Next Article in Journal
The Dark Side of Growth: Are Shadow Economies Undermining the Global Climate Goal?
Previous Article in Journal
Projection of Photovoltaic System Adoption and Its Impact on a Distributed Power Grid Using Fuzzy Logic
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Eco-Friendly Gallic Acid-Tailored Binder with Synergistic Polarity Sites for High-Loading Lithium–Sulfur Batteries

1
College of Chemistry and Chemical Engineering, Anhui University of Technology, Ma’anshan 243002, China
2
College of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310058, China
*
Authors to whom correspondence should be addressed.
Sustainability 2025, 17(12), 5240; https://doi.org/10.3390/su17125240
Submission received: 9 April 2025 / Revised: 15 May 2025 / Accepted: 19 May 2025 / Published: 6 June 2025
(This article belongs to the Special Issue Sustainable Materials and Technologies for Battery Manufacturing)

Abstract

:
The development of polymer binders with tailored functionalities and green manufacturing processes is highly needed for high-performance lithium–sulfur batteries. In this study, a readily hydrolyzable 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane is utilized to prepare a water-based binder. Specifically, the acrolein produced by hydrolysis undergoes in situ polymerization to form a linear polymer, while the other hydrolyzed product, pentaerythritol, physically crosslinks these polymer chains via hydrogen bonding, generating a network polymer (BTU). Additionally, gallic acid (GA), a substance derived from waste wood, is further introduced into BTU during slurry preparation, forming a biphenol-containing binder (BG) with a multi-hydrogen-bonded structure. This resilience and robust cathode framework effectively accommodate volumetric changes during cycling while maintaining efficient ion and electron transport pathways. Furthermore, the abundant polar groups in BG enable strong polysulfide adsorption. As a result, sulfur cathode with a high mass loading of 5.3 mg cm−2 employing the BG (7:3) binder still retains an areal capacity of 4.7 mA h cm−2 after 50 cycles at 0.1 C. This work presents a sustainable strategy for battery manufacturing by integrating renewable biomass-derived materials and eco-friendly aqueous processing to develop polymer binders, offering a green pathway to high-performance lithium–sulfur batteries.

1. Introduction

Lithium-ion batteries (LIBs) have revolutionized energy storage for portable electronics and play an increasingly vital role in electric vehicles (EVs) and distributed energy storage systems [1,2,3,4]. However, conventional LIBs based on intercalation chemistry and transition metal materials face challenges in meeting the growing demand for higher energy density and lower cost [5,6]. In contrast, lithium–sulfur (Li-S) batteries offer a promising alternative due to their exceptionally high theoretical specific capacity (1675 mAh g−1) and energy density (2600 Wh kg−1), coupled with low material costs and environmental friendliness [7,8,9,10,11].
Despite these advantages, several challenges hinder the large-scale commercialization of Li-S batteries, primarily due to issues related to the sulfur cathode. The significant volume change (~80%) between sulfur (2.03 g cm−3) and lithium sulfide (1.66 g cm−3) can lead to cathode structural collapse and rapid capacity decay [12,13,14,15]. Additionally, the shuttle effect of polysulfides results in reduced Coulombic efficiency and increased internal resistance, while sulfur’s low conductivity limits reaction kinetics [16,17]. Therefore, stabilizing the sulfur cathode structure and effectively mitigating polysulfide dissolution are key to advancing Li-S battery technology [18].
To address these challenges, extensive research has explored strategies for improving cathode materials [19,20,21,22]. Highly conductive and porous carbon materials, such as carbon nanotubes [23,24] and nanospheres [25,26,27], are widely applied as sulfur hosts. To further accelerate polysulfide conversion, various metal catalysts have been combined with carbon materials in cathode or coated on the surface of separator, significantly improving cycling stability and rate performance [28]. Recently, rational electrolyte engineering, such as the introduction of redox mediators and selective anion screening, has been reported to effectively regulate the redox reactions of dissolved LiPSs [29,30,31,32]. Despite these advances, the substantial volume expansion of the sulfur cathode remains problematic, leading to microcrack formation and progressive electrode degradation over cycling [33,34,35]. If left unaddressed, these microcracks expand, eventually compromising structural integrity and battery performance [36,37].
Thus, developing innovative strategies to mitigate microcracks and suppress the polysulfide shuttle effect is critical for advancing the commercialization of lithium–sulfur (Li-S) batteries. As essential components of cathode fabrication, polymer binders play a pivotal role in maintaining electrode integrity and inhibiting polysulfide diffusion [9,38,39]. For instance, a self-healing binder was engineered through ring-opening polymerization of lipoic acid (with inherent disulfide bonds) combined with a thiol radical–polyphenol Michael addition reaction using gallic acid, enabling intrinsic crack repair and polysulfide anchoring [40]. Similarly, Liu et al. proposed a multifunctional binder where flame-retardant properties were introduced through the incorporation of hexachlorocyclotriphosphazene, while dynamic covalent bonds and polar groups provided self-healing capability and lithium polysulfide adsorption, respectively [41]. Although these binders address critical challenges in electrode stability and shuttle effects, scalability, cycling durability, and industrial processing compatibility remain unresolved.
Here, we present an environmentally friendly binder (BG) for sulfur cathodes, synthesized via acid-catalyzed hydrolysis and polymerization of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane [42] (Figure 1a). This process yields a polar polymer (BTU) with dense hydrogen-bonding sites for lithium polysulfide (LiPS) adsorption. When composited with gallic acid (GA), the resulting BG binder further strengthens LiPS retention through synergistic hydrogen-bonding and π-π interactions (Figure 1b). Compared to conventional polyvinylidene fluoride (PVDF)-based binders, the BG-based sulfur cathode demonstrates superior LiPS confinement and cycling stability, as confirmed by ultraviolet–visible (UV-Vis) spectroscopy analysis of the binder–LiPS mixture, X-ray photoelectron spectroscopy (XPS) analysis of Li anode surfaces, and cyclic voltammetry (CV) kinetics studies. The hydrogen-bond-rich 3D network dynamically accommodates sulfur’s volumetric expansion during cycling, preserving long-term electrode integrity, as evidenced by field-emission scanning electron microscopy (SEM) coupled with energy-dispersive spectroscopy (EDS) mapping. Systematic optimization confirms that a BTU-to-GA mass ratio of 7:3 achieves optimal electrochemical performance by balancing polar site density and structural flexibility. Finally, we assessed the applicability of the BG (7:3) binder under high-sulfur loadings and elevated current densities, exploring its viability for practical Li-S battery applications.

2. Materials and Methods

2.1. Experimental Materials

The 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane, gallic acid (GA), ammonium persulfate (APS), and sodium hydroxide (NaOH) used in this study were all purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). N-methylpyrrolidone (NMP) and sublimed sulfur (S, purity 99.95%) were provided by Aladdin (Shanghai, China). The Super P conductive agent was supplied by Timcal (Baudieu, Switzerland). The Ketjen black (KB) conductive agent was purchased from Suzhou Jilong Energy Technology Co., Ltd. (Suzhou, China).

2.2. Synthesis of BG (7:3) Binder

First, dissolve 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane (1 g) and APS (0.05 g) in deionized water (35 g), adjust the solution pH to 5 with NaOH, and react at 60 °C for 48 h under argon protection. The resulting product was freeze-dried and denoted as BTU. Secondly, when preparing the electrode slurry, dissolve BTU and GA in a 7:3 mass ratio in deionized water, stir for 3 min in a homogenizer, and use the resulting solution directly as a binder, denoted as BG (7:3).

2.3. Preparation of Electrodes

To prepare the sulfur cathode, sublimed sulfur and conductive agent (KB) were uniformly mixed in a ratio of 7:3, and then heated at 155 °C for 12 h to obtain the S/KB composite material. Next, a slurry containing a mass ratio of 8:1:1 of S/KB composite, Super P, and binder was cast onto carbon-coated aluminum foil and vacuum-dried at 60 °C for 12 h. The resulting electrodes were punched into 12 mm diameter disks for use. Each cathode contained 1.0–1.2 mg of sulfur per cm2. To prepare electrodes with a high active material loading, nickel foam was used as the current collector.

2.4. Characterization

Fourier transform infrared spectroscopy (FTIR-850) was employed to analyze the chemical structure of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane before and after hydrolysis. The surface morphology of the sulfur cathode and lithium anode was observed using a field emission scanning electron microscope (SEM, model Nano SEM430) (Product of Carl Zeiss Optics Co., Ltd. Guangzhou, China). The elemental composition of the sulfur cathode was determined using an X-ray photoelectron spectrometer (XPS, product of Thermo Scientific, Massachusetts, USA) with an Al X-ray source. The measurements of Nuclear Magnetic Resonance (NMR) spectra were carried out on a Mercury VX-300 spectrometer (Bruker (Beijing) Scientific Technology Co., Ltd. (Beijing, China)).

2.5. Electrochemical Measurement

Using the Neware CT-4008 battery testing system, constant–current charge–discharge cycling tests were conducted on half-cells, with the test voltage range set at 1.7–2.8 V. CR2025 coin cells were assembled in an argon-filled glove box, using Celgard 2400 as the separator material. The half-cell electrolyte was composed of 1 M LiTFSI dissolved in a 1:1 (volume ratio) mixture of 1,3-dioxolane (DOL) and 1,2-dimethoxyethane (DME) solvents, with an additional 1% lithium nitrate (LiNO3) added as an auxiliary additive. For the electrochemical cycling performance of the low-loading sulfur cathode, the 0.5 C battery was activated under 0.3 C conditions for the first two cycles, the 1 C battery was activated under 0.3 C conditions for the first three cycles, and the 2 C battery was activated under 0.3 C conditions for the first five cycles. For the cycling performance of the high-loading sulfur cathode, the first cycle of the 0.1 C and 0.2 C batteries was activated under 0.03 C conditions. Electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) tests were both performed using the CHI760E electrochemical workstation. CV tests were conducted within a voltage range of 1.7–2.8 V, at scanning rates of 0.05, 0.10, 0.15, 0.2, and 0.25 mV s−1. The diffusion coefficient of lithium ions (Li+) was calculated using the classical Randles–Sevcik equation:
I p = ( 2.69   ×   10 5 ) n 1.5 A D Li + 0.5 v 0.5 C Li +
where Ip represents the peak current, C Li + represents the concentration of lithium ions in the electrolyte, A represents the electrode surface area, v represents the scanning rate (V s−1), n represents the number of electrons in the reaction (n = 2), and D Li + represents the diffusion coefficient of lithium ions (cm2 s−1). EIS tests were conducted at open-circuit voltage, with a frequency range of 1 MHz to 1 Hz, and an AC voltage amplitude of 10 mV. All tests were conducted at room temperature.

2.6. Preparation of Li2S6 Solution

Sulfur powder (S) and lithium sulfide (Li2S) were added to a solution of 1,2-dimethoxyethane (DME) and 1,3-dioxolane (DOL) in a 1:1 volume ratio, with a molar ratio of 5:1 for S and Li2S. The mixture was then reacted at 70 °C for 24 h to obtain a lithium polysulfide (LiPS) solution with an average molecular formula of Li2S6.

3. Results

3.1. Structure of BG Binder

Using 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane as the raw material, the BTU binder was synthesized via one-step radical polymerization [43,44]. Subsequently, BTU and GA were mixed in a specific ratio to prepare the BG (7:3) binder. As shown in Figure 2a,b, 1H NMR spectroscopy confirmed the complete hydrolysis of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane, as evidenced by the disappearance of proton peaks corresponding to C=C double bonds at 4.2~5.8 ppm. In the 1H NMR spectrum of BTU, new proton peaks appeared at 1 ppm, attributed to protons on the terminal carbon of the BTU chain, while a characteristic –CHO proton peak was detected at 9.7 ppm. Additionally, the emergence of an –OH proton peak at 1.7 ppm further supported the successful synthesis of BTU. The appearance of the C=O bond and the disappearance of the C=C bond in Fourier transform infrared (FTIR) spectroscopy further indicated the successful preparation of the BTU binder (Figure 2c). Moreover, a broad stretching vibration peak at 3423 cm−1, attributed to –OH groups from pentaerythritol, confirmed the presence of hydrogen-bonding interactions.
The solubility of GA in water is significantly affected by temperature, and generally, its solubility is lower in cold water. As shown in Figure 3a,b, we added GA (0.01 g) into water (1 mL) with BTU (0.023 g) and without BTU, respectively. It was observed that GA could not completely dissolve at room temperature, appearing as a white opaque suspension. After placing it in a 50 °C oven for 10 min, GA completely dissolved. Subsequently, when placed in a 3 °C environment, GA precipitated. At room temperature, GA dissolved in water containing BTU, forming a transparent liquid. After placing it in a 3 °C environment for 10 min, GA did not precipitate. The proportion of GA was gradually increased by mass. When the mass ratio of GA to BTU reacheds 7:3, partial precipitation of GA occurred at 3 °C, indicating that GA had reached its solubility limit. This phenomenon is attributed to hydrogen-bonding interactions between GA and the BTU network, which prevent GA from precipitating in aqueous solution. These findings confirm the strong interactions between BTU and GA, enabling the formation of an effective binder for Li-S batteries through synergistic hydrogen-bonding interactions.

3.2. Electrochemical Performance

To figure out the optimal BTU-to-GA ratio, electrochemical cycling tests were conducted on BG binders with different compositions. Figure 4a presents the cycling performance of batteries with various binders at 0.5 C. After 150 cycles, the battery with the BG (7:3) binder delivers a reversible capacity of 928.9 mAh g−1, outperforming BG (8:2), BTU, BG (6:4), and PVDF. This enhancement is attributed to the ability of GA, at a 7:3 ratio, to effectively regulate the polar sites of BTU, leading to superior LiPS adsorption and accelerated Li2S redox reactions. Even at higher current densities, the BG (7:3) binder enables excellent performance, as shown in Figure 4b,c. At 1 C, the battery achieves a reversible capacity of 845.4 mAh g−1 after 150 cycles, while at 2 C, it retains 727.1 mAh g−1, after 200 cycles. In contrast, the PVDF-based battery suffers from rapid capacity decay, reaching only 449.3 mAh g−1 and 455.7 mAh g−1 at 1 C and 2 C, respectively. Figure 4d–f illustrate the typical charge–discharge profiles of Li-S batteries, where BG (7:3) exhibits lower polarization and extended discharge plateaus compared to BTU and PVDF. This suggests enhanced redox kinetics, improved lithium-ion conductivity, and higher active material utilization. Therefore, subsequent analyses primarily focus on the BG (7:3) binder.
As shown in Figure 4g–j, the BG (7:3) battery delivers higher reversible discharge capacities across various current densities (0.1, 0.3, 0.5, 1, and 2 C), reaching 1158.2, 1065.8, 1028.1, 968.8, and 861.7 mAh g−1, respectively, and surpassing both BTU and PVDF. When the current density is returned to 0.1 C, the BG (7:3) battery can still achieve a reversible discharge capacity of 1063.9 mAh g−1. Notably, as the current rate increases, the capacity disparity between BG (7:3) and PVDF widens, signifying faster sulfur redox kinetics in BG (7:3)-based cells.
The practical application of the BG (7:3) binder in high-energy-density Li-S batteries with high-sulfur–mass loadings was further validated. As shown in Figure 4k, at a S loading of 3 mg cm−2, the battery maintains a capacity of 813.2 mAh g−1 after 100 cycles at a current density of 0.2 C. When the S loading increases to 3.7 and 5.3 mg cm−2, the capacities remain high at 916.6 and 696.9 mAh g−1, respectively, after 100 cycles at 0.1 C. Figure 4l reveals that the charge–discharge profiles for a S loading of 3 mg cm−2, remain nearly identical between the 10th and 30th cycles, indicating stable redox kinetics and long-term cycling stability. This superior performance is attributed to the BG (7:3) binder’s strong LiPS adsorption capability and its ability to preserve electrode integrity, even under high-sulfur loading conditions. Additionally, the electrochemical performance of the Li-S batteries in this work was compared with that reported in the relevant literature, as shown as Table 1.
The lithium-ion (Li+) diffusion rate is a critical parameter governing polysulfide (LiPS) conversion kinetics in Li-S batteries [45]. To evaluate this, CV tests were performed at scan rates ranging from 0.05 to 0.25 mV s−1, as shown in Figure 5a–c. The two reduction peaks observed at 2.3 V (C1) and 2.04 V (C2) signify the stepwise reduction of S8 to long-chain lithium polysulfides (Li2Sx, 4 ≤ x ≤ 6), culminating in the formation of Li2S2/Li2S. Conversely, the oxidation peak at 2.33 V (A) corresponds to the reverse process, starting from the oxidation of Li2S2/Li2S to Li2Sx and culminating in a reversion to S8. All cathodes exhibited increasing peak current densities and well-defined redox peaks at higher scan rates, confirming robust electrochemical stability. The linear correlation between peak current (Ip) and the square root of the scan rate (v0.5) indicates a diffusion-limited redox process. As shown in Figure 5d–f, the BG(7:3)-based battery showed the steepest slope across the three peaks, indicative of superior Li+ diffusivity. Using the Randles–Sevcik equation [48], Li+ diffusion coefficients ( D Li + ) were calculated. Table 2 further illustrated that the characteristic peak values ( D Li + ) for the BG(7:3)/S battery surpassed those of the BTU/S and PVDF/S batteries.
The BG (7:3) cathode demonstrated significantly higher D Li + values than those of BTU and PVDF, which we attribute to its enhanced LiPS adsorption capability. This adsorption shortens the electron transfer pathway and mitigates polysulfide shuttling, thereby improving reaction kinetics. Furthermore, the CV curves of the BG (7:3) cathode during initial cycles (Figure 5g–i) showed negligible shifts in peak positions (<50 mV) and retained >95% of their original intensity, underscoring exceptional electrochemical reversibility and cycling stability. In contrast, BTU and PVDF cathodes exhibited progressive polarization and capacity fade, consistent with their lower D Li + values.
Notably, the BG (7:3)-based cathode exhibits a positive shift in reduction peaks and a negative shift in oxidation peaks compared to its BTU and PVDF counterparts, indicating reduced polarization and enhanced reversibility (Figure 6a). The BG (7:3)-based cathode exhibits a negative shift in peak as compared to BTU and PVDF counterparts, which we attribute to its better ionic conductivity. In addition, its network structure effectively enables a compact electrode structure, which shortens the electron transfer pathway, thereby improving reaction kinetics. The strong polysulfides absorbability of BG (7:3) leads to a much suppressed shuttling effect, resulting in a higher response current and larger integrated area in the CV profile. Furthermore, the BG (7:3) cathode shows sharper and more intense redox peaks, reflecting faster reaction kinetics and improved lithium-ion diffusion. This is further corroborated by Tafel analysis (Figure 6b–d), where the BG (7:3) binder demonstrates significantly lower polarization slopes at critical peaks: (1) peak C1 (Li2Sn to Li2S/Li2S): 45.8 mV dec−1, (2) peak C2 (S8 to Li2Sn): 21.1 mV dec−1, and (3) peak A (Li2S to S8): 88.26 mV dec−1. These values are markedly lower than those of BTU (51.57, 41.61, 109.37 mV dec−1) and PVDF (51.92, 43.67, 126.62 mV dec−1), confirming the superior conversion kinetics enabled by the BG (7:3) binder. The reduced Tafel slopes align with its enhanced polysulfide adsorption capability and efficient charge transfer, which collectively suppress the shuttle effect and stabilize the sulfur cathode.

3.3. Absorbability of Binder to Polysulfide Lithium

To further elucidate the superior adsorption capability of the BG (7:3) binder toward lithium polysulfides (LiPSs) and its ability to maintain cathode integrity, in situ electrochemical impedance spectroscopy (EIS) and UV-Vis spectroscopy comparisons were conducted, along with a post-cycling cathode morphology analysis. As shown in Figure 7a, in situ EIS measurements were performed on Li-S batteries with the BG (7:3) binder during a 0.1 C charge–discharge process. The Nyquist plots consist of three components: electrolyte resistance (R0), interface resistance (RS), and Warburg impedance (W0). During the discharge process, a slight increase in the high-frequency semicircle was observed, attributed to electrolyte viscosity changes caused by partial LiPS dissolution. As the discharge depth increases, R0 decreases due to LiPS oxidation into insoluble Li2S, accompanied by a reduction in the semicircle diameter. Simultaneously, a new semicircle emerged at lower frequencies, reflecting an increase in Rs due to Li2S deposition. These changes were fully reversible during charging, demonstrating the BG (7:3) binder’s efficient LiPS adsorption and rapid redox kinetics. In contrast, PVDF-based batteries exhibited irreversible increases in R0 and RS during cycling (Figure 7b), indicative of poor LiPS confinement and sluggish reaction dynamics, likely due to extended charge transfer paths and higher energy barriers.
The LiPS absorbability of the BG (7:3) binder was evaluated using UV-Vis spectroscopy (Figure 7c). After 8 h of static adsorption with the Super P-BG (7:3) mixture, the LiPS solution turned a faint yellow color, whereas the Super P-BTU-containing solution remained a darker yellow, and the Super P-PVDF-containing solution showed no visible change. The BG (7:3) sample also displayed attenuated absorbance near 400 nm, characteristic of LiPS, further verifying its strong chemisorption capability via polar functional groups. Further investigation of the Li anode surface composition was conducted via XPS characterization (Figure 7d). The S 2p spectrum of the BG (7:3)-based Li anode displayed four peaks at 162.0, 163.4, 167.1, and 169.1 eV, which can be attributed to polysulfides, sulfates, and thiosulfates, respectively. Compared to BTU and PVDF binders, the BG (7:3) binder showed significantly weaker sulfate/thiosulfate peaks, confirming its ability to suppress LiPS shuttling.
Furthermore, the PVDF-based batteries exhibited rapid capacity decay during cycling. Post-cycling SEM and EDS analyses of the cathodes revealed the underlying cause. As shown in Figure 7e, the PVDF-based cathode surface displayed severe protrusions and cracks after 100 cycles, resulting from active material disconnection during volume changes. This led to rapid capacity fade and poor cycling stability. In contrast, the BG (7:3) binder, with its hydrogen-bond-rich structure, effectively buffered sulfur volume expansion, preserving cathode integrity without significant structural damage. EDS mapping (Figure 7e–h) confirmed a more uniform sulfur distribution and higher sulfur content in the BG (7:3) cathode, attributed to polar site-mediated LiPS adsorption and shuttle effect suppression.

4. Conclusions

In summary, we developed an environmentally friendly BG (7:3) binder, which greatly improved the electrochemical performance and cycling stability of Li-S batteries compared to conventional binders. By tuning the polarity sites and polymer network structure through the synergistic integration of BTU and GA, the BG (7:3) binder effectively mitigates sulfur volume expansion-induced cathode cracking while optimizing lithium polysulfide (LiPS) adsorption and redox kinetics, as evidenced by post-cycling SEM morphology, CV measurements, and XPS analyses. These advancements give Li-S batteries exceptional cycling stability and capacity retention. Specifically, the BG (7:3)-based battery delivers a reversible capacity of 727.1 mAh g−1 after 200 cycles at 2 C, and even under a sulfur loading of 5.3 mg cm−2, it retains 890 mAh g−1 after 50 cycles at 0.1 C. This work not only addresses critical challenges in Li-S battery commercialization but also underscores the pivotal role of sustainable material design in advancing next-generation energy storage technologies.

Author Contributions

Conceptualization, X.J. and B.J.; methodology, X.J., S.L. and J.W.; software, X.J. and S.L.; validation, X.J. and J.W.; formal analysis, X.J.; investigation, X.J.; resources, X.H.; data curation, X.J. and J.Z.; writing—original draft preparation, X.J.; writing—review and editing, C.W., J.Z. and B.J.; visualization, X.J.; supervision, X.H. and B.J.; project administration, C.W. and B.J.; funding acquisition, C.W. and B.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (no. 22108002 and 22108238), the Excellent Young Scholars Program of Natural Science Foundation Anhui Province (No. 2408085Y005), and the Excellent Youth Scholars Program of Higher Education Institutions of Anhui Province (No. 2024AH030008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors are grateful to the Shiyanjia Lab (https://www.shiyanjia.com) (accessed on 24 January 2025) for characterizing the materials used in this study. We express our gratitude to English editing software for refining the language and checking for grammatical errors in our manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Guo, R.; Yang, Y.; Huang, X.L.; Zhao, C.; Hu, B.; Huo, F.; Liu, H.K.; Sun, B.; Sun, Z.; Dou, S.X. Recent Advances in Multifunctional Binders for High Sulfur Loading Lithium-Sulfur Batteries. Adv. Funct. Mater. 2023, 34, 2307108. [Google Scholar] [CrossRef]
  2. Li, Z.; Wan, Z.; Lin, Z.; Zheng, M.; Zheng, J.; Qian, S.; Wang, Y.; Song, T.; Lin, Z.; Lu, J. A highly elastic and Li-ion conductive binder enables stable operation of silicon microparticle anodes in high-capacity and high-energy-density pouch cells. Energy Environ. Sci. 2025, 18, 2365–2380. [Google Scholar] [CrossRef]
  3. Tian, X.; Cheng, Y.; Zhou, Y.; Zhang, B.; Wang, G. Long-cycling and high-loading lithium–sulfur battery enabled by free-standing three-dimensional porous NiCo2O4 nanosheets. Appl. Energy 2023, 334, 120694. [Google Scholar] [CrossRef]
  4. Xi, G.; Zhang, Z.; Zhong, L.; Wang, S.; Xiao, M.; Han, D.; Huang, S.; Meng, Y. Novel Aliphatic polycarbonate binders for Solvent-free manufacturing High–loading cathodes of high-performance lithium-ion batteries. Chem. Eng. J. 2024, 485, 149983. [Google Scholar] [CrossRef]
  5. Jin, B.; Dolocan, A.; Liu, C.; Cui, Z.; Manthiram, A. Regulating Anode-Electrolyte Interphasial Reactions by Zwitterionic Binder Chemistry in Lithium-Ion Batteries with High-Nickel Layered Oxide Cathodes and Silicon-Graphite Anodes. Angew. Chem. Int. Ed. 2024, 63, e202408021. [Google Scholar]
  6. Lin, X.; Liu, X.; Tong, Y.; Zhou, X.; Li, J.; Song, J.; Feng, X.; Liu, R.; Shi, L.; Yu, A.; et al. A Poly(Ionic Liquid)-Based Polymer Binder for Endurable Lithium-Sulfur Batteries. Adv. Funct. Mater. 2024, 34, 2406985. [Google Scholar] [CrossRef]
  7. Liang, S.; Zhang, J.; Jia, C.; Luo, Z.; Zhang, L. Water soluble polymer binder with good mechanical property and ionic conductivity for high performance lithium sulfur battery. Carbon 2024, 222, 118807. [Google Scholar] [CrossRef]
  8. Liu, H.; Wu, Q.; Guan, X.; Liu, M.; Wang, F.; Li, R.; Xu, J. Ionically Conductive Self-Healing Polymer Binders with Poly(ether-thioureas) Segments for High-Performance Silicon Anodes in Lithium-Ion Batteries. ACS Appl. Energy Mater. 2022, 5, 4934–4944. [Google Scholar] [CrossRef]
  9. Mu, P.; Sun, C.; Gao, C.; Li, L.; Zhang, H.; Li, J.; Li, C.; Dong, S.; Cui, G. Dual Network Electrode Binder toward Practical Lithium–Sulfur Battery Applications. ACS Energy Lett. 2023, 8, 3733–3741. [Google Scholar] [CrossRef]
  10. Reddy, B.S.; Ahn, H.-J.; Ahn, J.-H.; Cho, G.-B.; Cho, K.-K. Cost-effective water-soluble three-dimensional cross-linked polymeric binder for high-performance lithium–sulfur batteries. J. Energy Storage 2023, 66, 107400. [Google Scholar] [CrossRef]
  11. Yu, Z.; Gao, T.; Le, T.; Wang, W.; Wang, L.; Yang, Y. A homemade self-healing material utilized as multi-functional binder for long-lifespan lithium–sulfur batteries. J. Mater. Sci. Mater. Electron. 2019, 30, 5536–5543. [Google Scholar] [CrossRef]
  12. Chen, J.; Geng, X.; Wang, C.; Hou, X.; Wang, H.; Rong, Q.; Sun, N.; Liu, W.; Hu, L.; Fu, X.; et al. An interweaving 3D ion-conductive network binder for high-loading and lean-electrolyte lithium–sulfur batteries. J. Mater. Chem. A 2024, 12, 11038–11048. [Google Scholar] [CrossRef]
  13. Gao, Q.; Shen, Z.; Guo, Z.; Li, M.; Wei, J.; He, J.; Zhao, Y. Metal Coordinated Polymer as Three-Dimensional Network Binder for High Sulfur Loading Cathode of Lithium–Sulfur Battery. Small 2023, 19, 2301244. [Google Scholar] [CrossRef] [PubMed]
  14. Gong, Q.; Hou, L.; Li, T.; Jiao, Y.; Wu, P. Regulating the Molecular Interactions in Polymer Binder for High-Performance Lithium-Sulfur Batteries. ACS Nano 2022, 16, 8449–8460. [Google Scholar] [CrossRef]
  15. Wang, W.; Hou, M.; Han, F.; Yu, D.; Liu, J.; Zhang, Q.; Yu, F.; Wang, L.; He, M. Three-in-one LaNiO3 functionalized separator boosting electrochemical stability and redox kinetics for high-performance Li-S battery. J. Energy Chem. 2023, 82, 581–591. [Google Scholar] [CrossRef]
  16. Guo, R.; Wang, D.; Ding, P.; Chen, Y.; Zhao, H. Dual Cross-Linked Multifunctional Binder for High-Performance Lithium–Sulfur Batteries. ACS Appl. Energy Mater. 2023, 6, 8590–8598. [Google Scholar] [CrossRef]
  17. Li, D.; Wang, W.; Liu, J.; He, M. Hierarchical lamellar single-walled carbon nanotube aerogel interlayers for stable lithium-sulfur batteries with high-sulfur-loading. Chem. Eng. J. 2023, 461, 142031. [Google Scholar] [CrossRef]
  18. Chen, Z.; Lu, M.; Qian, Y.; Yang, Y.; Liu, J.; Lin, Z.; Yang, D.; Lu, J.; Qiu, X. Ultra-Low Dosage Lignin Binder for Practical Lithium-Sulfur Batteries. Adv. Energy Mater. 2023, 13, 2300092. [Google Scholar] [CrossRef]
  19. Zhao, M.; Peng, H.-J.; Li, B.-Q.; Huang, J.-Q. Kinetic Promoters for Sulfur Cathodes in Lithium-Sulfur Batteries. Acc. Chem. Res. 2024, 57, 545–557. [Google Scholar] [CrossRef]
  20. Jin, B.; Lai, T.; Manthiram, A. High-Mass-Loading Anode-Free Lithium-Sulfur Batteries Enabled by a Binary Binder with Fast Lithium-Ion Transport. ACS Energy Lett. 2023, 8, 3767–3774. [Google Scholar] [CrossRef]
  21. Guo, D.; Thomas, S.; El-Demellawi, J.K.; Shi, Z.; Zhao, Z.; Canlas, C.G.; Lei, Y.; Yin, J.; Zhang, Y.; Hedhili, M.N.; et al. Electrolyte engineering for thermally stable Li-S batteries operating from −20 °C to 100 °C. Energy Environ. Sci. 2024, 17, 8151–8161. [Google Scholar] [CrossRef]
  22. Guo, D.; Li, M.; Hedhili, M.N.; Tung, V.; Li, Y.; Lai, Z. Asymmetric cathode membrane with tunable positive charge networks for highly stable Li-S batteries. Energy Storage Mater. 2020, 25, 33–40. [Google Scholar] [CrossRef]
  23. Cao, Y.; Li, X.L.; Zheng, M.S.; Yang, M.P.; Yang, X.L.; Dong, Q.F. Ultra-high Rates and Reversible Capacity of Li-S Battery with a Nitrogen-doping Conductive Lewis Base Matrix. Electrochim. Acta 2016, 192, 467–474. [Google Scholar] [CrossRef]
  24. Jia, X.; Liu, B.; Liu, J.; Zhang, S.; Sun, Z.; He, X.; Li, H.; Wang, G.; Chang, H. Fabrication of NiO-carbon nanotube/sulfur composites for lithium-sulfur battery application. RSC Adv. 2021, 11, 10753–10759. [Google Scholar] [CrossRef]
  25. Fang, Y.; Yao, Y.; Yang, H.; Fan, Y.; Nomura, N.; Zhou, W.; Ni, D.; Li, X.; Jiang, W.; Qiu, P.; et al. Incorporating Cobalt Nanoparticles in Nitrogen-Doped Mesoporous Carbon Spheres through Composite Micelle Assembly for High-Performance Lithium–Sulfur Batteries. ACS Appl. Mater. Interfaces 2021, 13, 38604–38612. [Google Scholar] [CrossRef]
  26. Yang, F.; Huang, K. I, N Co-doped hierarchical micro/mesoporous carbon modified separator for enhanced electrochemical performances of lithium-sulfur batteries. Mater. Res. Express 2021, 8, 115002. [Google Scholar] [CrossRef]
  27. Yu, C.-H.; Yen, Y.-J.; Chung, S.-H. Nanoporosity of Carbon-Sulfur Nanocomposites toward the Lithium-Sulfur Battery Electrochemistry. Nanomaterials 2021, 11, 1518. [Google Scholar] [CrossRef]
  28. Zheng, F.; Zhang, Y.; Ding, G.; Xiao, Y.; Wei, L.; Su, J.; Wang, C.; Chen, Q.; Wang, H. Pentagon Defects Accelerating Polysulfides Conversion Enabled High-Performance Sodium-Sulfur Batteries. Adv. Funct. Mater. 2023, 34, 2310598. [Google Scholar] [CrossRef]
  29. Li, Z.; Ma, Z.; Wang, Y.; Chen, R.; Wu, Z.; Wang, S. LDHs derived nanoparticle-stacked metal nitride as interlayer for long-life lithium sulfur batteries. Sci. Bull. 2018, 63, 169–175. [Google Scholar] [CrossRef]
  30. Song, Y.-W.; Shen, L.; Yao, N.; Feng, S.; Cheng, Q.; Ma, J.; Chen, X.; Li, B.-Q.; Zhang, Q. Anion-Involved Solvation Structure of Lithium Polysulfides in Lithium-Sulfur Batteries. Angew. Chem. Int. Ed. 2024, 63, e202400343. [Google Scholar] [CrossRef]
  31. Liu, Y.; An, Y.; Fang, C.; Ye, Y.; An, Y.; He, M.; Jia, Y.; Hong, X.; Liu, Y.; Gao, S.; et al. Surface-localized phase mediation accelerates quasi-solid-state reaction kinetics in sulfur batteries. Nat. Chem. 2025, 17, 614–623. [Google Scholar] [CrossRef]
  32. Lai, T.; Bhargav, A.; Manthiram, A. Lithium Tritelluride as an Electrolyte Additive for Stabilizing Lithium Deposition and Enhancing Sulfur Utilization in Anode-Free Lithium-Sulfur Batteries. Adv. Funct. Mater. 2023, 33, 2304568. [Google Scholar] [CrossRef]
  33. Chen, Z.; Chen, T.; Wang, J.; Li, P.; Liu, J.; Chen, W.; Yang, Z.; Deng, Y.; Chang, J.; Yang, Y. A Low-Dosage Flame-Retardant Inorganic Polymer Binder for High-Energy-Density and High-Safety Lithium-Sulfur Batteries. Adv. Energy Mater. 2024, 14, 2401568. [Google Scholar] [CrossRef]
  34. Jin, B.; Yang, L.; Zhang, J.; Cai, Y.; Zhu, J.; Lu, J.; Hou, Y.; He, Q.; Xing, H.; Zhan, X.; et al. Bioinspired Binders Actively Controlling Ion Migration and Accommodating Volume Change in High Sulfur Loading Lithium-Sulfur Batteries. Adv. Energy Mater. 2019, 9, 1902938. [Google Scholar] [CrossRef]
  35. Wan, Z.; Li, S.; Tang, W.; Dai, C.; Yang, J.; Lin, Z.; Qiu, J.; Ling, M.; Lin, Z.; Li, Z. Exploring the optimal molecular weight of polyacrylic acid binder for silicon nanoparticle anodes in lithium-ion batteries. J. Energy Chem. 2025, 105, 76–86. [Google Scholar] [CrossRef]
  36. Jin, B.; Wang, D.; Zhu, J.; Guo, H.; Hou, Y.; Gao, X.; Lu, J.; Zhan, X.; He, X.; Zhang, Q. A Self-Healable Polyelectrolyte Binder for Highly Stabilized Sulfur, Silicon, and Silicon Oxides Electrodes. Adv. Funct. Mater. 2021, 31, 2104433. [Google Scholar] [CrossRef]
  37. Wu, Z.; Ma, Y.; Li, S.; Que, L.; Chen, H.; Hao, F.; Tao, X.; Xing, H.; Ye, J.; Qian, D.; et al. Damage-Tolerant and Self–Repairing Web–Like Borate Type Binder Enable Stable Operation of Efficient Si-Based Anodes. Small 2024, 20, 202401345. [Google Scholar] [CrossRef]
  38. He, Y.; Jing, X.; Lai, T.; Jiang, D.; Wan, C.; Postnikov, P.S.; Guselnikova, O.; Xu, L.; He, X.; Yamauchi, Y.; et al. Amphipathic emulsion binder for enhanced performance of lithium-sulfur batteries. J. Mater. Chem. A 2024, 12, 12681–12690. [Google Scholar] [CrossRef]
  39. Wang, W.; Hua, L.; Zhang, Y.; Wang, G.; Li, C. A Conductive Binder Based on Mesoscopic Interpenetration with Polysulfides Capturing Skeleton and Redox Intermediates Network for Lithium Sulfur Batteries. Angew. Chem. Int. Ed. 2024, 63, e202405920. [Google Scholar] [CrossRef]
  40. Wen, Y.; Lin, X.; Sun, X.; Wang, S.; Wang, J.; Liu, H.; Xu, X. A biomass-rich, self-healable, and high-adhesive polymer binder for advanced lithium-sulfur batteries. J. Colloid Interface Sci. 2024, 660, 647–656. [Google Scholar] [CrossRef]
  41. Liu, M.; Chen, P.; Pan, X.; Pan, S.; Zhang, X.; Zhou, Y.; Bi, M.; Sun, J.; Yang, S.; Vasiliev, A.L.; et al. Synergism of Flame-Retardant, Self-Healing, High-Conductive and Polar to a Multi-Functional Binder for Lithium-Sulfur Batteries. Adv. Funct. Mater. 2022, 32, 2205031. [Google Scholar] [CrossRef]
  42. Van der Heyden, K.; Babooram, K.; Ahmed, M.; Narain, R. Protein encapsulation and release from degradable sugar based hydrogels. Eur. Polym. J 2009, 45, 1689–1697. [Google Scholar] [CrossRef]
  43. Zhou, W.; Zhang, H.; Liu, Y.; Zou, X.; Shi, J.; Zhao, Y.; Ye, Y.; Yu, Y.; Guo, J. Preparation of calcium alginate/polyethylene glycol acrylate double network fiber with excellent properties by dynamic molding method. Carbohydr. Polym. 2019, 226, 115277. [Google Scholar] [CrossRef]
  44. Şarkaya, K.; Yildirim, M.; Alli, A. One-step preparation of poly(NIPAM-pyrrole) electroconductive composite hydrogel and its dielectric properties. J. Appl. Polym. Sci. 2021, 138, 50527. [Google Scholar] [CrossRef]
  45. Yi, H.; Lan, T.; Yang, Y.; Zeng, H.; Zhang, T.; Tang, T.; Wang, C.; Deng, Y. A robust aqueous-processable polymer binder for long-life, high-performance lithium sulfur battery. Energy Storage Mater. 2019, 21, 61–68. [Google Scholar] [CrossRef]
  46. Chen, F.; Li, H.; Chen, T.; Chen, Z.; Zhang, Y.; Fan, X.; Zhan, L.; Ma, L.; Zhou, X. Constructing crosslinked lithium polyacrylate/polyvinyl alcohol complex binder for high performance sulfur cathode in lithium-sulfur batteries. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125870. [Google Scholar] [CrossRef]
  47. Lin, X.; Wen, Y.; Ma, D.; Li, J.; Zhu, Z.; Wang, S.; Liu, H.; Xu, X.; Huang, X. A zwitterionic polymer binder Integrating multiple dynamic interactions enables High-Performance Lithium-Sulfur batteries. Chem. Eng. J. 2025, 512, 162808. [Google Scholar] [CrossRef]
  48. Mu, P.; Zhang, S.; Zhang, H.; Li, J.; Liu, Z.; Dong, S.; Cui, G. A Spidroin-Inspired Hierarchical-Structure Binder Achieves Highly Integrated Silicon-Based Electrodes. Adv. Mater. 2023, 35, 2303312. [Google Scholar] [CrossRef]
Figure 1. (a) The hydrolysis process of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane; (b) schematic of the functionality of the BG binder: effectively adsorbing polysulfides and maintaining structural integrity during the electrochemical cycling of the S cathode.
Figure 1. (a) The hydrolysis process of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane; (b) schematic of the functionality of the BG binder: effectively adsorbing polysulfides and maintaining structural integrity during the electrochemical cycling of the S cathode.
Sustainability 17 05240 g001
Figure 2. (a) 1H NMR spectrum of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane; (b) 1H NMR spectrum of BTU; (c) FTIR of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane and BTU.
Figure 2. (a) 1H NMR spectrum of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane; (b) 1H NMR spectrum of BTU; (c) FTIR of 3,9-divinyl-2,4,8,10-tetraoxaspiro-[5.5]-undecane and BTU.
Sustainability 17 05240 g002
Figure 3. Temperature-dependent solubility of GA (a) without and (b) with BTU.
Figure 3. Temperature-dependent solubility of GA (a) without and (b) with BTU.
Sustainability 17 05240 g003
Figure 4. Electrochemical performance of sulfur cathodes with different binders. Cycling performance of different binders at 0.5 C (a), 1 C (b), and 2 C (c); GCD curves of BG (7:3) (d), BTU (e), and PVDF (f) cathodes at 0.5 C; (g) rate performance of Li-S batteries based on different cathodes; GCD curves of BG (7:3) (h), BTU (i), and PVDF (j) cathode at different rates from (g); (k) cycling performance of high-sulfur-loading BG (7:3) cathode at 0.1 C and 0.2 C; (l) GCD curves of BG (7:3) with a high-sulfur loading of 3 mg cm−2 at different cycles.
Figure 4. Electrochemical performance of sulfur cathodes with different binders. Cycling performance of different binders at 0.5 C (a), 1 C (b), and 2 C (c); GCD curves of BG (7:3) (d), BTU (e), and PVDF (f) cathodes at 0.5 C; (g) rate performance of Li-S batteries based on different cathodes; GCD curves of BG (7:3) (h), BTU (i), and PVDF (j) cathode at different rates from (g); (k) cycling performance of high-sulfur-loading BG (7:3) cathode at 0.1 C and 0.2 C; (l) GCD curves of BG (7:3) with a high-sulfur loading of 3 mg cm−2 at different cycles.
Sustainability 17 05240 g004
Figure 5. (a) CV curves of the sulfur cathode with BG (7:3) binder, (b) BTU binder, and (c) PVDF binder at scan rates ranging from 0.05 to 0.25 mV s−1; (d) shows the oxidation CV peak current versus the square root of scan rate for the three binders; (e,f) show the reduction CV peak current versus the square root of scan rate for the three binders; (gi) show the CV curves of the three binders at a scan rate of 0.1 mV s−1.
Figure 5. (a) CV curves of the sulfur cathode with BG (7:3) binder, (b) BTU binder, and (c) PVDF binder at scan rates ranging from 0.05 to 0.25 mV s−1; (d) shows the oxidation CV peak current versus the square root of scan rate for the three binders; (e,f) show the reduction CV peak current versus the square root of scan rate for the three binders; (gi) show the CV curves of the three binders at a scan rate of 0.1 mV s−1.
Sustainability 17 05240 g005
Figure 6. (a) Cyclic voltammetry (CV) curves of electrodes with different binders at a scan rate of 0.1 mV s−1; (b) Tafel plot of peak A; (c) peak C1; and (d) peak C2.
Figure 6. (a) Cyclic voltammetry (CV) curves of electrodes with different binders at a scan rate of 0.1 mV s−1; (b) Tafel plot of peak A; (c) peak C1; and (d) peak C2.
Sustainability 17 05240 g006
Figure 7. In situ electrochemical impedance spectroscopy (EIS) characterization of Li-S batteries with BG (7:3) (a) and PVDF (b) binders; (c) UV-Vis spectra of polysulfide solutions after being adsorbed for 8 h by different binder/carbon black composites; (d) S 2p XPS spectra of the Li anode after 100 cycles using BG (7:3), BTU, and PVDF binders; (e) SEM images and corresponding EDS mapping images of sulfur cathodes with BG (7:3), BTU, and PVDF binders; EDS mapping images of sulfur cathodes with BG (7:3) (f), BTU (g), and PVDF (h) binders.
Figure 7. In situ electrochemical impedance spectroscopy (EIS) characterization of Li-S batteries with BG (7:3) (a) and PVDF (b) binders; (c) UV-Vis spectra of polysulfide solutions after being adsorbed for 8 h by different binder/carbon black composites; (d) S 2p XPS spectra of the Li anode after 100 cycles using BG (7:3), BTU, and PVDF binders; (e) SEM images and corresponding EDS mapping images of sulfur cathodes with BG (7:3), BTU, and PVDF binders; EDS mapping images of sulfur cathodes with BG (7:3) (f), BTU (g), and PVDF (h) binders.
Sustainability 17 05240 g007
Table 1. Electrochemical performance of lithium–sulfur batteries with different binders.
Table 1. Electrochemical performance of lithium–sulfur batteries with different binders.
BinderCathode MaterialSulfur Loading (mg cm−2)Electrochemical PerformanceRef.
PILSuper/S3.6686.7 mAh g−1 at 0.1 C after 100 cycles[6]
HMM/PAAKetjen black/S4.2813.6 mAh g−1 at 0.1 C after 90 cycles[7]
PVA-BA0.07Acetylene–carbon black/S3.5947 mAh g−1 at 0.2 C after 300 cycles[10]
PNAVSKetjen black/S1.0647.8 mAh g−1 at 1 C after 500 cycles[14]
CPSKetjen black/S8.5627 mAh g−1 at 0.5 C after 100 cycles[16]
CSEGKetjen black/S4.1641.4 mAh g−1 at 0.2 C after 100 cycles[45]
LiPAA/PVACarbon/S1.2654.2 mAh g−1 at 0.5 C after 200 cycles[46]
PLMKetjen black/S1.0741.1 mAh g−1 at 1 C after 500 cycles[47]
BG(7:3)Ketjen black/S3.0813.2 mAh g−1 at 0.1 C after 100 cyclesThis work
BG(7:3)Ketjen black/S1.2928.9 mAh g−1 at 0.5 C after 150 cyclesThis work
Table 2. List of calculated lithium-ion diffusion coefficients.
Table 2. List of calculated lithium-ion diffusion coefficients.
Binders D Li I p A D Li I p C 1 D Li I p C 2
BG (7:3)1.60 × 10−73.13 × 10−89.74 × 10−8
BTU8.68 × 10−81.88 × 10−86.81 × 10−8
PVDF4.72 × 10−81.55 × 10−82.36 × 10−8
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

Jing, X.; Liu, S.; Wang, J.; Wan, C.; Zhu, J.; He, X.; Jin, B. Eco-Friendly Gallic Acid-Tailored Binder with Synergistic Polarity Sites for High-Loading Lithium–Sulfur Batteries. Sustainability 2025, 17, 5240. https://doi.org/10.3390/su17125240

AMA Style

Jing X, Liu S, Wang J, Wan C, Zhu J, He X, Jin B. Eco-Friendly Gallic Acid-Tailored Binder with Synergistic Polarity Sites for High-Loading Lithium–Sulfur Batteries. Sustainability. 2025; 17(12):5240. https://doi.org/10.3390/su17125240

Chicago/Turabian Style

Jing, Xulong, Shuyu Liu, Jiapei Wang, Chao Wan, Juan Zhu, Xiaojun He, and Biyu Jin. 2025. "Eco-Friendly Gallic Acid-Tailored Binder with Synergistic Polarity Sites for High-Loading Lithium–Sulfur Batteries" Sustainability 17, no. 12: 5240. https://doi.org/10.3390/su17125240

APA Style

Jing, X., Liu, S., Wang, J., Wan, C., Zhu, J., He, X., & Jin, B. (2025). Eco-Friendly Gallic Acid-Tailored Binder with Synergistic Polarity Sites for High-Loading Lithium–Sulfur Batteries. Sustainability, 17(12), 5240. https://doi.org/10.3390/su17125240

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