Van-Phu
Vu
a,
Hye-Mi
So
a,
Areum
Kim
a,
Jin Young
Lee
a,
Minsub
Oh
*a and
Seungmin
Hyun
*ab
aDepartment of Nano-Devices & Displays, Korea Institute of Machinery and Materials (KIMM), Daejeon, 34103, Republic of Korea. E-mail: odong@kimm.re.kr; hyun@kimm.re.kr
bDepartment of Nano-Mechatronics, University of Science and Technology (UST), Daejeon, 34113, Republic of Korea
First published on 8th September 2025
Polymer binders are crucial in electrodes, as they both hold together active material particles and conductive additives and firmly bond the composite to the current collector. Thus, they maintain the mechanical integrity of battery systems and stabilize electron pathways during repeated cycling. However, mechanical stresses such as bending, stretching, and volumetric changes can generate internal fractures that disrupt conductive pathways, detach active particles, and compromise electrode–collector interfaces, ultimately degrading electrochemical performance. Although conventional binders provide adequate adhesion and processability, they are inherently passive and cannot respond to such structural damage. Once cracks form or particle contact is lost, they cannot re-establish connectivity, causing irreversible capacity loss. In contrast, self-healing polymer binders (SHPBs), a new class of smart materials, can autonomously repair the mechanical and structural damage incurred during battery operation. Their unique ability to re-establish chemical or physical bonds within the polymer matrix enables them to effectively mend microcracks, preserving electrode cohesion and conductive networks. These adaptive properties offer several compelling advantages, e.g., improved mechanical resilience and extended cycle life. They also mitigate internal short circuits and potential thermal runaway, enhancing safety. Furthermore, SHPBs support consistent electrochemical performance by maintaining interfacial integrity among active materials, conductive additives, and current collectors. This reduces the need to maintain or replace batteries and/or their components, improving the cost-effectiveness and environmental sustainability of energy storage systems. In contrast to earlier reviews that focused on binders for Si-based lithium-ion batteries, this review explores recent advancements in the molecular design strategies and healing mechanisms of SHPBs, and their impact on cell-level performance across battery platforms such as lithium-ion, lithium–sulfur, and emerging sodium-based batteries. We discuss critical challenges, key future research directions, and opportunities for advancing resilient, safe, high-energy-density batteries with prolonged cycle lives.
Polyvinylidene fluoride (PVDF) is one of the most commonly used polymer binders in lithium-ion batteries (LIBs) owing to its good adhesion and stable electrochemical properties.6,7 However, PVDF mainly relies on weak van der Waals forces to retain active particles, which becomes a limitation when electrode materials undergo large physical changes during cycling, such as the expansion and contraction of Si anodes. These repeated changes often lead to cracking and a loss of contact between components.8,9 Other conventional binders, such as carboxymethyl cellulose (CMC), styrene–butadiene rubber (SBR), polyacrylic acid (PAA), and poly(ethylene oxide) (PEO), interact more effectively with polar surfaces and offer better wettability and processability.10 Nonetheless, they often lack sufficient mechanical strength and adaptability to the continuous physical stress encountered during cycling.11–13 As illustrated in Fig. 1, each conventional binder material exhibits trade-offs across key functional parameters. For instance, PVDF offers high oxidation and thermal stability but poor adhesion and low ionic conductivity, whereas PEO and PAA provide better ionic transport yet weak tensile strength and elasticity. CMC forms non-covalent bonds with surface oxygen on Si and enhances cycling stability in anodes composed of nano-sized Si, but it has poor mechanical durability and cannot adhere to freshly fractured micro-sized Si because of its insufficient surface oxygen. Therefore, it has a limited ability to buffer large volume changes during cycling.14,15 These radar plots offer a semi-quantitative visual comparison based on literature-reported data on the key attributes required for durable, high-performance binders. This comparative analysis clearly highlights that no single commercial binder fulfills all performance demands, particularly under the repeated stress conditions associated with modern electrode materials. Fig. 1 therefore provides a rationale for the development of advanced polymer binders with multifunctional capabilities.
![]() | ||
| Fig. 1 Comparative analysis of the key functional properties of commonly used polymer binders. The comparison is semi-quantitative, as the data were collected from various literature sources with differing experimental conditions and methods. Reproduced with permission from ref. 241. Copyright 2024, Royal Society of Chemistry. | ||
One of the most significant challenges in modern batteries is the mechanical degradation of electrodes during operation.16,17 This problem is especially pronounced in high-capacity electrode materials, such as Si anodes and sulfur cathodes, which undergo substantial size changes that can lead to cracking, delamination, and capacity fading.18–20 These structural failures disrupt electron pathways, reduce conductivity, and degrade battery performance over time. Additionally, these mechanical changes continuously introduce fresh surfaces, accompanied by new interactions among the binder, electrolyte, and active materials. For example, the electrolyte can penetrate cracks or the voids created when the binder separates from the active material and/or adjacent components, thereby enabling side reactions at these newly exposed surfaces. All of these phenomena can accelerate both mechanical and chemical degradation, further contributing to electrode aging.21–24
To address these challenges, self-healing polymer binders (SHPBs) have emerged as a promising way to reinforce electrode integrity and extend battery lifespan. Dynamic chemical interactions enable these materials to autonomously heal without requiring external additives or energy input.15,25,26 Unlike conventional binders, which passively maintain component cohesion, SHPBs actively respond to damage by repairing cracks and restoring disrupted interfaces via reversible chemical and/or physical bonding.27–29 This self-repair capability enables SHPBs to accommodate the repetitive mechanical stress caused by volume fluctuations during cycling, thereby preserving structural cohesion and electrical contact. By maintaining internal interfaces and conductive networks, SHPBs significantly improve cycling stability and mitigate capacity loss.30 Moreover, by sealing microcracks and maintaining close contact between active materials, SHPBs help minimize the formation of fresh reactive surfaces that would otherwise be exposed to the electrolyte. This prevents undesirable side reactions, such as the continuous formation of unstable solid electrolyte interphase (SEI) layers, and excessive lithium consumption and electrode degradation caused by electrolyte intrusion into damaged regions.31,32 These combined advantages make SHPBs especially attractive for next-generation battery technologies that demand high mechanical resilience, long-term performance, and high energy density under extreme conditions. Given the growing interest and rapid progress in this field, a systematic and application-oriented review is timely; however, most reviews thus far have focused on binders for Si-based lithium-ion batteries.
In this review, we provide a comprehensive overview of the recent progress in developing SHPBs for battery electrodes. We begin by outlining the essential functions of polymer binders and the specific challenges they face in high-capacity electrodes subjected to intense structural stress. We then highlight key molecular design strategies and bonding mechanisms that enable self-healing, supported by performance outcomes demonstrated across representative case studies in various battery technologies. Finally, we identify existing limitations and propose future research directions aimed at accelerating the practical implementation of SHPBs in next-generation energy storage systems.
Fundamentally, a typical polymer binder must meet several essential requirements. First, it should strongly adhere to the active materials, conductive additives, and the current collector, thereby binding them together to maintain electrode cohesion and mechanical robustness. Second, it must possess sufficient tensile strength to accommodate the large volume fluctuations that occur during cycling, particularly in high-capacity electrodes. Third, it must balance solubility: it should dissolve in processing solvents to fabricate electrodes, but thereafter, it must remain insoluble in the electrolyte to preserve structural and electrochemical stability during operation.
Selecting an appropriate binder thus involves a multifaceted evaluation of several criteria, including electrochemical stability, mechanical performance, thermal resistance, environmental impact, and cost-effectiveness. By carefully considering each of these factors, binders can be identified that improve efficiency, prolong cycle life, and are compatible with sustainable manufacturing practices. As binder technologies evolve, their development should continue to align with these priorities to meet the demands of future energy storage systems.
Beyond these core requirements, modern binder design is increasingly focusing on multifunctionality to address the complex challenges posed by next-generation batteries (Fig. 2). As our understanding of electrochemical processes deepens, the binder has shifted from being considered a passive support component to an active material contributing to improved performance. For instance, self-healing binders can autonomously repair microstructural damage, thus enhancing mechanical resilience and extending service life. Some advanced binders also integrate ion and electron transport functions, forming interconnected networks that accelerate redox kinetics and increase power output. In addition, electrochemical stability, especially resistance to oxidation and reduction, is critical for compatibility with high-voltage or chemically reactive electrodes. Meanwhile, intrinsically conductive polymer binders are emerging as promising alternatives to traditional insulating systems.35 By providing both structural support and conductive pathways, these binders decrease the need for conductive additives, thereby increasing the proportion of active material and boosting the energy density. Furthermore, binders with low polarity and high ionic conductivity improve the mobility of lithium ions at the electrode–electrolyte interface, which enhances rate capability. Ultimately, polymer binders are now engineered as multifunctional materials that actively influence the electrochemical behavior, stability, and longevity of battery systems. Designing such advanced binders requires a nuanced understanding of their interactions with electrode architectures, the principal active materials, and electrolytes, enabling tailored solutions for high-performance, next-generation energy storage.
![]() | ||
| Fig. 2 Conceptual diagram illustrating the ideal multifunctional properties required for advanced polymer binders in high-performance battery electrodes. | ||
C bonds47) and are commonly used in supramolecular polymeric networks.48–51 Although individually weaker, these interactions collectively allow polymer chains to realign and recover mechanical integrity. For example, Su et al.52 developed a multifunctional polymer binder named 1-phenyl-3-(2-pyridyl)-2-thiourea (PPTU) by engineering hydrogen bonds between a conductive polymer, poly(3,4-ethylenedioxythiophene)–poly(styrenesulfonate) (PEDOT:PSS), and a stretchable polymer, poly(ether-thiourea). The PPTU binder can encapsulate silicon nanoparticles (SiNPs) and form a continuous three-dimensional (3D) network structure. This network significantly improves both the mechanical stability and electronic conductivity of the electrode. Moreover, thanks to its excellent self-healing properties, PPTU effectively handles the large volume expansion of Si particles, helping preserve electrode integrity during cycling.
![]() | ||
| Fig. 3 Schematic illustration of an intrinsic self-healing system, wherein damaged regions autonomously heal through polymer matrix reorganization and the regeneration of (a) dynamic covalent and/or (b) non-covalent bonds at the fractured interfaces. Reproduced with permission from ref. 242. Copyright 2021, Wiley-VCH. (c) Illustration of representative reversible bonds and shape-memory-assisted healing mechanisms59 in intrinsic self-healing polymer networks. Reproduced with permission from ref. 59. Copyright 2024, American Chemical Society. | ||
Supramolecular interactions, such as π–π stacking53,54 and host–guest interactions,55,56 further facilitate reversible self-assembly and enhance resilience.57 Additionally, certain polymers exhibit a shape-memory effect, where external stimuli (e.g., heat, magnetism, or electricity) trigger the material to return to its original configuration, enabling crack closure and healing.58–60 This mechanism addresses a key limitation of traditional self-healing systems, which often rely on manually aligning the fractured interfaces to enable repair (Fig. 3c). For instance, Rehman et al.61 synthesized a polyurethane–polycaprolactone (PUPCL) copolymer network featuring a phase-separated structure composed of chemically cross-linked hard segments, namely, 1,6-hexamethylene diisocyanate (HDI), N,N,N′,N′-tetrakis(2-hydroxypropyl)ethylenediamine (HPED), and triethanol amine (TEA), along with thermally responsive soft segments of HDI and polycaprolactone-diol (PCL-diol). PUPCL exhibited stable shape memory behavior, triggered by heating slightly above its melting temperature, where the soft PCL segments melted to enable chain mobility, while the crosslinked network prevented complete flow or degradation. Leveraging its shape-memory-assisted self-healing capability, this system achieved a high self-healing efficiency of 95.1%. Similarly, Huang and co-workers62 developed thermoplastic vulcanizates that self-healed under thermal, magnetic, or light stimuli, with prestretched samples achieving high healing efficiencies. Together, these intrinsic mechanisms enable repeated repair cycles and enhance the mechanical integrity of electrodes over extended use.
![]() | ||
| Fig. 4 Schematic illustration of extrinsic self-healing mechanisms in polymer systems. (a) Microcapsule-based healing strategies, where the healing agent (monomer) is either encapsulated alone (single capsule), combined with a separate capsule containing a catalyst (dual capsule), or co-encapsulated within the core and shell layers of a single capsule (all-in-one configuration). Reproduced with permission from ref. 67. Copyright 2015, Elsevier. (b) Microvascular-based self-healing systems employing engineered vascular networks. These include one-dimensional (1D) channels or fibers filled with healing agents, as well as more complex two-dimensional (2D) and 3D networks that require precise fabrication to ensure continuous flow and connectivity. Reproduced with permission from ref. 68. Copyright 2017, Elsevier. | ||
Another strategy mimics the cardiovascular networks found in biological systems. These materials contain interconnected hollow channels filled with healing agents that flow toward damaged areas and polymerize in situ.72–74 A third type involves phase-separated agents physically embedded within the matrix. When cracks form, these agents migrate to the cracks and facilitate repair via physical interactions or chemical reactions.75,76 Although extrinsic systems offer rapid, effective healing, they are often single-use systems at each damage site unless the healing agent is replenished. Nonetheless, they provide versatile solutions, particularly for structural or packaging components in battery systems where periodic mechanical failure is expected.77
Another hybrid strategy integrates extrinsic reservoirs (e.g., capsules or vascular channels) into polymers that already exhibit intrinsic self-repair.82 This enables an immediate initial healing response by releasing a healing agent, followed by long-term self-repair through reversible bond reformation. These systems are especially valuable in environments prone to frequent or large-scale damage, offering both rapid and sustained healing. Despite their advantages, hybrid systems often involve more complex synthesis and higher costs, which may limit their scalability in battery applications. Nonetheless, they offer a powerful design paradigm for engineering resilient and durable binders in advanced energy storage systems.
From both practical and ecological perspectives, self-healing polymers present a compelling solution by prolonging the lifespan of batteries, reducing the frequency of replacement of batteries and/or their components, and delaying their end of life, thereby minimizing electronic waste. This section reviews applications of SHPBs across different battery chemistries, including lithium-ion, Li–S, and emerging sodium-based systems. Representative studies are highlighted to demonstrate how molecular binder design and healing mechanisms directly affect electrochemical performance, structural integrity, and long-term reliability in each context.
![]() | ||
| Fig. 5 Representative failure mechanisms in Si anodes, highlighting structural breakdown and SEI instability during cycling. Reproduced with permission from ref. 243. Copyright 2018, Royal Society of Chemistry. | ||
| Self-healing polymer binder | T g (°C) | Healing mechanism | Healing conditions | Healing efficiency (%) | Si mass loading (mg cm−2) | ICE (%) | Highest rate performance (mA h g−1 @ C-rate) | Cycling stability (% (cycles @ C-rate)) | Ref. |
|---|---|---|---|---|---|---|---|---|---|
| SHP/PEG | N/A | Hydrogen bonds | RT, 3 h | N/A | 0.5–0.7 | 83.0 | 1300 @ 0.5C | 80% (150 @ 0.5C) | 101 |
| PAA–UPy | N/A | Hydrogen bonds | RT, <1 min | N/A | 0.4–0.6 | 86.4 | 2662 @ 5C | 74% (110 @ 1C) | 102 |
| PAA–P(HEA-co-DMA) | 83 | Hydrogen bonds, catechol groups | N/A | N/A | ∼1.0 | 89.3 | 2550 @ 1 A g−1 | 93.8% (220 @ 1 A g−1) | 103 |
| UPy–PEG–UPy | N/A | Hydrogen bonds | RT, 3 h | N/A | N/A | 81.0 | 1847 | 78.7% (400) | 104 |
| Alg–C–CS | N/A | Hydrogen bonds | N/A | N/A | ∼0.56 | 74.5 | N/A | 750 mA h g−1 (100) | 105 |
| Fe–PDBP | 12.8 | Coordination bonds | RT, 24 h | N/A | 0.7 | 81.9 | 1.52 mA h cm−2 @ 1C | 81.9% (350 @ 1C) | 106 |
| BC10–g | N/A | Hydrogen bonds, borate ester bonds | RT, 3 h with electrolyte | N/A | 0.25–2.1 | 82.7 | 2750 @ 0.2C | 87.3% (100 @ 0.2C) | 107 |
| PAU-g-PEG | N/A | Hydrogen bonds | RT, 24 h | N/A | 0.5–1.0 | 70.4 | 2500 @ 3C | 53.3% (350 @ 0.5C) | 108 |
| Poly(ether-thioureas) | N/A | Hydrogen bonds | 4 h | N/A | 1.2 | N/A | 1917 @ 4.2 A g−1 | 85.6% (250 @ 0.42 A g−1) | 109 |
| CA–PAA | −40 | Hydrogen bonds | RT | N/A | 0.6–1.8 | 89.5 | 4000 @ 1C | 90% (300 @ 0.1C) | 110 |
| GCS/DA–PEG | N/A | Imine bonds | 40 °C, 5 min | 85.7 | ∼0.5 | 82.2 | 2716 @ 3C | 64.7% (150 @ 0.5C) | 111 |
| GCS-I-OSA-10 | N/A | Schiff base | RT | N/A | 0.35–0.45 | 72.7 | 2185 @ 5C | 1355 mA h g−1 (200 @ 1C) | 112 |
| Starch/PVA/ST | N/A | Hydrogen bonds, borate ester bonds | RT, 3 min | N/A | 0.3–0.5 | N/A | 1283 @ 4 A g−1 | 41.8% (300 @ 1 A g−1) | 113 |
| PVA/LB-30 | N/A | Borate ester bonds | N/A | N/A | 0.6–0.7 | 74.2 | 2211.5 @ 400 mA g−1 | 79.9% (180 @ 400 mA g−1) | 114 |
| DA-grafted PAA | N/A | Ionic bonds, hydrogen bonds | RT | N/A | 1 | 83.3 | 2671.6 @ 1C | 68.8% (100 @ 0.4 A g−1) | 115 |
| PAA–TUEG | N/A | Hydrogen bonds | RT, 2 h | 81 | 0.5–1 | 87.2 | 1000 @ 2C | 82% (300 @ 0.5C) | 116 |
| P-BIAN/PAA | N/A | Electrostatic hydrogen bonds | N/A | N/A | ∼0.87 | 87.5 | 2100 @ 500 mA g−1 | ∼95% (600 @ 500 mA g−1) | 117 |
| PAA–DA/PVA | N/A | Hydrogen bonds | N/A | N/A | N/A | N/A | 2168.7 @ 4 A g−1 | 50.8% (500 @ 4 A g−1) | 118 |
| PAA–PEO | N/A | Hydrogen bonds | RT, 12 h | 95.6 | 0.8–1.5 | 90.3 | 1567 @ 1C | 42% (200 @ 0.5C) | 119 |
| Al/Alg-TUEG | N/A | Metal coordination, hydrogen bonds | RT, 2 h | 90 | N/A | 87.2 | 250 @ 5C | 77.4% (300 @ 0.5C) | 120 |
| PEDOT:PAA:PA | N/A | Hydrogen bonds, electrostatic interactions | 4 h with electrolyte | 81.9 | ∼1.2 | 94.0 | 2084 @ 5C | 74% (250 @ 0.5C) | 121 |
| PAA–UPy 5%/PEO | 19.2 | Hydrogen bonds | RT, 12 h | 96.2 | 0.8–1.5 | 92.3 | 1855 @ 1C | 51.5% (200 @ 0.5C) | 122 |
| Guar gum–CA | N/A | Hydrogen bonds, coordination bonds | RT | N/A | 0.6–1.4 | 93.0 | 2416 @ 4 A g−1 | 1184 mA h g−1 (740 @ 2 A g−1) | 123 |
| PLSA75 | 94.2 | Electrostatic interactions, hydrogen bonds | RT, 30 min | ∼93 | 0.4–0.6 | 85.3 | 811 @ 5C | 82.1% (300 @ 0.5C) | 124 |
| PAA–DABBF | N/A | Dynamic carbon radicals | RT, a few minutes | N/A | 0.45–0.5 | 77.9 | 1831.1 @ 5C | 1774.45 mA h g−1 (500 @ 0.5C) | 125 |
| PVA@LB | 131.3 | Hydrogen bonds | N/A | N/A | 0.7–0.9 | 6.49 | 1133 @ 6 A g−1 | 87.8% (250 @ 325 mA g−1) | 126 |
| PAA–LS | 100.7 | Coordination bonds | A few seconds | N/A | 1.5–2.0 | 63.9 | 2087.2 @ 0.5 A g−1 | 997.3 mA h g−1 (450 @ 0.5 A g−1) | 127 |
| TA-c-PAA | 59.71 | Hydrogen bonds | RT, 30 min | N/A | ∼0.4 | 83.7 | 1599 @ 2C | 1742 mA h g−1 (450 @ 0.25C) | 128 |
| LiCB | N/A | Boronic ester bonds | RT, 72 h | N/A | 1.1 ± 0.05 | 82.2 | 220 @ 4C | 600 mA h g−1 (200 @ 1C) | 129 |
| GGC | 219.5 | Hydrogen bonds | RT, interval | N/A | 1.0–1.2 | N/A | 2062.1 @ 2 A g−1 | 96% (1200 @ 1 A g−1) | 130 |
| xPAA–B–DA | N/A | Boronic ester bonds | N/A | N/A | 0.5–0.7 | 74.9 | 2259.6 @ 5C | 70.8% (100 @ 0.5C) | 131 |
| PAA–CA–PEI | −34 | Amide bonds, hydrogen bonds | RT, 30 min | N/A | 0.6–1.9 | 89.7 | 1928.4 @ 1C | 74.9% (500 @ 0.2C) | 132 |
| PDA–Al–BFPU | N/A | Metal-ion coordination, hydrogen bonds, S–S bonds | 30 °C, 1 h | 90 | 1.2 | 84.5 | 3631.8 @ 0.1C | 76.8% (200 @ 0.3C) | 133 |
| PAA–TUEG | N/A | Hydrogen bonds, ionic bonds | RT, 2 h | 80 | N/A | N/A | 1249 @ 2C | 793 mA h g−1 (180 @ 1C) | 134 |
| PAA–TP | N/A | Hydrogen bonds | N/A | N/A | 1.0–1.5 | 86.7 | N/A | 2072 mA h g−1 (300 @ 0.5C) | 135 |
| SA–STB–GN | 59 | Boronic ester bonds | RT, 48 min | N/A | 0.8–1.0 | 90.3 | N/A | 81.2% (150 @ 0.2C) | 136 |
| BDSA–DPA–PEGCE | N/A | Supramolecular interactions, S–S | RT, 2 h | N/A | ∼1.2 | 85.2 | 729 @ 2C | 81.2% (150 @ 1C) | 137 |
| PVA–4FBA–PEI | N/A | C N, B–O–C |
RT, 2 min | N/A | ∼0.38 | N/A | 954.7 @ 1C | 82.1% (200 @ 0.2C) | 138 |
| PCL12-IU | −35.3 | Hydrogen bonds | RT, 20 h | N/A | ∼1.2 | N/A | 287 @ 7C | 71% (200 @ 2C) | 139 |
| LiPAA–TA–SS | N/A | Hierarchical hydrogen bonds | RT, ∼10 min | N/A | 0.4–1.0 | 91.0 | 1678 @ 2C | 77.9% (400 @ 1C) | 140 |
| SA@borax | N/A | Boronate ester bonds | 25 °C, 2 h | N/A | N/A | 70.9 | N/A | 1655.8 mA h g−1 (500 @ 0.5 A g−1) | 141 |
| PDB | N/A | Hydrogen bonds | RT | N/A | 0.9 | 86.2 | 1973 @ 8.0 A g−1 | 2538 mA h g−1 (200 @ 0.5 A g−1) | 142 |
| β-CD-VI | N/A | Hydrogen bonds, host–guest interactions | 25 °C, 3 min | N/A | 1.3 | 98.1 | N/A | 91% (100 @ 0.2C) | 143 |
| TA–IA | N/A | Hydrogen bonds | RT | N/A | 0.7–1.2 | 66.7 | 756.1 @ 5 A g−1 | ∼88.4% (400 @ 2 A g−1) | 144 |
| CS–CA | N/A | Hydrogen bonds | RT, 10 min | N/A | 1.4–1.6 | 74.1 | 558.9 @ 5 A g−1 | ∼99.8 (400 @ 2 A g−1) | 145 |
| LA–MMA | N/A | S–S bonds, hydrogen bonds | RT, 30 s | N/A | 2.0–2.5 | N/A | 158.1 @ 0.1C | 50.29% (2000 @ 2C) | 146 |
| LiPAA/3D-s-PU | −38.8 | S–S bonds | RT, 5 h | 73.2 | 13 | 82.4 | 600 @ 2C | 74.38% (150 @ 0.5C) | 147 |
| SHIR-A1 | −62.6 | Hydrogen bonds, coordination bonds | RT | N/A | 1.52 (Si/C); 1.5 (SiOx) | N/A | 220 @ 2C | 83.3% and 86.8% (300 @ 0.5C) | 148 |
One of the earliest strategies was reported by Larcher et al.,98 who emphasized that effective Si binders must maintain continuous contact between active particles. They also highlighted that beyond elasticity, the chemical affinity between the binder and active surfaces is crucial for ensuring long-term stability. Since then, several polymer binders have demonstrated superior self-healing behavior compared with conventional options such as PVDF. For example, Mi et al.99 introduced a composite binder composed of SiNPs, sodium alginate (SA), and polyaniline (PANI), synthesized via in situ polymerization (Fig. 6a). Dynamic hydrogen bonding within the resulting Si–SA–PANI system allowed the binder to adapt to Si volume changes, delivering a high specific capacity of 1099.5 mA h g−1 after 200 cycles at 1 A g−1 and an ICE of 77.6%. Song et al.100 developed a bilayer binder system with a gradient in its mechanical properties, wherein a stiff PAA inner layer offered structural support, and a soft, self-healing bifunctional polyurethane (BFPU) outer layer absorbed stress and repaired damage. The BFPU exhibited elastomeric behavior with a low Tg of −42.5 °C; macroscopic scratches healed rapidly within 1 h, achieving a visible recovery efficiency of 96.7%.
![]() | ||
| Fig. 6 (a) Schematic illustration of the fabrication route and self-healing processes for Si–SA–PANI composites. Reproduced with permission from ref. 99. Copyright 2018, Elsevier. (b) Chemical structure of the PAA–UPy supramolecular polymer and UPy–UPy dimers. Optical images show the self-healing behavior of the PAA–UPy binder. Reproduced with permission from ref. 102. Copyright 2018, Wiley-VCH. (c) Schematic illustrations of lithiation/delithiation in Si anodes using conventional and SHPET binders, along with the synthesis pathway of SHPET and the hydrogen-bonding modes of thiourea and urea units, including trans/trans and cis/trans conformations. Reproduced with permission from ref. 109. Copyright 2020, Elsevier. (d) Structural formulae and graphical representations of Alg and β-CD polymer binders. The side chains (R) in the β-CD polymer may include 2,3-dihydroxypropyl groups in monomeric, dimeric, or trimeric forms, with glyceryl moieties as the connecting bridges in the polymer backbone. Reproduced with permission from ref. 152. Copyright 2014, the American Chemical Society. (e) Schematic design of the electrode illustrating the spatial distribution of SHP/CB within the Si layer and the corresponding cycling stability of thick Si electrodes fabricated by different methods. Reproduced with permission from ref. 159. Copyright 2015, Wiley-VCH. (f) Schematic illustration of the synthesis pathway and proposed mechanism of a 3D conductive polymer binder, along with the surface morphologies of the cathodes with self-healing CPB (above) and PVDF (below) after 100 cycles. Reproduced with permission from ref. 161. Copyright 2018, Elsevier. | ||
Tensile testing confirmed a healing efficiency of 93.1%, based on a recovered tensile strength of 5.9 MPa and 700% elongation, compared with 6.34 MPa and 800%, respectively, in the pristine state. This “rigid-to-soft” design helped mitigate lithiation-induced stress and maintain electrode integrity. As a result, the Si anode achieved an ICE of 89% and retained 97% of its capacity after 100 cycles at 1.2 A g−1. At a higher areal capacity of 2.7 mA h cm−2, the capacity retention was 88% after 200 cycles at 2 A g−1.
Ureido-pyrimidinone (UPy) motifs, known for forming stable quadruple hydrogen bonds, have emerged as promising polymeric building blocks for rapid, reversible self-healing.149,150 Inspired by this property, Zhang et al.102 incorporated UPy units into linear PAA, forming a supramolecular network with strong mechanical and healing capabilities (Fig. 6b). This network efficiently repaired microcracks and suppressed fracture propagation, improving Si anode longevity. Wang et al.151 further advanced this concept with a randomly branched hydrogen-bonded polymer binder, forming a stretchable coating that healed both small and large cracks within 5 h. The system retained 80% of its initial capacity after 90 cycles at 0.4 A g−1 with 50% Si loading, corresponding to an areal capacity of ∼1 mA h cm−2, although this is still below the commercial target for LIBs (3 mA h cm−2). Chen and co-workers109 developed a self-healing poly(ether-thioureas) (SHPET) binder that combined dynamic hydrogen bonding with cross-linked thiourea units to balance flexibility and mechanical strength (Fig. 6c). The polymer exhibited ultrafast self-healing, completely recovering in just 1 min under light pressure and within 4 h without an external force. The thiourea groups also enhanced adhesion to the Si surface, effectively maintaining electrode integrity during cycling. As a result, the Si@SHPET anode delivered a high capacity of 3744 mA h g−1 at 0.42 A g−1 and retained 85.6% of its capacity after 250 cycles at 4.2 A g−1, demonstrating the potential of this design strategy for high-performance Si anodes.
To enhance structural resilience, Choi et al.152 introduced β-cyclodextrin (β-CD)-polymer-based hyperbranched binders that leveraged hydrogen bonding to stabilize the microstructure and enable self-healing (Fig. 6d). In a follow-up study, the same research group investigated supramolecular crosslinked binders based on highly branched α-, β-, and γ-cyclodextrin polymers, exploiting dynamic host–guest interactions. In this system, each CD polymer functioned as the host, providing multiple –OH groups for hydrogen bonding with Si, while a dendritic adamantane derivative (6AD), containing six terminal guest units, served as the guest molecule. The assembly efficiency and network stability were influenced by the cavity size of the cyclodextrins and the 6AD content (ranging from 0 to 25.5 wt%). Among the tested systems, the β-CD/6AD binder demonstrated the most stable cycling performance, effectively preserving adhesion and electrode integrity by mitigating mechanical stress during repeated cycling. Furthermore, owing to its large ring structure and abundant hydroxyl groups, β-CD has also been employed as a building block in triblock copolymers in solid-state electrolytes, where it enhances their mechanical strength, ionic conductivity, and interfacial stability, all functions analogous to its role in binder design.153–158
To meet commercial areal capacity standards (≥3 mA h cm−2), Chen et al.159 proposed a 3D spatial distribution strategy for self-healing polymers (SHPs) within the entire electrode architecture, coupled with controlled Si particle size optimization (Fig. 6e). Unlike previous approaches that applied SHPs merely as a thin surface coating on Si particles, this study involved repeatedly blade coating the SHP/carbon black (CB) composite onto Si-coated Cu current collectors under mild heating conditions. This method allowed the SHP/CB composite to uniformly infiltrate the electrode throughout its entire thickness, creating a 3D network that reinforced the mechanical integrity of the structure. This study also emphasized the critical influence of Si particle size on both self-healing behavior and electrochemical performance. Among the particles evaluated, Si particles with an average diameter of ∼0.8 μm offered the most favorable balance of high coulombic efficiency, low material cost, and long-term cycling stability, with 80% capacity retention after 500 cycles. This favorable trade-off arises from the fact that larger particles are more cost-effective and scalable but prone to severe pulverization and contact loss during cycling. On the other hand, smaller particles improve interfacial contact and self-healing efficiency but are expensive to produce and difficult to scale. The ∼0.8 μm particles provided an optimal compromise between these competing factors. As a result of the synergistic effects of the 3D distribution of the SHP/CB composite and the optimized particle size, the Si–SHP/CB electrode exhibited exceptional cycling stability even at high areal capacities (3–4 mA h cm−2), approaching the performance benchmarks necessary for practical LIBs.
Building on this, Yuan et al.160 designed a crosslinked, lithiated self-healing ionomer binder tailored for Si/Gr anodes. The ionomer featured a 3D network structure and a layered nanospherical morphology, which collectively enhanced adhesion and mechanical cohesion. Through hydrogen bonding with active particles and lithium-containing chains that facilitated Li+ transport, the binder effectively maintained the electrode architecture during cycling. Impressively, the electrode thickness increased by only ∼30%, and the performance remained stable at a commercial-level areal capacity of 3 mA h cm−2 using only 1.8 wt% binder. Furthermore, Ma et al.161 synthesized a 3D conductive polymer binder with a hierarchical walnut-kernel-like structure via emulsion polymerization (Fig. 6f). This conductive binder formed multidimensional interactions with LiFePO4 particles, enhancing both adhesion and ionic conductivity. The 3D network exhibited self-healing, recovering from mechanical disruption during cycling, thereby more effectively maintaining the structural stability and cycling performance compared with conventional binders such as PVDF. Similarly, a 3D crosslinked polymer binder for SiOx anodes was developed through ionic bonding between chitosan (CS) and ethylenediaminetetraacetic acid (EDTA).162 The dynamic ionic interactions (–COO−⋯NH3+) and hydrogen bonding in the CS–EDTA binder established a flexible yet robust network that improved adhesion and prevented electrode delamination. This reversible structure enabled self-repair at fracture sites, enhancing mechanical resilience and long-term cycling stability. The SiOx@CS–EDTA electrode exhibited a reversible capacity of 721 mA h g−1 with 78% retention after 200 cycles at 1.0 A g−1, and the corresponding SiOx@CS–EDTA//NCM622 full cell delivered a high energy density of 402 W h kg−1.
In addition, environmentally friendly conductive self-healing hydrogel binders are also being explored for electrode repair mechanisms because of their low cost and high safety.163,164 Yu and co-workers165 advanced this concept by developing a conductive hydrogel binder termed ESVCA with a 3D interconnected electron-conducting network. They synthesized this binder by initiating the crosslinking of PEDOT:PSS in a solution containing polyvinyl alcohol (PVA) and 4-carboxybenzaldehyde (CBA) using ammonium persulfate (APS). In ESVCA, PEDOT:PSS provided electronic conductivity, while the hydroxyl groups in PVA facilitated Li+ transport and dynamic hydrogen bonding with the PSS sulfonate groups. This binder rapidly self-healed, as evidenced by the immediate restoration of LED illumination after being cut and manually rejoined. It also maintained electrical conductivity under 300% strain, after which it partially recovered its mechanical properties, confirming both conductive and structural healing. The resulting Si–ESVCA electrode achieved a capacity of 1786 mA h g−1 at 0.5 A g−1 and retained 71.3% of its capacity after 200 cycles at RT.
The variety of dynamic noncovalent interactions has led to multiple strategies for developing innovative SHPBs for LIBs capable of accommodating large volume changes. For instance, Fe3+–tris(catechol) complexes were used to construct dynamically crosslinked acrylate copolymers with pendant dopamine groups, yielding Si anodes with enhanced self-repair capability and improved electrochemical performance (Fig. 7a).106 The resulting binder demonstrated autonomous healing behavior, with a crack gradually closing over 24 h at RT without external stimuli, highlighting its intrinsic ability to restore structural integrity over time. In parallel, metal–ligand coordination has effectively enhanced binder-to-active particle and binder-to-binder interactions while enabling autonomous damage repair. For example, Kim et al.166 developed a CMC-based binder by in situ crosslinking CMC with poly(ethylene glycol)diglycidyl ether (PEGDE), 1-(3-aminopropyl)imidazole (Im), and Zn2+ ions during electrode drying (Fig. 7b). Their method uniformly distributed the binder throughout the electrode matrix. The incorporation of reversible Zn2+–imidazole coordination bonds introduced dynamic crosslinking that not only facilitated self-healing but also improved mechanical cohesion. Although these dynamic interactions slightly restricted polymer chain mobility, as reflected by an increase in the Tg from −35.9 °C (PEGDE–Im) to −30.2 °C (PEGDE–Im–Zn2+), the polymer remained sufficiently flexible for healing and stress dissipation. The resulting CMC–PEGDE–Im–Zn2+ binder exhibited high elasticity, promoting tight particle contact and stabilizing the interfacial structure throughout cycling, which supported consistent and efficient ion transport. The reversible Zn2+–Im coordination bonds further contributed to the structural integrity and long cycle life. In full-cell tests, this system achieved a high areal capacity of 3 mA h cm−2 after 120 cycles at 0.5C (1C = 0.65 A g−1) (Fig. 7c).
![]() | ||
| Fig. 7 (a) Schematic illustration of the self-healing mechanism involving Fe3+–catechol coordination within a binder network for Si anodes (left), and the molecular structure of poly(dopamine methacrylamide (DMA)-co-butyl acrylate (BA)-co-polyethylene glycol diacrylate (PEGDA)), termed PDBP binder (right). Chemical stability of the binder against the electrolyte and the long-term cycling performance of Si electrodes based on Fe–PDBP@pH10 measured over five months. The self-healing capability of the binder was visualized on the macroscale before and after 24 h. Reproduced with permission from ref. 106. Copyright 2019, the American Chemical Society. (b) Schematic illustration of the supramolecular network formed on the Si/C composite via in situ crosslinking between PEGDE–Im–Zn2+ and CMC, along with the process for preparing the metal–ligand complex PEGDE–Im–Zn2+.166 (c) Rate capability at various C-rates and cycling performance of the full cell at 0.5C with an LiNi0.8Co0.15Al0.05O2 (NCA) loading of 26.83 mg cm−2. Reproduced with permission from ref. 166. Copyright 2021, Wiley-VCH. | ||
Xu and colleagues103 employed a “spring-expander” concept to design a binder with soft and hard domains by copolymerizing PAA with poly(2-hydroxyethyl acrylate-co-dopamine methacrylate) (P(HEA-co-DMA)) incorporating hydroxyl and catechol functionalities from dopamine side chains (Fig. 8a). The resulting multi-network polymer, formed in situ during electrode fabrication, combined mechanical toughness with flexibility through hydrogen bonding. The polymer exhibited a Tg of 83 °C and intrinsic self-healing ability enabled by reversible hydrogen bonds, which enhanced its structural resilience under repeated cycling (Fig. 8b). As a result, this binder achieved excellent electrochemical performance in Si microparticle (SiMP) anodes, with a capacity of 2394 mA h g−1 after 220 cycles at 1 A g−1 and 93.8% capacity retention (Fig. 8c). Building on this, a similar Ni2+ coordination strategy was employed to construct a 3D binder network integrating soft and hard domains and self-recoverable features.167 In this system, flexible gellan gum chains coordinated with Ni2+ were interlaced with linear PVA chains (denoted as PGN), effectively buffering Si volume expansion during cycling. The resulting PGN-5 binder demonstrated rapid self-healing, returning to its original shape after repeated stretching, indicative of elastic, spring-like healing behavior.
![]() | ||
| Fig. 8 (a) Chemical structures of poly(2-hydroxyethyl acrylate-co-dopamine methacrylate), termed P(HEA-co-DMA), and their interactions with Si. (b) The photographs show the self-healing and elastic resilience tests of the binder. (c) Rate capability at various specific current densities and cycling performance of the electrodes with different binders. Reproduced with permission from ref. 103. Copyright 2018, Elsevier. (d) Chemical structure and intermolecular interactions of the hard and soft phases in the BDP SHPB, enabling enhanced mechanical properties, reversible self-healing ability, and high ionic conductivity for Si-based LIBs. (e) Self-healing ability and elastic resilience of the BDP polymer after cutting, healing, and stretching tests. (f) Cycling performance of the NCM523//Si/C@BDP full cell. Reproduced with permission from ref. 137. Copyright 2024, Elsevier. (g) Schematic of the boronic ester binder illustrating covalent crosslinking and self-healing via hydroxyl–PEO/boronic ester interactions, and fast Li+ transport facilitated by PEO chains. (h) Self-healing capability of the BC10–g polymer film demonstrated by its re-formation into a cohesive structure after cutting and electrolyte-assisted healing, capable of supporting an external weight. Reproduced with permission from ref. 107. Copyright 2020, Wiley-VCH. | ||
Zhang et al.137 also developed a multifunctional binder composed of (1,1′-biphenyl)-4,4′-diamino-2,2′-disulfonic acid (BDSA), 3,3′-dithiodipropionic acid (DPA), and poly(ethylene glycol)bis(carboxymethyl)ether (PEGCE), termed BDP, featuring distinct hard and soft phases to further enhance Si/C anode performance. The hard phase, based on S–S-bonded DPA and BDSA, contributed mechanical robustness, self-healing capability, and stress resistance through S–S dynamic covalent bonds and supramolecular interactions (Fig. 8d). The soft PEGCE-based phase increased flexibility and facilitated Li+ transport via ether oxygen coordination.168–170 The BDP polymer exhibited autonomous self-healing behavior, rejoining cleanly after being cut and recovering stretchability to nearly 300% within 2 h at RT (Fig. 8e). Pre-lithiation of BDSA helped mitigate initial lithium loss. The BDP binder effectively mitigated swelling, stabilized the SEI, and delivered an ICE of 85.2% with excellent cycling and rate performance in both half and full cells (NCM523//Si/C) (Fig. 8f).
A natural-polymer-based binder strategy was introduced using an ionically conductive boronic crosslinker (BC) containing pendant phenylboronic acid groups.107 The BC formed dynamic boronic ester bonds with vicinal diols in guar gum (BC–g) via spontaneous RT reactions (Fig. 8g), eliminating the need for external stimuli. This dynamic network provided mechanical flexibility, self-healing ability, and enhanced ionic conductivity owing to the PEO segments. Although conventional scratch–recovery tests were unsuitable because of the rigid covalent network of the polymer, a stacked film adhesion test under electrolyte-soaked conditions demonstrated supramolecular self-healing behavior driven by hydrogen bonding and dynamic boronic ester exchange (Fig. 8h). When applied in high-loading (>2 mg cm−2) SiNP electrodes, BC–g maintained cohesion and suppressed structural degradation, significantly extending cycle life. Thermally reversible dynamic networks have also been explored for their self-healing potential.171 Using DA click chemistry, 1,6-bismaleimide was reacted with furfurylamine-functionalized PAA to form a 3D crosslinked binder (DA–PAA).172 The DA reaction endowed the binder with thermal reversibility and strong adhesion to Si, resulting in excellent mechanical resilience and healing capacity at RT after a few hours. Thus, Si anodes with DA–PAA achieved a reversible capacity of 1076 mA h g−1 after 200 cycles at 0.5C, demonstrating that this design effectively stabilized high-capacity electrodes.
To mitigate this deterioration, researchers have developed polymer binders incorporating polar functional groups (–NH2, –COOH, and –OH) or cationic moieties (e.g., quaternary ammonium) from commercial polymers, natural materials, or water-soluble precursors.180–185 These groups or moieties interact with LiPS via electrostatic interactions, hydrogen bonding, or coordination chemistry, lengthening the cycle life by immobilizing the soluble intermediates.186 However, many of these binders lack sufficient structural robustness to accommodate repetitive volume changes or endure mechanical damage during processing. Table 2 summarizes recent advances in SHPBs specifically tailored for Li–S cathodes. This table highlights key features such as the polymer composition, self-healing strategies and efficiencies, electrolyte-to-sulfur (E/S) ratios, and cycling performance. Many SHPBs incorporate dynamic bonding networks that simultaneously facilitate crack repair and immobilize LiPS, enabling stable cycling even under practical conditions (e.g., with sulfur loadings of >6 mg cm−2 and using a lean electrolyte). Additionally, Table S2 provides detailed information on these systems, including the electrode architecture, mechanical properties, electrolyte composition, and specific electrochemical results.
| Self-healing polymer binder | T g (°C) | Healing mechanism | Healing conditions | Healing efficiency (%) | S content (wt%) | E/S ratio (μL mg−1) | High S loading (mg cm−2) | Cycling stability on areal capacity (mA h cm−2 (cycles @ C-rate)) | Ref. |
|---|---|---|---|---|---|---|---|---|---|
| PDEM–CT | N/A | Hydrogen bonds | RT, 60 min | 86.7 | 60 | N/A | 7.6 | N/A | 188 |
| SHM | N/A | Hydrogen bonds | N/A | N/A | 60 | N/A | 2.65 | N/A | 189 |
| 2S–PDMS | −50 | S–S bonds | RT, 8 h | 82.5 | 60 | >12 | 6.6 | 4.54 (50 cycles @ 0.42 mA cm−2) | 190 |
| c-LiPAACA–LAPONITE® | N/A | Electrostatic interactions and hydrogen bonds | N/A | N/A | 70 | N/A | 7.6 | 3.1 (100 cycles @ 2 mA cm−2) | 191 |
| SPI–PAM | N/A | Hydrogen bonds | RT, 3 min | N/A | 66.7% | N/A | 2.3 | N/A | 192 |
| PSPEG | N/A | S–S bonds | RT, 6 h | 97.6 | 91.2 | 4.5 | 7.6 | 6.47 (60 cycles @ 100 mA g−1) | 193 |
| SP–PA | N/A | Hydrogen bonds | N/A | N/A | 70 | 10 | 1 | N/A | 194 |
| A6D4-PA | 29.47 | Hydrogen bonds and π–π interactions | RT, 5 h with electrolyte | 86.13 | 69.6 | 10 | 4.0 | 3.0 (100 cycles @ 0.2C) | 195 |
| SHPU-400 | −10.02 | S–S bonds and hydrogen bonds | RT, 40 min | 95.5 | 60 | 30 | 8.6 | 8.0 (50 cycles @ 0.2C) | 196 |
| PVP–PEI | N/A | Hydrogen bonds | RT, 24 h | N/A | 70 | 8 | 7.1 | 6.6 (30 cycles @ 0.6 mA cm−2) | 197 |
| ZIP | −61.5 and 7.8 | Hydrogen bonds and electrostatic interactions | RT, 24 h with electrolyte | 98.2 | 70 | 10 | 8.5 | 6.6 (50 cycles @ 0.2C) | 198 |
| TDI–PPG600–DTA–HCCP | N/A | Hydrogen bonds and S–S bonds | RT, 60 min | 91.9 | 60 | 6 | 4.72 | 5.25 (100 cycles @ 0.1C) | 199 |
| RCB | N/A | Hydrogen bonds and ionic bonds | RT, 48 h | N/A | 80 | 12 | 7.0 | 6.2 (50 cycles @ 1C) | 200 |
| 10%-PVBST | N/A | Dynamic inter-chain associations | RT, 1.5 h | >80 | 70 | 7 | 6.4 | 5.7 (30 cycles @ 0.1C) | 201 |
| Cationic Im-grafted PMMA | N/A | Hydrogen bonds | RT, 3 h with electrolyte | N/A | 75 | 7 | 9.27 | N/A | 202 |
| LA–GA | N/A | S–S bonds | RT, 6 h | 95.13 | 70 | 12 | 6.5 | 5.5 (75 cycles @ 0.1C) | 203 |
| Zwitterionic LA–MPC | N/A | Hydrogen bonds, S–S bonds, and electrostatic interactions | RT, 6 h | 94.32% | 70 | 8.5 | 7.42 | 5.76 (150 cycles @ 0.1C) | 204 |
To overcome the abovementioned hurdles, Xie et al.187 designed a multifunctional self-healing binder by crosslinking PAA with cationic hydroxypropyl polyrotaxane (HPRN+) through dynamic boronic ester bonds, forming a PAA–B–HPRN+ network (Fig. 9a–c).
![]() | ||
| Fig. 9 (a) Preparation of cationic HPRN+–S from polyrotaxane (PR). (b) Synthesis of boronic-acid-modified PAA (PAA–B). (c) Formation of the crosslinked polymer binder PAA–B–HPRN+via dynamic boronic ester bonding between PAA–B and HPRN+–S. (d) Self-healing behavior of the binder after treatment at 60 °C for 6 h (left) and at 40 °C with one drop of DME/DOL electrolyte (right) applied to the scratched surface. (e) Healing performance evaluated at different healing times and temperatures. (f) Cycling performance of S cathodes measured at 1C, demonstrating prolonged stability. Reproduced with permission from ref. 187. Copyright 2022, Springer Nature. | ||
This system integrated (1) polyrotaxane for elasticity and volume buffering via a pulley-like sliding mechanism, (2) dynamic boronic ester linkages for reversible crack healing, and (3) functional groups (carboxyl, hydroxyl, and quaternary ammonium) that chemically anchored LiPS. Mechanical tests confirmed that the optimized formulation had improved tensile strength (∼1.26 MPa) and high elongation at break (425–502%), while its lap shear strength (∼0.99 MPa) was comparable to that of pristine PAA and far superior to that of PVDF. Their binder also exhibited time- and temperature-dependent self-healing behavior, with increasing recovery efficiency up to 48 h or 80 °C (Fig. 9d and e), attributed to reversible supramolecular and covalent interactions. It was also water-processable, thus avoiding harmful solvents. The resulting cathode demonstrated outstanding cycling stability, exhibiting only 0.064% capacity fading per cycle over 550 cycles at 1C, with ∼99% coulombic efficiency and strong areal capacity retention at sulfur loadings exceeding 11.5 mg cm−2 (Fig. 9f).
In addition to suffering damage during cycling and thus requiring mechanical healing, high-capacity sulfur cathodes experience complex structural and chemical evolution during multi-electron redox cycling, making it essential to stabilize the phase transitions between S/Li2S and LiPS. One emerging strategy involves the use of sulfur-containing polymers with dynamic S–S bonds. These materials not only promote structural repair but also mediate redox conversion. Organo-polysulfides and LiPS have been shown to facilitate the uniform redeposition of solid S/Li2S, restoring the cathode structure.205–209 For instance, Zuo et al.190 synthesized a poly(dimethylsiloxane) (PDMS)-based binder with reversible S–S bonds, termed 2S–PDMS (Fig. 10a and b), which enabled both self-healing and improved polysulfide conversion. The polymer showed remarkable flexibility (up to 2280% stretchability) and a low Tg of −50 °C, indicating high chain mobility at RT. It rapidly recovered 82.5% of its tensile strength within 30 min and maintained electrical conductivity after damage, demonstrating both mechanical and electrical self-healing (Fig. 10c). In Li–S cells, 2S–PDMS enhanced cathode stability, retaining 545.7 mA h g−1 after 400 cycles at 2C, with 79.6% capacity retention. In pouch cells with a sulfur loading of 6.6 mg cm−2, it delivered 4.54 mA h cm−2 after 50 cycles (Fig. 10d). The batteries also remained stable under continuous bending (Fig. 10e), underscoring the prospects of the binder for flexible applications.
![]() | ||
| Fig. 10 (a and b) Molecular structure and schematic of the self-healing mechanism in the 2S–PDMS binder, illustrating its role in maintaining cathode integrity and promoting efficient polysulfide conversion. (c) Scratch healing and tensile tests of the 2S–PDMS polymer conducted at various healing times at RT. (d) Cycling stability of soft-packed Li–S batteries using 2S–PDMS–S cathodes with a high S loading of 6.6 mg cm−2 (3 cm × 4 cm) at a current density of 0.42 mA cm−2. (e) Cycling performance of soft-packed Li–S batteries with a S loading of 2.2 mg cm−2, with the inset showing LED operation under repeated bending–unbending conditions. Reproduced with permission from ref. 190. Copyright 2016, Royal Society of Chemistry. (f) Schematic of the zipper-like sulfur electrode integrating a PSPEG polymer binder and a CPS/S composite (AB: acetylene black), depicting how dynamic –Sx– bonds enable structural self-healing and redox mediation. (g) Self-healing behavior of the PSPEG polymer. (h) Electrochemical data showing enhanced cycling performance in Li–S cells at a sulfur loading of 2.5 mg cm−2 (0.2C) and high areal capacities of 4.41 and 6.47 mA h cm−2 at 5.2 and 7.6 mg cm−2, respectively, under 100 mA g−1 with an E/S ratio of 4.5. (i) Comparison of their morphologies reveals the structural benefits of the zipper design over conventional electrodes. Reproduced with permission from ref. 193. Copyright 2021, Elsevier. | ||
Inspired by the interlocking mechanism of zippers, Zeng et al.193 engineered a composite electrode by combining a sulfur nanocomposite (CPS/S, composed of sulfur and organo-polysulfide chains grafted onto a carbon host) with an organo-polysulfide polymer binder (denoted as PSPEG) containing dynamic –Sx– linkages (Fig. 10f). In this design, the dynamic –Sx– bonds in both the binder and sulfur composite serve dual roles: (1) acting as redox mediators to facilitate phase transitions between sulfur, polysulfides, and Li2S, and (2) functioning as “zipper sliders” and “zipper teeth” to autonomously repair mechanical cracks and maintain electrode integrity. The PSPEG binder demonstrated excellent RT self-healing, regaining ∼50% of its tensile strength and elongation within 0.5 h and achieving nearly full recovery (97.6%) after 6 h. It also exhibited remarkable stretchability (545%) and supported a 200 g load after healing, confirming strong mechanical resilience (Fig. 10g). Moreover, the binder remained chemically stable after 72 h of electrolyte exposure. This innovative configuration effectively mitigated sulfur agglomeration and structural collapse during cycling. The cathode retained a capacity of 812 mA h g−1 after 300 cycles at 1C and an areal capacity of 6.47 mA h cm−2 at a sulfur loading of 7.6 mg cm−2 and a low electrolyte/sulfur (E/S) ratio of 4.5 μL mg−1 (Fig. 10h and i). This dual functionality highlights the unique value of sulfur-based dynamic bonds in both structural and chemical stabilization.
In parallel, zwitterionic polymer systems have emerged as a promising binder strategy for Li–S batteries owing to their unique dual-charge molecular structure.210–213 Containing both positively and negatively charged functional groups within a single backbone, these materials enable strong electrostatic and hydrogen bonding interactions with LiPS, effectively anchoring soluble species and mitigating the shuttle effect. Moreover, their internal dipole structure facilitates Li+ dissociation and transport by creating a favorable ionic environment that enhances ion mobility. Building on this concept, Lin et al.204 developed a multifunctional zwitterionic polymer binder termed PLM by incorporating lipoic acid (LA) and the zwitterionic monomer 2-methacryloyloxyethyl phosphorylcholine (MPC) (Fig. 11a). The resulting PLM binder featured a dynamic, reversible crosslinked network composed of S–S bonds, electrostatic interactions, and hydrogen bonding, which together endowed the material with intrinsic self-healing ability and strong affinity for LiPS, which enabled trapping these intermediates. Scratch–recovery tests revealed 94.3% healing efficiency within 6 h at RT (Fig. 11b), while 180° peel tests showed significantly enhanced adhesion strength (1.33 N) compared with that of PVDF (0.59 N). In addition, phosphate and carboxylate anions in the binder enhanced Li+ transport via electrostatic interactions, thereby accelerating the redox kinetics and improving the overall electrochemical performance (Fig. 11c). A nitrogen–phosphorus-based flame-retardant mechanism was also introduced to enhance thermal safety, while the water-processable formulation of PLM supported environmentally sustainable fabrication. Electrochemical evaluations confirmed the effectiveness of this design. The PLM-based cathode exhibited a low capacity fading rate of only 0.041% per cycle over 500 cycles at 1C. Under a moderate sulfur loading of 7.42 mg cm−2, it maintained an areal capacity of 5.76 mA h cm−2 after 150 cycles. Even under challenging conditions, at a high sulfur loading of 11.45 mg cm−2 and with a lean electrolyte volume of 6.4 μL mg−1, the cathode delivered 7.64 mA h cm−2 after 80 cycles, surpassing the areal capacity of current commercial LIBs (Fig. 11d–f). These results highlight the versatility, multifunctionality, and scalability of zwitterionic polymer binders for high-performance, long-lasting, and safer Li–S battery technologies. In summary, SHPBs in Li–S batteries are designed not only to mitigate mechanical degradation, as in LIBs, but also to address chemical instability and polysulfide migration. Designing binders that reflect the distinct electrochemical environments and failure mechanisms of each battery system is critical for fully realizing their practical potential.
![]() | ||
| Fig. 11 (a) Chemical structure and schematic of dynamic interactions within the PLM binder. (b) Visual demonstration of the self-healing capability of the PLM binder through scratch recovery tests. (c) Illustration of the structural integrity and multifunctionality of the cycled PLM-based cathode. (d) Long-term cycling performance of the PLM cathode at 1C. (e) Rate capability of the PLM cathode at various current densities. (f) Cycling stability of the PLM cathode under high S loading (7.42 mg cm−2) with a lean electrolyte (8.5 μL mg−1). Reproduced with permission from ref. 204. Copyright 2025, Elsevier. | ||
Rechargeable sodium-ion batteries (SIBs) have garnered increasing interest for large-scale energy storage owing to the abundance of sodium resources, low raw material costs, and an electrochemical potential comparable to that of lithium (−2.71 V for Na/Na+vs. −3.04 V for Li/Li+).214–216 Although SIBs generally offer lower energy density than LIBs, their cost-effectiveness and environmental advantages make them especially well-suited for stationary applications, where material sustainability and scalability often outweigh energy density requirements (Fig. 12).217,218 Despite these benefits, SIBs face significant material-related challenges. The large ionic radius of Na+ (∼1.02 Å), which is approximately 1.4 times that of Li+ (∼0.76 Å), causes more severe volumetric expansion, especially in alloy-type anode materials such as alloys containing Sn and black phosphorus (BP). These repeated expansion–contraction cycles can severely degrade adjacent components, including particle fracture and pulverization, ultimately disrupting electron pathways, challenges similar to those observed in Si-based LIB anodes.219,220
![]() | ||
| Fig. 12 (a) Comparison of energy densities of M–S and M-ion batteries, the elemental abundance of M in the upper continental crust, and raw material prices of M2CO3, where M = Li or Na. Reproduced with permission from ref. 244. Copyright 2023, John Wiley and Sons. (b) Cost analysis of battery systems, where the theoretical energy density was calculated based on the total mass of anode and cathode active materials. Reproduced with permission from ref. 245. Copyright 2020, The Royal Society of Chemistry. (c) Comparison of key physical and chemical properties of Li and Na. Reproduced with permission from ref. 246. Copyright 2022, John Wiley and Sons. | ||
To mitigate these effects, Callegari et al.221 developed a SHPB based on dynamic quadruple hydrogen bonding by functionalizing PEG telechelics with UPy end-groups (UPyPEGnUPy), blended with PEO, for BP anodes with active mass loadings of 1.26–2.5 mg cm−2 (Fig. 13a and b). This system spontaneously self-healed at RT without requiring external stimuli and effectively accommodated the mechanical stress induced by Na+ intercalation/deintercalation. Among the tested ratios, the 50
:
50 UPyPEG–PEO blend showed the most balanced performance, exhibiting higher crystallinity and improved phase continuity while retaining a low Tg (<−34 °C) after healing, reflecting sustained chain mobility and dynamic bond reformation. It fully recovered from surface scratches and restored electrical conductivity within 2 h (Fig. 13c), and it withstood continuous bending for over 4 h without fracture (Fig. 13d). Compared with a conventional CMC–PAA binder system, the UPy-based binder significantly enhanced the specific capacity of the BP anode from ∼250 to 1750 mA h g−1 (Fig. 13e). The multiple UPy hydrogen bonds enabled autonomous crack repair, thereby preserving structural integrity and ensuring long-term electrical connectivity.
![]() | ||
| Fig. 13 (a) Molecular design and synthesis scheme of the UPy-terminated PEG telechelics (UPyPEGnUPy) used as a SHPB. (b) Schematic comparison of electrode interactions in the UPyPEGnUPy and conventional CMC–PAA binders. (c) Optical microscopy images showing the self-healing process of the polymer binder with different blend compositions. (d) Time-dependent bending test results for various polymer blends. (e) Rate performance comparison of BP anodes utilizing the SHPB versus the conventional CMC–PAA binder. Reproduced with permission from ref. 221. Copyright 2021, American Chemical Society. | ||
Luo et al.222 introduced a hydrogen-bonding-based self-healing strategy for organic SIBs using sodium rhodizonate dibasic (SRD) as the active material and SA as the binder (Fig. 14a and b). Upon sodiation, SRD significantly expands in volume, eventually pulverizing the electrode. However, the hydroxyl-rich SA binder autonomously repaired microcracks and bound the fragmented SRD particles through hydrogen bonding (Fig. 14c). As a result, the SRD–SA electrode exhibited enhanced long-term cycling and rate capability, maintaining a capacity of 140 mA h g−1 after 500 cycles at 50 mA g−1, with a minimal capacity fading rate of 0.051% per cycle (Fig. 14d and e). This study demonstrated that chemical interactions between functional groups in the binder and active material can effectively suppress structural degradation, offering a general approach to improving electrode durability. In another study, Huang and co-workers223 employed tetrahydroxy-1,4-benzoquinone disodium salt (TBDS) as a self-healing organic cathode material for SIBs. Their design leveraged intermolecular hydrogen bonding between hydroxyl and carbonyl groups, which enabled the autonomous healing of microcracks caused by repeated sodiation-induced volume changes (Fig. 14f and g). Compared with control compounds lacking hydroxyl or sodium functionalities, TBDS showed superior electrochemical performance, delivering 76.6 mA h g−1 at 2.0 A g−1 after 1000 cycles (Fig. 14h and i).
![]() | ||
| Fig. 14 (a) Schematic representation of hydrogen bonding interactions between SRD and SA. (b) Electrochemical sodiation/desodiation mechanism of SRD. (c) Magnified images of the SRD–SA electrode after 100 cycles. Comparison of the (d) cycling performance and (e) rate capability of SRD electrodes using SA and PVDF binders. Reproduced with permission from ref. 222. Copyright 2017, Elsevier. (f) Schematic illustration of the intermolecular hydrogen bonding between TBDS molecules, along with a scanning electron microscopy image of TBDS particles. (g) Morphology of the TBDS electrodes before (top) and after cycling (bottom). (h) Cycling performance and coulombic efficiency of TBDS at 2000 mA g−1. (i) Rate capability of TBDS electrodes at various current densities in SIBs. Reproduced with permission from ref. 223. Copyright 2023, John Wiley and Sons. | ||
Together, these studies demonstrated that although still in the early stages, SHPB strategies can be effectively adapted to sodium-based battery chemistries. By addressing mechanical degradation and preserving structural integrity, self-healing designs have the potential to enhance both performance and cycle life in SIBs, thereby broadening the application scope of SHPBs across diverse electrochemical systems.
Despite these advances, several challenges must still be addressed to realize the practical application of SHPBs in commercial battery systems. One important consideration is the environmental and safety impact of conventional fluorine-containing binders, such as PVDF, which often require toxic solvents (e.g., N-methyl-2-pyrrolidone) and raise concerns regarding biodegradability and end-of-life processing. Developing fluorine-free SHPBs that can be processed with greener solvents or dry-coating techniques will be essential for sustainable battery manufacturing. Similarly, the use of renewable or bio-derived polymers offers a promising route to reduce the overall environmental footprint. However, these materials have not yet been extensively studied in the context of SHPB development, and they present additional challenges related to compositional consistency, chemical purity, and processing scalability. These challenges are schematically illustrated in Fig. 15.
Many SHPBs reported to date have been evaluated only under idealized laboratory conditions, such as in coin-type half-cells with a low active material loadings (typically <1 mg cm−2), high binder content (>10 wt%), and excess electrolyte volumes.224 However, real-world battery manufacturing demands much stricter conditions, including high mass loadings (>3–5 mg cm−2),225,226 lean electrolyte-to-capacity ratios (E/C < 5 μL mA h−1),227,228 and minimal inactive material contents.229 Moreover, achieving the appropriate trade-offs between mechanical strength, ionic/electronic conductivity, and self-healing efficiency remains a core bottleneck.230 For example, incorporating flexible polymer domains to promote healing may reduce mechanical stiffness or electrical conductivity. Additionally, dynamic supramolecular interactions often degrade at elevated temperatures or voltages. Although SHPBs exhibit excellent healing in controlled environments, their long-term performance remains largely untested under practical conditions, such as high-rate cycling, wide temperature fluctuations, or high-voltage operation. Very few studies have assessed the retention of self-healing properties over 300–1000 cycles or under repeated mechanical fatigue relevant to pouch or cylindrical cells.231 Moreover, most current SHPB systems are tailored to Si-based anodes, whereas a limited number of studies have explored other battery systems, such as sodium-, potassium-, and zinc-based aqueous batteries as well as all-solid-state and flexible/stretchable batteries.
To fully realize their potential in advanced battery systems, future research on SHPBs should address several key challenges and strategic opportunities.
(1) Rational molecular design of multifunctional binders: a primary challenge in SHPB development is achieving an optimal balance between mechanical robustness, ionic/electronic conductivity, and healing efficiency. Although dynamic supramolecular bonds, such as hydrogen bonding, S–S linkages, and borate esters, enable efficient self-repair, their standalone use may not always provide sufficient adhesion strength or charge transport for long-term operation. Future strategies should focus on integrating these reversible motifs with chemically and mechanically resilient polymer backbones. Designing multifunctional binders that synergistically combine rapid intrinsic healing, shape-memory-assisted recovery, high ionic and electronic conductivity, electrolyte compatibility, and mechanical durability will be essential. For example, incorporating functional components such as ion-conductive segments (e.g., ethylene oxide chains –CH2–CH2–O– or sulfonate groups –SO3H), redox-active moieties (e.g., quinones or nitroxide radicals), or polysulfide-trapping groups (e.g., amino or thiol functionalities) may further improve the electrochemical performance.232
Moreover, to meet commercial standards, decreasing the binder content while maintaining high energy density is crucial. Current SHPBs often exceed 10 wt% in lab-scale electrodes, whereas practical systems typically use less than 3 wt%. This gap presents a significant challenge: achieving effective self-healing when the binder comprises only a small portion of the electrode composition, especially relative to the much higher proportions of active materials and conductive additives. As such, future binder designs must retain robust mechanical integrity and autonomous repair functionality even at low contents, ensuring long-term electrode performance under realistic operating conditions.
(2) Enhancing healing kinetics and operational reliability: binders must be able to self-heal rapidly and repeatedly at RT and in dry environments. Promising approaches include hybrid bonding strategies that combine dynamic covalent and non-covalent interactions, or low-energy reversible systems. A key factor governing healing performance is the Tg of a polymer, which strongly influences chain mobility and the kinetics of bond reformation. Lower Tg values generally facilitate faster healing by enhancing segmental motion and flexibility. Systematic evaluation using techniques such as differential scanning calorimetry or dynamic mechanical analysis should be incorporated into binder characterization protocols to clarify the relationship between thermal properties and healing behavior. In parallel, standardized evaluation metrics, such as healing time, mechanical strength recovery, and conductivity restoration, must be established for consistent benchmarking and cross-study comparison.
Healing efficiency is a critical parameter influencing long-term electrochemical performance. Binders with high healing efficiency can restore interfacial adhesion and conductive pathways following mechanical damage, thereby reducing impedance buildup, minimizing voltage hysteresis, and extending the cycle life. In contrast, limited healing may only temporarily delay structural failure, eventually resulting in contact loss and capacity degradation. Therefore, healing efficiency should be assessed not only as a performance metric of the binder but also as a key indicator of device-level reliability and durability.
(3) Improving chemical and electrochemical stability: SHPBs should maintain their structural and functional integrity in harsh battery environments, including continuous exposure to reactive intermediates (e.g., LiPS in Li–S batteries) and frequent redox cycling. Future designs should prioritize materials that are both chemically robust and dynamically self-healing with the ability to resist oxidative, reductive, and hydrolytic degradation over extended cycling.
(4) Prioritizing sustainability and scalable processing: sustainable design is becoming increasingly central to the development of next-generation battery materials. Future SHPBs should prioritize the use of renewable or bio-derived polymer resources (e.g., lipoic acid, polysaccharides, silk protein, lignin, and CS), which have only been explored in a limited number of studies to date. Expanding research efforts in this direction, particularly toward toxic-solvent-free systems and recyclable or dry-processable binder designs, will be essential to align SHPB development with the broader goals of sustainability, safety, and industrial scalability.233–235 To ensure industrial viability, scalable manufacturing techniques should be considered, such as slurry casting, screen printing, or dry-processable battery electrode fabrication. Life cycle assessments and eco-efficiency evaluations should guide material and process development to align with global sustainability goals.
(5) Leveraging artificial intelligence (AI) and machine learning (ML): AI and ML are powerful tools for accelerating the rational design of SHPBs by uncovering structure–property relationships, identifying key functional moieties, and optimizing binder formulations based on large-scale experimental and computational datasets.236 Looking ahead, the integration of AI and ML tools with high-throughput experimentation could enable the predictive modeling of binder behavior, such as self-healing efficiency, electrochemical stability, or interfacial adhesion under varying conditions. For instance, models capable of predicting the effects of functional groups (e.g., carbonyls, ether linkages, and hydrogen bond donors) on capacity retention, or flagging detrimental moieties such as aromatic amines, could significantly reduce the time required to design and validate effective SHPBs for specific electrode chemistries.237 In addition, computational simulation techniques such as density functional theory and molecular dynamics are expected to further contribute to the design of advanced SHPBs. By enabling an atomistic-level understanding of dynamic bond formation, exchange reaction kinetics, chain mobility, and self-healing mechanisms, these tools can support the predictive screening of functional groups and molecular architectures. As simulation models become more efficient and better integrated with experimental workflows, they increasingly hold promise for accelerating the development of next-generation SHPBs with tailored performance across diverse battery chemistries.238 However, these methods remain computationally demanding and difficult to scale, underscoring the need for more efficient modeling strategies and advanced simulation tools.
(6) Establishing standardized testing protocols for SHPBs aligned with industrial relevance: a major limitation in the current SHPB literature is the lack of unified testing and reporting protocols. This makes it difficult to compare results across studies or assess the practical viability of SHPBs under real battery conditions. To address this gap, future research should focus on developing a systematic, industry-aligned testing framework that enables quantitative, reproducible, and meaningful comparisons across different SHPB systems. Such a framework will guide rational design, performance optimization, and technology transfer from laboratory prototypes to commercial battery applications. Key domains for SHPB evaluation should include.
(i) Chemical and thermal stability: the chemical integrity of SHPBs must be assessed before and after electrochemical cycling to determine whether the functional groups responsible for self-healing are preserved. Fourier transform infrared and Raman spectroscopy can be used to monitor functional group retention, whereas thermogravimetric analysis can evaluate thermal degradation thresholds (e.g., Td5% and Td50%). In addition, X-ray photoelectron spectroscopy (XPS) can be employed to examine changes in the oxidation states and interfacial bonding. These measurements should be conducted under representative stress conditions, including elevated temperatures (45–60 °C) and voltage limits (up to 4.5 V for cathodes or down to 0.01 V for anodes). Additionally, electrolyte uptake and swelling behavior should be quantified to evaluate binder–electrolyte interactions. Excessive swelling may compromise mechanical integrity and interfacial adhesion, whereas insufficient uptake could hinder ionic conductivity. These parameters provide valuable insight into the dimensional stability and interfacial compatibility of SHPBs over extended cycling.
(ii) Mechanical and self-healing properties: qualitative scratch-healing demonstrations alone are insufficient to characterize self-healing performance. Quantitative mechanical testing, such as ASTM D638 tensile measurements,239 should be conducted on pristine and healed specimens to determine the recovery of tensile strength, Young's modulus, elongation at break, and toughness. Fatigue resistance under cyclic loading, assessed via dynamic mechanical analysis or repeated strain testing, can simulate electrode expansion and shrinkage during battery cycling. Furthermore, adhesion strength to active materials and current collectors should be measured using 90°/180° peel tests or shear tests, as interfacial cohesion is crucial for long-term electrode stability.
(iii) Ionic and electronic conductivity and electrochemical performance: electrochemical transport properties should be evaluated before and after mechanical damage or healing. Electrochemical impedance spectroscopy (EIS) and cyclic voltammetry offer insights into ionic and electronic conductivity, charge transfer resistance, redox stability, and potential side reactions involving the binder. SHPBs should be tested under realistic cell conditions using full-cell or pouch-cell configurations, with high areal loading (>3 mA h cm−2), low binder content (<3 wt%), lean electrolyte usage (E/C < 3–5 μL mA h−1), and extended cycling (>500–1000 cycles at ≥C/2). Trends in voltage hysteresis and increasing impedance can further indicate interfacial degradation, while post-cycling measurements of conductivity and capacity retention can help quantify self-healing effectiveness.
(iv) Morphological and interfacial characterization: post-cycling structural analysis is necessary to assess healing at the micro- and nanoscales. Scanning electron microscopy can reveal morphological features and crack closure, atomic force microscopy can map the modulus distribution and phase morphology, and X-ray diffraction can examine crystallinity changes. Complementary surface-sensitive techniques such as XPS and energy-dispersive X-ray spectroscopy provide elemental and chemical-state information, especially at critical interfaces between the binder and active material.
(v) Compatibility with dry electrode processing: given the growing adoption of solvent-free and roll-to-roll electrode fabrication, SHPBs should be assessed for their compatibility with dry-processing techniques. Key parameters include binder fibrillation efficiency under shear mixing, uniform dispersion of active and conductive powders, and the ability to form cohesive films without delamination. Moreover, maintaining mechanical integrity during calendaring and stacking, as well as ensuring continuous ion and electron transport in thick electrodes (>10 mg cm−2), are critical for reliable performance. Incorporating these dry-process-relevant metrics into standardized evaluation protocols will be essential to ensure that SHPBs meet both functional and manufacturing requirements under industrially realistic, solvent-free conditions.
(vi) Protocol transparency and data reporting standards: to facilitate reproducibility and support ML-assisted materials discovery, SHPB studies should follow the standardized reporting practices widely adopted in battery research and industrial testing.240 These include full disclosure of electrode formulations (e.g., binder content and active material ratios), detailed fabrication and cycling conditions (e.g., coating methods, drying temperature, C-rate, and number of cycles), and the presentation of raw datasets such as stress–strain curves, EIS spectra, cycling profiles, and healing efficiency metrics. Transparent reporting will enable meaningful cross-study comparisons and accelerate the rational design of next-generation SHPBs.
In summary, advancing SHPB development requires a multidisciplinary strategy that integrates rational molecular design, rapid and reliable healing behavior, environmental sustainability, and data-driven optimization. These future directions will not only address key limitations, such as the trade-offs between self-healing efficiency and electrochemical performance, but also open new pathways toward scalable, green manufacturing. By leveraging innovations in polymer science, computational modeling, ML, and sustainable processing, SHPBs are well-positioned to enable the next generation of high-performance, durable, and eco-friendly energy storage systems.
The supplementary information includes extended tables (Tables S1 and S2) summarizing SHPB chemistries and battery performance (Li-ion and Li–S), plus additional notes and references. See DOI: https://doi.org/10.1039/d5ta04403k.
| This journal is © The Royal Society of Chemistry 2025 |