Ho Mei Law†
ab,
Zilong Wang†c,
Shengjun Xuab,
Longyun Shend,
Baptiste Pyc,
Yuhao Wangc,
Renée Siegele,
Jürgen Senker
be,
Qingsong Wang
*bf and
Francesco Ciucci
*ab
aChair of Electrode Design for Electrochemical Energy Systems, University of Bayreuth, Weiherstraße 26, 95448 Bayreuth, Bavaria, Germany. E-mail: francesco.ciucci@uni-bayreuth.de
bBavarian Center for Battery Technology (BayBatt), University of Bayreuth, Universitätsstraße 30, 95447 Bayreuth, Germany
cDepartment of Mechanical and Aerospace Engineering, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong SAR, China
dDivision of Emerging Interdisciplinary Areas, The Hong Kong University of Science and Technology, Clear Water Bay, Hong Kong SAR, P. R. China
eDepartment of Chemistry, Inorganic Chemistry III, and Northern Bavarian NMR Centre, University of Bayreuth, Universitätsstraße 30, Bayreuth 95440, Germany
fDepartment of Chemistry, University of Bayreuth, Universitätsstraße 30, 95447 Bayreuth, Germany. E-mail: Qingsong.Wang@uni-bayreuth.de
First published on 19th August 2025
Sodium-ion batteries represent a more sustainable and, potentially, cost-effective alternative to lithium-ion technology, with sodium–metal anodes showing promise for achieving high-energy densities. However, the strong reactivity between sodium–metal and conventional liquid electrolytes leads to unstable solid electrolyte interphases, undermining battery performance and safety. To address this challenge, this work introduces a novel weakly solvating quasi-solid electrolyte. This electrolyte is fabricated via in situ polymerization of polyethylene glycol diacrylate with sodium bis(fluorosulfonyl)imide in a mixed solvent system of 2-methyltetrahydrofuran and cyclopentyl methyl ether, which enables targeted manipulation of the solvation of the sodium cation. Computational and spectroscopic analyses reveal that this design promotes anion-dominated solvation, facilitates the formation of a robust anion-derived solid electrolyte interphase, suppresses dendrite formation, and enhances stability and cell performance. Batteries using this weakly solvating solvent-based quasi-solid electrolyte achieve an average coulombic efficiency of 98.4% over 400 cycles (at 0.5 mA cm−2, 0.5 mAh cm−2 in half-cell tests) and retain a capacity of 1077 mAh g−1 (based on sulfur content) over 250 cycles when paired with sulfurized polyacrylonitrile cathodes. These findings establish a new paradigm for developing practical, high-performance sodium–metal batteries.
Broader contextSodium–metal batteries (NMBs) are rapidly emerging as a promising alternative for next–generation energy storage due to the abundance of sodium, high theoretical capacity, and cost advantages over lithium–ion batteries. However, the practical implementation of NMBs faces a critical challenge: the formation of unstable solid electrolyte interphases (SEIs) at the sodium metal surface. While numerous strategies have been developed to address this issue, most are limited by increased cost, environmental concerns, or safety risks. Here, we propose a transformative approach to electrolyte design by introducing the concept of “weakly solvating solvent-based quasi-solid electrolytes”, which simultaneously addresses several fundamental challenges. By integrating weakly solvating solvents (MTHF and CPME) in situ into a polymerized PEGDA matrix, we fundamentally alter the sodium ion solvation structures to promote anion–dominated coordination environments. This molecular–level strategy yields remarkable electrochemical performance in batteries when paired with sulfurized polyacrylonitrile cells, maintaining high capacity (1077 mAh g−1) after 250 cycles. Our approach represents a comprehensive solution that combines the safety benefits of quasi-solid systems with the performance advantages of controlled solvation chemistry, potentially accelerating the practical deployment of sodium-based energy storage for large–scale applications. |
Recent advances in weakly solvating electrolytes (WSEs) provide an alternative approach that significantly improves battery performance by promoting the formation of inorganic-rich SEIs.7,24–27 Conventional LEs use strongly solvating solvents that readily dissolve Na salts, facilitating the creation of abundant solvent-separated ion pairs (SSIPs, where one or more solvent molecules completely separate the Na+). These strong Na+–solvent interactions lead to the formation of mechanically weak, solvent-derived SEI on Na–metal with poor stability, as illustrated in Fig. 1. In contrast, WSEs use weakly solvating solvents that intentionally limit salt dissolution, thereby reducing Na+–solvent interactions while maximizing Na+–anion interactions.24,28 This alteration in solvation structure creates an anion-rich environment around Na+, even at moderate salt concentrations (in the order of 2 M).7,29 Consequently, the inorganic-rich anions (e.g., bis(fluorosulfonyl)imide (FSI−) from sodium bis(fluorosulfonyl)imide (NaFSI)) form anion-derived SEIs that are inherently stable, resulting in longer cycling life and improved rate capabilities.20,24 For example, including cyclopentyl methyl ether (CPME) in lithium batteries and 2-methyltetrahydrofuran (MTHF) in NMBs as a weakly solvating solvent has led to average CEs of 99.3% over 350 cycles30 and 99.6% over 100 cycles,18 respectively. While MTHF has been validated in Na systems, CPME, despite its proven effectiveness in lithium batteries, remains virtually unexplored in NMB applications. Additionally, these liquid-based WSEs still face fundamental safety concerns related to flammability and leakage.3,8,31–33
Quasi-solid electrolytes (QSEs) address the limitations of LEs through a safer polymer matrix structure.32–34 Comprising a polymer matrix, Na salts, and solvents, QSEs achieve ionic conductivities of 1 to 5 mS cm−1 at room temperature35,36 that are comparable to those of LEs. The in situ polymerization process used to create QSEs enhances electrode–electrolyte contact and, relative to solid-state electrolytes, reduces interfacial resistances, thus improving battery performance.32,37,38 Various polymers, such as pentaerythritol triacrylate,39 poly(ethylene glycol) diacrylate (PEGDA),33 poly(butyl acrylate),34 poly(vinylene carbonate),40 and poly(1,3-dioxolane),41 have been successfully utilized in in situ polymerized QSEs for NMBs, demonstrating promising results. Notably, PEGDA-based QSEs have exhibited exceptional Na metal stability over 2000 hours at 0.1 mA cm−2, attributed to the presence of carbonyl groups (CO), which modify the Na+ solvation structure.33 Despite these advances, current QSE systems for NMBs still employ conventional strongly solvating electrolytes,32–34,41 leading to solvent-derived SEIs.
This work introduces a strategy for controlling the Na+ solvation structure in the QSE through the integration of weakly solvating solvents into the polymer matrix. Our innovation centers on two key points. First, we developed a novel QSE system using in situ free radical polymerization of a liquid precursor, which combines 2 M of NaFSI in MTHF with PEGDA as the crosslinking agent. The CO groups in the PEGDA network weaken Na+–solvent interactions, shifting Na+ coordination toward anions. Second, we enhance this solvation structure by incorporating CPME, a weakly solvating co-solvent that has been previously unexplored in NMBs or QSE applications.30,42 This combination creates a weakly solvating solvent-based QSE (WS-QSE) that promotes the formation of anion-rich Na+ complexes, including contact ion pairs (CIPs, where an anion coordinates to a single Na+) and aggregate ion pairs (AIPs, where an anion coordinates to two or more Na+), and results in an anion-dominated Na+ solvation structure. In turn, this controlled solvation environment facilitates the formation of a stable, anion-derived SEI on the Na–metal surface through the reductive decomposition, as illustrated in Fig. 1.
Electrochemical performance tests confirmed the effectiveness of the system. In Na|carbon-coated aluminum (Al) half-cell tests, the WS-QSE achieved an impressive average CE of 98.4% (at 0.5 mA cm−2, 0.5 mAh cm−2) over 400 cycles. We also evaluated WS-QSE with a sulfurized polyacrylonitrile (SPAN) cathode, which was selected for its high theoretical capacity (1675 mAh g−1) and structural stability due to the presence of covalent sulfur-polyacrylonitrile bonds.13,43,44 The Na|WS-QSE|SPAN system showed remarkable cycling stability (all capacities based on sulfur content unless indicated otherwise), retaining a specific capacity of 1077 mAh g−1 (484.7 mAh gSPAN−1) after 250 cycles at 240 mA g−1 or 0.32 mA cm−2. Even with an increased areal capacity of 2.3 mAh cm−2, the battery could retain 1099 mAh g−1 (494.6 mAh gSPAN−1) over 150 cycles at 120 mA g−1 (0.25 mA cm−2). In short, this work develops a novel electrolyte design in NMBs, contributing to both a fundamental understanding of ion solvation behavior in QSEs and the practical development of next-generation energy storage systems.
To improve performance through molecular design, we chose two structurally similar cyclic ether-based solvents, MTHF and CPME. While both molecules contain a five-membered ring with an oxygen atom, they exhibit significant differences in their molecular structures. As shown in Fig. 2a, CPME features a bulky cyclopentyl ring connected to oxygen through a methyl bridge, which creates a greater steric hindrance and consequently solvates Na+ more weakly than MTHF. This difference in solvating ability is quantitatively supported by the dielectric constants (ε), which indicate an ability to dissociate ion pairs.10,24,47 CPME exhibits a lower value (ε = 4.7)30 than MTHF (ε = 6.97)26 (Fig. 2a and Table S1 in the SI). Dipole moments (μ, Table S1) further confirm this trend, with CPME showing a lower value (μ = 1.27 D) than MTHF (μ = 1.38 D). Both cyclic ethers demonstrate weaker solvation properties than the conventional linear ether-based solvent 1,2-dimethoxyethane (DME, ε = 7.2,24 μ = 1.71 D, Table S1), making them promising candidates for our weakly solvating solvent strategy.
Density functional theory (DFT) calculations revealed the specific electronic interactions between Na+ and the solvent molecules, providing molecular-level insights into their coordination behavior. Fig. S1 displays negative charge densities at the oxygen atoms of both MTHF and CPME in the electrostatic potential analysis, indicating that these oxygen atoms are preferential binding sites for Na+. However, CPME showed a lower negative electron density (−1.17 eV) at the oxygen atom compared to MTHF (−1.3 eV) (Fig. S1), suggesting weaker Na+–CPME interactions. Binding energy calculations, which quantify the thermodynamic stability of Na+–solvent complexes, further support these findings. As illustrated in Fig. 2b, the Na+–MTHF complex exhibits a higher negative binding energy (−1.41 eV) than Na+–CPME (−1.36 eV). These computational results align well with experimental observations: pure CPME cannot dissolve NaFSI at a 0.5 M concentration (Fig. S2a), while a MTHF/CPME mixture (50:
50, volume ratio) successfully dissolves NaFSI up to a 2.0 M concentration (Fig. S2b). Based on these results, we chose MTHF as the primary solvent for NaFSI dissociation, with CPME acting as a co-solvent to modulate the Na+ solvation structure due to its lower solvation power.
For the polymer component selection, we identified PEGDA as the optimal crosslinking agent due to its exceptional networking-forming capabilities,33,48 which facilitated the formation of a polymer network. Through thermally initiated free-radical in situ polymerization of the LE (2.0 M NaFSI in MTHF) with PEGDA, as discussed in Fig. S3 and the Methods section in the SI, we obtained a QSE with solid-like polymer characteristics after evaluating several candidate monomers (Fig. S4 and SI I). In addition to structural benefits, the carbonyl groups (CO) in PEGDA play a critical role in manipulating the Na+ solvation structure.33,49 The presence of C
O groups weakens Na+–solvent bonds33,34,49 reducing the formation of SSIPs (Fig. 1). To further promote an anion-dominated solvation structure, we incorporated CPME as a weakly solvating co-solvent (MTHF
:
CPME = 50
:
50 vol%) in the liquid precursor. This formulation, which we designate as WS-QSE, combines PEGDA's carbonyl interactions with CPME's inherently weak solvating properties to create an environment that favors CIPs and AIPs, significantly enhancing the stability of the resulting SEI layer (Fig. 1).
We used molecular dynamics (MD) simulations (Fig. 2c–e) to investigate the Na+ solvation environment in the WS-QSE and validate our design principles. Using the local minimum (2.9 Å) in the radial distribution function (RDF) as the shell boundary,37 we calculated coordination numbers (CNs) for various Na+–molecule interactions (Fig. 2f). These CNs quantify the number of solvent molecules or anions surrounding each Na+ and provide direct evidence for solvation structure. The simulations revealed distinct patterns of anion coordination: the CNs of Na+–OFSI− were 3.75, 3.75, and 4 for WS-QSE, QSE, and LE, respectively. In contrast, the CNs of Na+–solvent (encompassing both Na+–OMTHF and Na+–OCPME interactions), showing an inverse relationship: 0.31 for WS-QSE, 0.36 for QSE, and 1.5 for LE. This significant reduction in solvent coordination, while maintaining similar anion coordination, indicates a predominantly anion-dominated solvation environment in both QSE systems, with WS-QSE showing the most pronounced effect. The anion-to-solvent ratio (ASR) was used to compare solvation structures (Fig. 2g, h and Fig. S5).50,51 WS-QSE achieved the highest distribution of ASR ≥ 4 (18.4%), exceeding both QSE (17.6%) and LE (14.7%). This improved ratio can be attributed to the synergistic effects of adding PEGDA and CPME. While the LE showed high CNs of Na+–OFSI−, its substantial solvent binding resulted in an overall lower ASR ≥ 4. These computational findings provide preliminary evidence for an anion-dominated solvation structure in WS-QSE.
Informed by these solvation environment findings, we then analyzed the molecular orbital energy levels to evaluate the electrochemical stability of the different coordination configurations. The lowest unoccupied molecular orbital (LUMO) levels were calculated since they affect SEI formation, while the highest occupied molecular orbital (HOMO) levels influence oxidation stability at the cathode. As shown in Fig. 2i, the LUMO energies of MTHF (−0.24 eV) and CPME (−0.26 eV) are nearly identical, indicating comparable electrochemical stability. However, upon Na+ coordination, their complexes exhibited significantly lower LUMO levels: −5.57 eV for Na+–MTHF and −5.49 eV for Na+–CPME. According to the literature,52 this decrease in LUMO energies results from orbital hybridization between Na atomic orbitals and solvent molecular orbitals, leading to lower energy states that promote C–O bond cleavage and subsequent decomposition. Consequently, these coordinated complexes decompose more readily, resulting in less stable, organic solvent-derived SEIs.27 In contrast, the Na+–FSI− complexes display a much lower LUMO level (−2.16 eV) compared to the bare FSI− anion (3.06 eV). This notable energy difference suggests that Na+–FSI− complexes are more susceptible to reduction than the free FSI− anion, facilitating the formation of robust, inorganic-derived SEIs on the Na–metal surface. Thus, manipulating an anion-dominated solvation structure in the electrolyte promotes the formation of more stable, anion-derived SEIs.27
We used a suite of spectroscopic techniques to investigate the anion-dominated Na+ solvation structure in WS-QSE. Raman spectroscopy of the S–N–S stretching of FSI− (720 to 750 cm−1) revealed that CIPs and AIPs account for the majority (>70%) of the species in all electrolyte systems (Fig. 3d, e and Fig. S7). Raman spectroscopy revealed significant differences in AIP content across electrolyte systems, with WS-QSE containing the highest proportion (22%) compared to QSE (14%) and LE (11%) (Fig. S7), as evidenced by the characteristic peak shifts shown in Fig. 3e. This enhancement resulted from two synergistic effects: the strong Na+ binding to the polymer backbone, as evidenced by the shift of the CO peak of PEGDA from 1720 to 1727 cm−1 (Fig. 3c and d), and the reduced solvating power of CPME. These factors effectively reduced SSIPs from 17% in LE to only 4% in WS-QSE (Fig. 3e and Fig. S7), indicating the successful establishment of an anion-dominated Na+ solvation structure.
To validate our Raman spectroscopy findings with complementary analytical evidence, we employed magic-angle spinning (MAS) solid-state nuclear magnetic resonance (NMR) spectroscopy, providing direct molecular-level insights into the Na+ solvation environments. As shown in Fig. 3f and Fig. S8a, the 23Na MAS spectra displayed progressive upfield shifts from LE (−7.1 ppm) to QSE (−10.4 ppm) and to WS-QSE (−12.8 ppm). This trend indicates increasing electron shielding around Na+, with the enhanced shielding attributed to PEGDA incorporation, which expands the first coordination shell for QSE and WS-QSE. A similar shift (−11.4 ppm) was observed for the 23Na resonance of crystalline NaFSI (Fig. S8a), where the Na+ is surrounded by seven Lewis basic centers (Fig. S8c). The additional upfield shift observed for WS-QSE indicates a stronger Na+–anion interaction as promoted by the weakly solvating solvent.45
Complementary 19F MAS spectra (Fig. 3g and Fig. S8b) revealed a consistent pattern, with WS-QSE showing the highest downfield shifts (56.5 ppm) compared to QSE (56.2 ppm) and LE (56.0 ppm). These shifts are similar to the mean 19F chemical shift (56.8 ppm) observed in crystalline NaFSI (Fig. S8b) and suggest predominantly chelating coordination of FSI−. The decreased electron shielding around fluorine atoms in WS-QSE suggests a stronger interaction between Na+ and FSI− there compared to LE and QSE.18,46 The near-zero 19F chemical-shift anisotropies demonstrate that the FSI− anion reoriented rapidly and isotropically, despite confinement within the PEGDA matrix. These NMR analyses align well with our experimental observations from Raman spectroscopy (Fig. 3d and e) and computational findings from MD simulation (Fig. 2c–h), providing further evidence for an anion-dominated solvation structure in WS-QSE.
Symmetrical cells were examined to understand the stability of the Na/electrolyte interphase. As shown in Fig. 4e, WS-QSE demonstrated exceptional longevity, maintaining stability for 800 hours compared to QSE (350 hours) and LE (80 hours) in Na|Na cells, at a current density of 0.05 mA cm−2 with matched capacity density. This superior performance was achieved despite a slightly higher overpotential (0.1 V vs. Na+/Na) due to lower ionic conductivities (0.28 mS cm−1) for WS-QSE (Fig. S12), a characteristic trade-off noted in electrolytes with low solvating power.29,30 Optimizing ionic conductivity was beyond the scope of this work. However, it presents a clear avenue for future research through the inclusion of strongly solvating additives58 or modifications to the polymer backbone.
When the current density was increased from 0.01 to 1 mA cm−2 (Fig. 4f), WS-QSE maintained stable performance. In contrast, LE and QSE failed at much lower current densities (0.2 and 0.5 mA cm−2, respectively). This improved rate performance was observed at higher current and capacity densities (0.5 mA cm−2 and 0.5 mAh cm−2) as illustrated in Fig. S13a, and further confirmed under more demanding conditions of 1.0 mA cm−2 and 1.0 mAh cm−2, where WS-QSE consistently cycled for 40 hours, while QSE and LE failed within 3 hours and 1 hour, respectively (Fig. S13b–d). Surface characterization revealed the mechanisms underlying the enhanced performance of WS-QSE. As shown in Fig. 4g, scanning electron microscopy (SEM) analysis of the Na–metal anode surface after 130 hours of cycling demonstrated that WS-QSE produced a compact and uniform surface structure, indicating stable SEI formation. In contrast, QSE resulted in a loosely structured surface, while LE produced a porous and pulverized surface morphology, suggesting less effective surface protection. X-ray photoelectron spectroscopy (XPS) analysis provided insights into the SEI composition (Fig. 4h, i, and Fig. S14). Analysis of the F1s spectra (Fig. 4h) revealed characteristic peaks at 687 and 684 eV, corresponding to S–F and NaF bonds, respectively,59 resulting from NaFSI decomposition. Notably, both WS-QSE and QSE exhibited significantly stronger NaF (684 eV) signals compared to LE, indicating the enhanced reduction of Na+–FSI− complexes. This finding is consistent with their anion-dominated solvation structures previously observed through Raman spectroscopy (Fig. 3d, e and Fig. S7) and MD simulations (Fig. 2c–h). The abundance of NaF is particularly beneficial, as it has been reported to suppress Na dendrite growth and prolong the lifespan of the Na–metal anode.8,13 Further analysis of the N1s spectra (Fig. 4i) revealed additional insights into the interphase composition. Both WS-QSE and QSE showed two distinct peaks, namely S–N (399 eV) and Na3N (398 eV),9 which also originate from NaFSI decomposition. The presence of Na3N is particularly significant as it promotes rapid Na+ diffusion through the SEI and uniform Na deposition.17,60,61 Furthermore, its high mechanical strength (shear modulus: 23.95 GPa)62 is crucial for suppressing dendritic growth. When combined with the robust NaF (shear modulus: 31.4 GPa),63 this creates a composite inorganic SEI that simultaneously delivers fast ionic transport, mechanical stability, and dendrite suppression.61,63 The more pronounced Na3N signal observed in WS-QSE directly correlates with its superior rate performance and extended cycling stability observed in symmetrical cell tests, providing compelling evidence that our solvation structure design strategy effectively promotes the formation of a favorable SEI.
Cycling stability tests further validated the advantages of WS-QSE. After three formation cycles at 120 mA g−1, the cells were cycled at 240 mA g−1. As shown in Fig. 5b and Fig. S17, the WS-QSE system sustained a capacity of 1079 mAh g−1 over 100 cycles, significantly outperforming both the QSE (936 mAh g−1) and LE systems, with the latter failing within five cycles. This improved stability can be attributed to the protective PEGDA framework and the robust SEI formation previously characterized (Fig. 4g–i).
Post-cycling XPS analysis was used to provide key insights into the cathode-electrolyte interface (CEI) in a fully charged state (Fig. S18 and SI II). While both the WS-QSE (Fig. S18d) and QSE (Fig. S18c) systems showed signs of PEGDA backbone decomposition as indicated by strong CO intensities, the LE system exhibited high levels of SOx–F and C–S/C–S species (Fig. S18b), indicating that uncontrolled NaFSI salt decomposition formed an ineffective CEI. Importantly, all systems showed the presence of NaF, which helps suppress polysulfide dissolution.21 The absence of Na2S (∼160 eV)15,21 in all electrolytes (Fig. S18b–d) suggests that the electrochemical reactions remained largely reversible for the SPAN cathode. This finding provides support that the rapid failure of the LE-based cell originates not from the cathode, but from the inherent instability of the Na anode.
XPS depth profiling analysis was used to demonstrate that the SEI thickness follows the order WS-QSE ≥ QSE > LE (Fig. S19 and S20 with detailed discussion provided in SI III), with WS-QSE maintaining both stable C1s and F1s signals throughout 30 minutes of sputtering, compared to the rapid signal decay in LE and QSE. The superior interfacial stability of the WS-QSE is attributed to its protective role of PEGDA and the formation of a more uniform SEI layer (Fig. S19 and S20 and Fig. 4g), both factors leading to improved cycling stability with reduced capacity fade.
The rate capability testing revealed an interesting performance tradeoff. As illustrated in Fig. 5c and d and Fig. S21, the batteries with QSE exhibited marginally higher capacity (1134 mAh g−1 vs. 1097 mAh g−1) at low rates (120 mA g−1) due to its superior ionic conductivity (Fig. S12). However, the WS-QSE batteries demonstrated a better capacity retention at higher rates, maintaining 634 mAh g−1 at 2400 mA g−1, compared to QSE's 404 mAh g−1. This enhanced rate performance correlates directly with the NaF and Na3N-rich interphase structure identified by XPS analysis (Fig. 4h and i).
Most significantly, long-term testing with sulfur loadings of 1.35 mg cm−2 further supported the stability of WS-QSE. As illustrated in Fig. 5e and f, the WS-QSE had excellent cycling stability, maintaining a capacity of 1077 mAh g−1 (484.7 mAh gSPAN−1) over 250 cycles at 240 mA g−1 (or 0.32 mA cm−2). Even more impressively, at a higher sulfur loading of 2.1 mg cm−2, the cells sustained 1099 mAh g−1 (494.6 mAh gSPAN−1) over 150 cycles, achieving a high areal capacity of 2.3 mAh cm−2 at 120 mA g−1 (or 0.25 mA cm−2), as shown in Fig. 5g. These results surpass previously reported room temperature Na/sulfur batteries in both loading and capacity retention (Fig. 5h and Table S2),11–13,15,17,18,32,39,45,60,65 To demonstrate its practical applicability, we tested the WS-QSE in full cells with limited-excess Na anodes prepared by two different methods. Anodes prepared by mechanical rolling (N/P ratio = 6.7:
1, Fig. S22) retained a capacity of 1031 mAh g−1 after 20 cycles. Those prepared by electrodeposition (N/P ratio = 3
:
1, Fig. S23) retained 780 mAh g−1 over the same period. These results confirm the robustness of the WS-QSE electrolyte system under practical conditions with lower Na excess (Fig. S22 and S23).
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |