Tianqi Xuabc,
Zongxu Yaoabc,
Wei Jiangabc,
Yaxin Chend,
Chenmin Liuabc,
Yinshuang Guanabc,
Zhiqiang Tangabc and
Liang Dong*abc
aJiangsu Key Laboratory for Clean Utilization of Carbon Resources, China University of Mining and Technology, Xuzhou 221116, China. E-mail: dongl@cumt.edu.cn
bInternational Joint Laboratory of Minerals Efficient Processing and Utilization, Ministry of Education, China University of Mining and Technology, Xuzhou 221116, China
cSchool of Chemical Engineering & Technology, China University of Mining and Technology, Xuzhou 221116, China
dSchool of Materials Science and Physics, China University of Mining and Technology, Xuzhou, 221116, China
First published on 3rd July 2025
Coal-derived hard carbon is a highly promising anode material for sodium-ion batteries (SIBs); however, its inherent structural disorder, limited porosity, and sluggish ion transport kinetics severely restrict its electrochemical performance. To address these challenges, this study introduces C–Na–O interaction motifs via the co-pyrolysis of Xinjiang bituminous coal with glucose and sodium carbonate. The resulting hard carbon exhibits excellent electrochemical performance, delivering a high reversible capacity of 299.8 mA h g−1, an initial coulombic efficiency of 90.46%, and a retained capacity of 217 mA h g−1 after 800 cycles at 0.2 A g−1. Galvanostatic Intermittent Titration Technique (GITT) measurements and in situ Raman spectroscopy reveal the Na+ storage mechanism. Density functional theory (DFT) calculations demonstrate that, following oxygen incorporation, Na is adsorbed through synergistic interactions with carbon and oxygen atoms to form C⋯Na–O and C–Na–O configurations. Among them, the C–Na–O structure facilitates Na+ storage and diffusion, while the C⋯Na–O configuration contributes to enhancing the initial coulombic efficiency. This work presents a structurally integrated and scalable strategy for constructing fast-ion-conducting channels in coal-based hard carbon, offering new mechanistic insights and practical guidance for the development of high-performance and sustainable SIB anode materials.
In recent years, co-pyrolysis of biomass with coal has emerged as an effective strategy for tailoring the microstructure of coal-derived hard carbon. For example, studies by Aboyade et al.,8 Ding et al.,9 and Wu et al.10 have shown that the incorporation of biomass during pyrolysis can significantly improve the pore structure and morphology of the resulting carbon materials. Further, Chen et al.11 revealed that sucrose can form ester-linked cross-linked structures with lignite at around 400 °C, thereby enhancing the structural stability of the carbon precursor. Zhou et al.12 also demonstrated that glucose, rich in hydroxyl groups, can graft onto lignite during low-temperature carbonization, suppressing gas release and impurity formation, reducing structural defects, and ultimately improving the electrochemical performance of coal-derived hard carbon. Despite these benefits, co-pyrolysis of coal with glucose can only moderately enhance the initial coulombic efficiency (ICE), which typically remains in the range of 82–84%, falling short of achieving high-ICE hard carbon materials. Against this backdrop, researchers have begun exploring the use of inorganic sodium sources to further regulate ICE.13,14 Among them, sodium carbonate (Na2CO3) has attracted considerable attention due to its ability to improve pyrolysis efficiency, alter decomposition kinetics, and activate intrinsic mineral matter in coal. Previous studies have confirmed that Na2CO3 can accelerate volatile release and reduce the activation energy of coal pyrolysis,15 as reported by Valluri et al.,16 Liu et al.,17 and Yang et al.18 Additionally, Zhao et al.19 showed that sludge carbonized in a molten KOH/Na2CO3 system exhibited significant improvements in specific surface area and carbon content. However, whether Na2CO3 can synergistically promote carbon structural ordering and Na+-active site formation during the co-pyrolysis of coal and glucose—and how it regulates the sodium storage mechanism at the atomic level—remains largely unexplored and warrants further investigation.
Herein, we propose a dual-additive co-pyrolysis strategy that combines glucose and Na2CO3 to engineer the multi-scale structure of coal-derived hard carbon. Specifically, glucose, owing to its high oxygen content and propensity for low-temperature carbonization, contributes to the introduction of oxygen-containing functional groups and adjusts the degree of disorder in the carbon skeleton. These modifications are expected to improve Na+ diffusion pathways. Concurrently, Na2CO3 serves as a synergistic modifier with two key roles: on one hand, Na+ can intercalate into the interlayer spaces of carbon materials at high temperatures, thereby expanding the interlayer spacing and improving the Na+ kinetics. On the other hand, the alkaline environment generated upon Na2CO3 decomposition may catalyze the coal–glucose co-carbonization reaction, thereby promoting the development of micro- and mesoporous structures. This integrated co-pyrolysis route enables simultaneous tuning of pore structure, surface chemistry, and promotes the formation of C–Na–O Structure, ultimately offering a promising pathway for constructing high-capacity, long-cycle-life coal-based hard carbon anodes for advanced sodium-ion batteries.
To evaluate the necessity and effect of acid leaching on sodium storage performance, a modified preparation route was employed. Raw coal (RC) was first demineralized through sequential acid treatment. The procedure involved initial digestion in 5 M HCl solution (solid–liquid ratio 1:
30, stirred for 12 hours), followed by etching in 5 M HF under the same conditions to remove residual inorganic minerals. The acid-treated samples were thoroughly rinsed with deionized water until a neutral pH was achieved, and subsequently oven-dried at 105 °C to obtain deashed coal (CC). The CC was then combined with glucose at the same mass ratios used for the BCG series (i.e., 1
:
0, 1
:
0.5, 1
:
1, and 1
:
2), followed by identical ball milling and pyrolysis procedures under argon atmosphere. The resulting hard carbon products were labeled as CCG0, CCG0.5, CCG, and CCG2. This parallel preparation route allowed for a systematic comparison between acid-washed and unwashed precursors in terms of their structural features and electrochemical properties.
Sample | Element composition (wt%) | ||||||
---|---|---|---|---|---|---|---|
Ca | Ha | Ob | Na | Sa | Nac | ||
a Tested by EA.b By difference (100%–C%–H%–N%–S%).c Tested by ICP. | |||||||
Before carbonization | RC | 75.48 | 4.55 | 18.66 | 1.04 | 0.27 | — |
BCGNa0.1 | 54.99 | 5.14 | 39.34 | 0.53 | 0.00 | — | |
After carbonization | CG0 | 94.25 | 0.19 | 4.95 | 0.32 | 0.29 | — |
CGNa0.1 | 93.03 | 0.17 | 6.39 | 0.23 | 0.18 | 0.17 |
Based on DFT, the CASTEP module in Materials Studio was employed to calculate the adsorption and diffusion of Na atoms on the carbon surface. All calculations were carried out using the generalized gradient approximation (GGA) with the PBE21 exchange-correlation functional. K-points on the Brillouin zone were 1 × 1 × 2, and the energy cutoff of 517 eV. To ensure the stability of the optimized geometries, the force convergence was set to be 0.05 eV·Å−1.
Furthermore, Fig. 1e reveals that the CG0 sample exhibits highly disordered carbon layers, which facilitate Na+ adsorption and enhance capacity in the high-voltage region. The introduction of glucose reduces the edge disorder of the carbon layers, while the addition of Na2CO3 promotes the formation of abundant nanopores during pyrolysis, driven by the release of CO2. Na2CO3 decomposes at approximately 850 °C into Na2O and CO2, followed by further reaction at higher temperatures to form metallic Na.22 The resulting Na deposits onto the hard carbon surface within the sealed ceramic boat, facilitating the formation of C–Na–O interactions. To further confirm the existence of C–Na–O interaction structures in the coal-derived material, TOF-SIMS measurements were performed. As shown in Fig. 2, several characteristic secondary ion fragments were detected in the CGNa0.1 sample, including [NaCO3]− (m/z = 82.974), [NaCO4]− (m/z = 98.987), [NaHCO4]− (m/z = 99.985), [NaC3CO4]− (m/z = 122.969), [Na3CH2O5]− (m/z = 162.959), and [Na3CO3]+ (m/z = 128.972). These fragments are consistent with the decomposition patterns of sodium carboxylate species, indicating stable interactions between Na and the C/O framework. Based on the detection of fragments such as [Na3CH2O5]−, it is further inferred that Na may exist in both chemically bonded and physically adsorbed forms with C and O. Notably, Na2CO3 plays a critical role in both accelerating the pyrolysis of coal (Fig. S5†) and promoting the formation of an ordered microstructure, leading to the development of hard carbon with an interlayer spacing of up to 0.41 nm. This structural feature is beneficial for enhancing Na+ diffusion kinetics and rate performance. In addition, SEM-EDS and TEM-EDS analyses (Fig. S6†) reveal that the introduced Na is uniformly distributed within the material, further confirming the significant impact of Na incorporation on the electrochemical performance.
The structural evolution of CG0, CG, and CGNa0.1 samples was systematically examined, as illustrated in Fig. 3. The XRD patterns of the carbonized samples (Fig. 3a) show broad (002) and (100) diffraction peaks, characteristic of amorphous carbon structures. Compared with the samples before carbonization (Fig. S8a†), the characteristic peaks associated with glucose completely vanished after carbonization, confirming the complete decomposition of glucose during thermal treatment. Additionally, no sodium oxide diffraction peaks were detected in the CGNa0.1 sample, implying that the introduced sodium was embedded into the hard carbon matrix through electrostatic interaction. Moreover, the addition of glucose and Na2CO3 effectively suppressed the retention of mineral phases in coal. Raman analysis (Fig. 3b) indicated changes in carbon structural disorder, with the CG0 sample showing a highly disordered carbon layer structure and an AD/AG ratio of 3.31. Glucose doping greatly reduced this ratio, reflecting a clear decrease in defect density. Following the addition of Na2CO3, the AD/AG ratio of CGNa0.1 was close to that of CG, further highlighting the dominant role of glucose in controlling carbon layer ordering (Table S3†). In addition, The in-plane crystallite size La23,24 was calculated based on the AD1/AG ratio using the modified Tuinstra–Koenig relation:
La = Cλ/(AD1/AG) | (1) |
XPS analysis (Fig. 3d–f) reveals notable modifications in surface chemical composition. Compared to CG0, both CG and CGNa0.1 exhibit increased sp2-hybridized carbon and decreased sp3-hybridized carbon content, indicating enhanced graphitization and reduced structural disorder, which aligns well with the Raman results. Interestingly, CGNa0.1 shows higher absolute amounts of both sp2C and sp3C species compared to CG, which may be attributed to lattice distortions induced by Na incorporation; however, the sp2C/sp3C ratio is nearly close (Table S4†), suggesting that their degrees of order are similar. The O 1s spectra further confirm the enrichment of C–O–C functional groups in CG and CGNa0.1 (Table S5†), consistent with the FTIR results. In addition, the Na 1s spectrum of the CGNa0.1 sample (Fig. 3f) shows a prominent peak at approximately 1071.5 eV, attributed to Na, and another peak appears at around 1073.2 eV, which is ascribed to the C–Na–O interaction. Nitrogen adsorption–desorption isotherms (Fig. 3g) exhibit typical Type II characteristics, with CGNa0.1 achieving the highest BET surface area (7.02 m2 g−1) compared to CG0 (1.70 m2 g−1), reflecting its more developed pore structure. The pore size distribution (Fig. 3h) shows that the CG sample mainly exhibits a macroporous structure centered around 40–80 nm, which contributes to enhanced capacity in the high-voltage region. In contrast, CGNa0.1 demonstrates a significantly increased surface area and microporosity, which facilitate ion transport and electrolyte accessibility, thereby improving its electrochemical performance. To further investigate the microporous structures of the three samples, we compared the SAXS curves (q range: 0.06–1 Å) of CG0, CG, and CGNa0.1, and analysed their pore size distributions in the 0.6–2 nm range, as shown in Fig. 3i. The results indicate that CGNa0.1 contains a large number of ultramicropores, further confirming its abundant nanoporous architecture.
Fig. 4c compares the rate performance of CG0, CG, and CGNa0.1, with detailed charge–discharge profiles of CGNa0.1 shown in Fig. 4d. Within the low current density range of 20–30 mA g−1, both CG and CGNa0.1 maintained stable capacities. However, CG exhibited significant capacity decay at higher current densities (100–500 mA g−1), followed by a sharp drop at 500–1000 mA g−1. In contrast, CG0 displayed a continuous decline in capacity across the entire range from 20 to 1000 mA g−1. All samples showed good rate recovery when the current density was returned to 200 mA g−1, indicating structural resilience. For CGNa0.1, the characteristic desodiation “gentle slope” in the 0.5–0.9 V range, observed under 20–30 mA g−1, gradually disappeared at higher current densities. This phenomenon suggests that with increasing current, the contribution from intercalation and pore-filling mechanisms decreases, while the disturbance to surface-adsorbed Na+ during extraction is reduced. The weakened interaction leads to the disappearance of the “gentle slope” in the charging curve, reflecting changes in sodium storage kinetics under high-rate conditions. Fig. 4e presents the cyclic voltammetry (CV) curves of CGNa0.1 during the first three cycles. A slight irreversible capacity near 0.5 V is observed in the initial cycle, while the subsequent cycles show well-overlapped profiles, indicating good electrochemical reversibility and cycling stability.
To further investigate the cause of capacity degradation of CG0 under high current densities, cycling tests were performed at 200 and 500 mA g−1 for 100 cycles, as shown in Fig. 4f. The severe capacity fading of CG0 is primarily attributed to structural degradation and surface cracking, which was confirmed by post-cycling SEM observations. Fig. 4g shows that CGNa0.1 maintains 86.3% of its capacity after 800 cycles at 200 mA g−1, demonstrating excellent long-term cycling stability. Moreover, we investigated the effects of precursor mixing ratios and intrinsic mineral components on sodium storage performance (Fig. 4h, S7c and d†). It was observed that, in addition to CG0 (8:
1
:
1) and CCG0 (8
:
1
:
1), both CGx (8
:
1
:
1) and CCGx (8
:
1
:
1) exhibited a distinct “gentle slope” in the charge profile between 0.5–0.9 V, similar to CGNa0.1. However, this feature was absent in CG (90
:
5
:
5), suggesting that the 8
:
1
:
1 slurry ratio improves electrode conductivity and structural integrity, thereby facilitating Na+ insertion and increasing capacity. When the discharge capacity exceeds a threshold (∼400 mA h g−1), the formation of C–Na–O interactions is promoted, which benefits the reversibility of desodiation. Nevertheless, the increased proportion of conductive agent and binder leads to a significant reduction in initial coulombic efficiency. In contrast, CGNa0.1, with the introduction of an external sodium source, effectively lowers the capacity threshold for C–Na–O formation, which improves the initial efficiency despite a slight capacity loss. Furthermore, as shown in Fig. 3a and S8a and b†, minerals such as kaolinite and pyrite in raw coal are converted into electrochemically active iron oxides and SiO2 during pyrolysis, while no obvious impurity peaks are detected in acid-treated CCG0, indicating effective mineral removal. Compared to CG0, acid treatment improves the initial coulombic efficiency from 66.6% to 72.0%, but reduces the reversible capacity from 255.1 to 244.5 mA h g−1. Although these mineral-derived phases can contribute to capacity, they primarily induce irreversible reactions. Overall, CGNa0.1 achieves the best comprehensive performance among all samples by balancing capacity and efficiency without relying on environmentally harmful acid treatments or complex recovery processes.
The Na+ storage behavior of coal-derived hard carbon was investigated through kinetic analysis using cyclic voltammetry (CV) curves of CG0, CG, and CGNa0.1 at various scan rates (Fig. 5a, S9a and b†). The relationship between peak current (i) and scan rate (v) was evaluated using the equation i = avb, where a and b are constants related to the charge storage mechanism.28 The value of b was obtained by plotting log(i) versus log(v); b = 0.5 corresponds to a diffusion-controlled process, while b = 1.0 indicates a capacitive-dominated behavior,29 as shown in Fig. 5b, S9c and d.† In the plateau region, the b values for CG0, CG, and CGNa0.1 were 0.530, 0.422, and 0.358, respectively, confirming that Na+ storage in this region is predominantly diffusion-controlled. In contrast, the sloping regions exhibited b values of 0.924, 0.941, and 0.913, respectively, suggesting that capacitive contributions dominate in these voltage ranges.30 Furthermore, with increasing scan rates, CGNa0.1 exhibited the highest fitting correlation (R2 = 0.9852) for the reduction peak P, indicating superior kinetic stability and consistent Na+ storage behavior under varying scan rates.
In in situ Raman spectra (Fig. 5c), as Na+ enters the hard carbons, it affects the intensity and position of the G and D bands.30,31 From open-circuit voltage to 0.5 V, the positions of the G and D bands remain nearly constant, suggesting that Na+ exhibits a high diffusion coefficient, corresponding to the adsorption stage within the sloping voltage region. As the battery is discharged down to 0.2 V, the D band peak position begins to shift, and a new peak emerges between the D and G bands, which is likely related to side reactions such as electrolyte decomposition and solvation effects occurring during the discharge process.32 With continued discharge, the D band intensity gradually diminishes and eventually disappears, indicating the filling of defect sites in the electrode material by Na+. During charging, the newly emerged peak between the D and G bands progressively weakens, suggesting that the side reactions induced during discharge are at least partially reversible; meanwhile, the D and G band intensities show partial recovery, further confirming the reversible nature of Na+ storage behavior in the hard carbon material.
Electrochemical impedance spectroscopy (EIS) was employed to further investigate the electrochemical kinetics of CG0, CG, and CGNa0.1 electrodes (Fig. 5d). The equivalent circuit model used for fitting (Fig. S10†) includes multiple resistive and capacitive components. The high-frequency intercept represents the internal resistance (RS), encompassing the ohmic resistance of the separator, electrode, and current collector. The first semicircle corresponds to the charge transfer resistance (Rct) and the double-layer capacitance (Cdl) at the electrode/electrolyte interface. The second semicircle reflects surface film resistance (RSEI) and its associated capacitance (CSEI), related to SEI layer formation. The low-frequency Warburg element (W) represents the diffusion resistance of Na+ ions in the bulk of the active material.33 The extracted parameters (RS, Rct, and W) for all three samples are summarized in the table within Fig. 5d. Notably, CGNa0.1 exhibited the lowest RS value, indicating excellent electronic conductivity. Additionally, its Rct was also the lowest among the samples, suggesting faster charge transfer kinetics and more favorable interfacial electrochemical reactions.
The Na+ storage mechanism of CGNa0.1 was further elucidated via the galvanostatic intermittent titration technique (GITT), as shown in Fig. 5e. The Na+ diffusion coefficient (D) was calculated using the following equation.30
![]() | (2) |
To further investigate the electrochemical kinetics of Na+ in CGNa0.1 during cycling, galvanostatic charge–discharge tests were conducted at a current density of 200 mA g−1. EIS measurements were performed at the 10th, 59th, and 258th cycles, and the Na+ diffusion coefficients were calculated using eqn (3) and (4).34 The corresponding results are presented in Fig. 6. Fig. 6a presents the reversible capacity of CGNa0.1 over 258 cycles at 200 mA g−1 and 60 cycles at 500 mA g−1. Overall, the material exhibits good cycling stability, although some fluctuation appears after the EIS2 measurement. As shown in Fig. 6b–d, from the 10th to the 59th cycle, RSEI and Rct decrease significantly while RS remains relatively stable, indicating a steady activation process with minimal capacity fluctuation. However, from the 59th to the 258th cycle, Rct increases from 2.02 Ω to 3.04 Ω, and the RSEI fitting signal weakens, with both parameters showing instability—likely contributing to the increased capacity fluctuation at later stages. Nevertheless, the Na+ diffusion coefficient increases from 10−10 cm2 s−1 to 10−9 cm2 s−1 and remains stable, indicating consistently favorable Na+ diffusion kinetics throughout the cycling process.
DNa+ = 0.5[(RT/(An2F2σωC))]2 | (3) |
Z′ = Rtotal + σωω−1⁄2 | (4) |
To further assess the practical electrochemical performance of CGNa0.1, a coin-type full cell was assembled using commercially available O3-type layered sodium oxide (O3-NFM-Na) as the cathode and CGNa0.1 as the anode. As shown in Fig. 7b, a stepwise formation protocol was employed to activate both electrodes: the cell was initially charged to 3.0 V at 20 mA g−1, followed by charging to 3.4 V at 50 mA g−1, and finally to 4.0 V at 100 mA g−1. This gradual activation strategy enables full utilization of electrode materials and promotes stable cycling performance. The cycling stability of the O3-NFM-Na//CGNa0.1 full cell at 20 mA g−1 is presented in Fig. 7c. After 40 cycles, the cell maintained a stable charge capacity of 2.34 mA h, slightly exceeding the capacity recorded in the first cycle post-formation, suggesting improved electrode compatibility and enhanced sodium-ion transport kinetics following full activation. Fig. 7d illustrates the rate capability of the full cell under various current densities ranging from 20 mA g−1 to 200 mA g−1. The cell exhibited a slight capacity increase at moderate rates, followed by a gradual decline at higher current densities. Notably, when the current density was decreased from 200 mA g−1 to 40 mA g−1, the charge capacity recovered significantly—from 0.54 mA h to 1.47 mA h—demonstrating excellent rate reversibility and structural stability. These results collectively highlight the strong compatibility, kinetic resilience, and practical application potential of the O3-NFM-Na//CGNa0.1 full cell configuration for sodium-ion energy storage systems.
Adsorption site | ENa/C (eV) | EC (eV) | ENa (eV) | Ef (eV) |
---|---|---|---|---|
a ⋯ Represents a weak interaction, – represents a strong interaction. | ||||
Top site | −6855.36 | −5548.65 | −1304.09 | −2.62 |
C⋯Na–O | −6856.30 | −3.56 | ||
C–Na–O | −6856.35 | −3.61 |
The electronic characteristics of the C–Na–O structure were further analyzed, with detailed results presented in Table 3 and Fig. S11.† As shown in Fig. S11,† the O and C atoms carry Mulliken charges of −0.58e and −0.51e, respectively. Upon Na adsorption at the Top site, O and C gain 0.04e and 0.03e in Mulliken charge, respectively, suggesting that their presence influences the Na adsorption behavior at the carbon ring center. Further taking into account the interaction between C, O, and Na, in the C⋯Na–O configuration, Na donates 0.07e, O accepts 0.21e, and C donates 0.01e. Because the influence of C on Na is relatively weak, this configuration results in a net charge accumulation of 0.13e. In comparison, in the C–Na–O configuration, Na donates 0.19e, and O and C gain 0.19e and 0.07e, respectively, leading to an overall charge-neutral configuration with minimal disturbance to the surrounding environment. Moreover, to further elucidate the electronic properties of the structure, we calculated the partial density of states (PDOS) of Na, as shown in Fig. 9. The presence of pronounced free electron states near the Fermi level indicates the metallic nature of Na in the structure,35 which is particularly prominent in the C–Na–O configuration. This metallic characteristic enhances the electronic conductivity of the material,36 thereby improving its cycling stability. Therefore, the C–Na–O configuration exhibits superior electronic stability, which is expected to better promote Na storage and diffusion.
Structure | Atom | Atomic population | Total charge | Net charge | |||
---|---|---|---|---|---|---|---|
s | p | ||||||
Up | Dn | Up | Dn | ||||
Before Na adsorption | C | 0.67 | 0.63 | 1.82 | 1.39 | 4.51 | −0.51 |
O | 0.92 | 0.92 | 2.51 | 2.23 | 6.58 | −0.58 | |
C⋯Na–O | Na | 1.01 | 1.01 | 2.96 | 2.95 | 7.93 | 1.07 |
O | 0.91 | 0.91 | 2.48 | 2.49 | 6.79 | −0.79 | |
C | 0.65 | 0.62 | 1.81 | 1.42 | 4.50 | −0.50 | |
C–Na–O | Na | 1.00 | 1.00 | 2.91 | 2.90 | 7.81 | 1.19 |
C | 0.64 | 0.62 | 1.87 | 1.45 | 4.58 | −0.58 | |
O | 0.91 | 0.91 | 2.47 | 2.48 | 6.77 | −0.77 |
![]() | ||
Fig. 9 Projected densities of states (PDOS) plots of the three configurations. (a) Top site; (b) C—Na—O; (c) C—Na—O. |
The diffusion behavior of Na within the carbon matrix was categorized into three distinct modes: surface diffusion, edge diffusion, and bulk diffusion. To gain deeper insight, two representative diffusion pathways extending from the surface to the interior were defined, as illustrated in Fig. 10a. In Pathway 1, Na diffuses via the C⋯Na–O configuration, whereas in Pathway 2, diffusion occurs through the C–Na–O configuration. The corresponding energy barriers for two pathways are also presented in Fig. 10a. During surface diffusion, the energy barriers for two pathways are calculated to be 0.93 eV and 0.75 eV, respectively, suggesting that Na can spontaneously form both C⋯Na–O and C–Na–O configurations on the Carbon surface, with the latter offering a lower diffusion barrier and thus promoting more favorable surface transport. In the case of edge diffusion, both configurations encounter higher diffusion barriers, and the C–Na–O pathway, being more strongly influenced by surrounding C atom, exhibits a significantly higher energy barrier than the C⋯Na–O configuration. In comparison, bulk diffusion remains energetically unfavorable, with energy barriers of 1.71 eV in Pathway 1 and 1.75 eV in Pathway 2, indicating that disruption of the C⋯Na–O and C–Na–O structures requires external perturbation due to their high intrinsic stability.
To further evaluate the effect of Na+ doping, Na+ derived from Na2CO3 were introduced to construct C⋯Na1+—O and C–Na2+—O configurations, as shown in Fig. 10b. The promoted bulk diffusion barriers were then recalculated. As a result, the barrier in Pathway 1 decreased significantly from 1.71 eV to 1.05 eV, which can be attributed to the relatively weak influence of C on the Na1+ in the C⋯Na1+—O configuration, while the diffusing Na still experiences C interaction, resulting in a slightly higher barrier. In contrast, the C–Na2+—O structure maintains charge neutrality and causes minimal perturbation to the surrounding framework, and the diffusing Na experiences weaker C interaction, leading to an even lower diffusion barrier of only 0.11 eV in Pathway 2. To verify whether the optimized diffusion energy barrier of 0.11 eV is consistent with the experimentally fitted sodium-ion diffusion coefficients, eqn (5) was used to calculate the corresponding diffusion coefficient. The result was 1.38 × 10−10 cm2 s−1, which was in good agreement with the experimental data fitted from EIS1 (2.25 × 10−10 cm2 s−1), further reinforcing the correlation between experimental observations and theoretical calculations.
D = D0e(−((Eb/(kBT)))) | (5) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta03632a |
This journal is © The Royal Society of Chemistry 2025 |