the batteries with the P-Nb2O5 cathodes. (b) Discharge/charge curves at 0.2 C. (c) Cycling performances of the P-Nb2O5, G-Nb2O5, and CNT electrodes at 0.2 C for 100 cycles. (d) Cycling performances of the P-Nb2O5, G-Nb2O5, and CNT electrodes at different current rates (between 0.2 and 0.5). (e) Charge/discharge performances at different current rates (between 0.2 and 0.5). (f) EIS spectra of the batteries with the P-Nb2O5 and G-Nb2O5 electrodes. (g) Long-time cycling diagram for the P-Nb2O5, G-Nb2O5, and CNT cathodes at 1 C. (h) High load cycle performance graph for the P-Nb2O5.
Figure 5. (a) CV curves of the batteries with the P-Nb2O5 cathodes. (b) Discharge/charge curves at 0.2 C. (c) Cycling performances of the P-Nb2O5, G-Nb2O5, and CNT electrodes at 0.2 C for 100 cycles. (d) Cycling performances of the P-Nb2O5, G-Nb2O5, and CNT electrodes at different current rates (between 0.2 and 0.5). (e) Charge/discharge performances at different current rates (between 0.2 and 0.5). (f) EIS spectra of the batteries with the P-Nb2O5 and G-Nb2O5 electrodes. (g) Long-time cycling diagram for the P-Nb2O5, G-Nb2O5, and CNT cathodes at 1 C. (h) High load cycle performance graph for the P-Nb2O5.
Nanomaterials 13 00777 g005
Figure 6. (a) UV-vis spectra and optical images of the P-Nb2O5 and G-Nb2O5 adsorbed LiPSs solutions. (b) Li2S oxidation LSV curve. (c) Symmetric battery cv curves for the P-Nb2O5. (d) Symmetric battery cv curves for the P-Nb2O5 and G-Nb2O5. (e,g) CV curves at different sweep speeds. (f,h) Linear regression line for peak A, peak B, and peak C.
Figure 6. (a) UV-vis spectra and optical images of the P-Nb2O5 and G-Nb2O5 adsorbed LiPSs solutions. (b) Li2S oxidation LSV curve. (c) Symmetric battery cv curves for the P-Nb2O5. (d) Symmetric battery cv curves for the P-Nb2O5 and G-Nb2O5. (e,g) CV curves at different sweep speeds. (f,h) Linear regression line for peak A, peak B, and peak C.
Nanomaterials 13 00777 g006
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

Wang, J.; Chen, L.; Zhao, B.; Liang, C.; Wang, H.; Zhang, Y. Porous Nb2O5 Formed by Anodic Oxidation as the Sulfur Host for Enhanced Performance Lithium-Sulfur Batteries. Nanomaterials 2023, 13, 777. https://doi.org/10.3390/nano13040777

AMA Style

Wang J, Chen L, Zhao B, Liang C, Wang H, Zhang Y. Porous Nb2O5 Formed by Anodic Oxidation as the Sulfur Host for Enhanced Performance Lithium-Sulfur Batteries. Nanomaterials. 2023; 13(4):777. https://doi.org/10.3390/nano13040777

Chicago/Turabian Style

Wang, Jianming, Lu Chen, Bo Zhao, Chunyong Liang, Hongshui Wang, and Yongguang Zhang. 2023. "Porous Nb2O5 Formed by Anodic Oxidation as the Sulfur Host for Enhanced Performance Lithium-Sulfur Batteries" Nanomaterials 13, no. 4: 777. https://doi.org/10.3390/nano13040777

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