Next Article in Journal
Scindapsus Aureus Resistive Random-Access Memory with Synaptic Plasticity and Sound Localization Function
Previous Article in Journal
Tailoring Co Distribution in PtCo Alloys for Enhanced Oxygen Reduction Reaction Activity and Durability in Fuel Cells
Previous Article in Special Issue
Review of Biomass-Derived Carbon Nanomaterials—From 0D to 3D—For Supercapacitor Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Activated Carbon from Spartina alterniflora and Its N-Doped Material for Li-Ion Battery Anode

School of Science, China University of Geosciences (Beijing), Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(9), 658; https://doi.org/10.3390/nano15090658
Submission received: 7 April 2025 / Revised: 22 April 2025 / Accepted: 23 April 2025 / Published: 26 April 2025

Abstract

:
The rampant growth of Spartina alterniflora has been wreaking havoc on the coastal ecosystems, leading to a serious environmental challenge in recent years. One potential solution to this issue involves converting Spartina alterniflora into activated carbon, offering a potential remedy for pollution while creating value in energy storage applications. Herein, through a facile carbonization process with sodium hydroxide activation, we successfully transformed obsolete Spartina alterniflora into a porous carbon material (called SAC) and its nitrogen-doped derivative (denoted as SANC) by using melamine as the nitrogen source in a similar procedure. The amorphous structure of these materials was confirmed to enhance lithium-ion storage and electrolyte permeation, making them ideal for use as anodes in lithium-ion batteries. As a result, both SAC and SANC, derived from Spartina alterniflora, exhibited outstanding electrochemical performance including high capacity (456.7 and 780.8 mA h g−1 for SAC and SANC, respectively, at the current density of 6 mA g−1), excellent rate performance (from 6 to 600 mA g−1) and long-term cycling stability. Notably, compared to SAC, its N-doped derivative SANC showed superior properties in the battery (retaining a reversible capacity of 412.9 mA h g−1 at the current density of 6 mA g−1 even after 600 repeated charge–discharge cycles), demonstrating the significantly positive impact of heteroatom doping. This work not only offers a strategy to mitigate environmental challenges but also demonstrates the potential for converting waste biomass into a valuable resource for energy storage applications.

1. Introduction

As one of the advanced electrochemical energy storage devices, lithium-ion (Li-ion) batteries have been widely used in modern applications including smart mobile machines, electric devices and vehicles [1,2]. Traditionally, graphite was most applied as anode material due to its excellent conductivity and perfect cycling stability, with a theoretical capacity of 372 mA h g−1. Typically, graphite is a crucial and expensive raw carbon either extracted or artificially manufactured from petroleum coke raw materials [3]. Therefore, it is imperative to enhance new electrode materials to promote the energy storage performance of Li-ion batteries and meet the increasing energy demands.
Nowadays, biomass has been receiving increasing attention in energy storage and conversion due to its advantages such as affordability, accessibility, rapid regeneration and eco-friendliness [4,5,6]. Carbon materials derived from biomass usually exhibit micro-nanostructures, large specific surface areas and structural stability, which can facilitate the reversible storage of Li-ion and shorten their transport pathways [7,8,9]. Moreover, heteroatom doping (e.g., B, N, P and S) can regulate the active sites for enhancing the interfacial reactivity and conductivity of derived carbon materials, thereby improving the electrochemical performance [10,11,12,13]. Various biomass sources have been used as precursors for carbonaceous materials and their derivatives, serving as anodes in Li-ion batteries (Table S1) [14,15]. Xiong et al. [16] explored S and N co-doped carbon using waste cotton balls through a simple pyrolysis method and used it as anode material in high-performance Li-ion battery. By a carbonization procedure, Sopon Butcha et al. [17] converted harmful floating plant water hyacinth to useful anode material with high specific capacity and perfect rate performance. Agricultural residues like corn straw [18] and sugarcane bagasse [19] have also been used as high specific capacitance electrodes and showed great potential in energy storage applications. In recent years, the widespread reproduction of Spartina alterniflora brought significant environmental pollution, serious ecological impacts and economic losses. Due to its abundant cellulose, lignin and pore structure [20,21], Spartina alterniflora can benefit from the storage of Li-ion and should be a potential anode material in Li-ion batteries.
In this paper, a porous carbon material (denoted as SAC) was successfully prepared from the obsolete Spartina alterniflora through a pyrolyzation method with a straightforward activation process using sodium hydroxide (as shown in Scheme 1). After post-doping with melamine as the nitrogen source, its N-doped activated derivative (denoted as SANC) was also obtained (Scheme S1). Various characterization techniques were employed to characterize the composition, porous structure and morphology of these carbon materials, including X-ray diffraction (XRD), Raman spectroscopy, Brunauer–Emmett–Teller (BET) measurements, X-ray photoelectron spectroscopy (XPS), scanning electron microscope (SEM), transmission electron microscope (TEM) and energy dispersive spectroscopy (EDS). As the anode materials for Li-ion batteries, SAC and SANC originated from Spartina alterniflora showed excellent electrochemical performance including high capacity (456.7 and 780.8 mA h g−1 for SAC and SANC, respectively, at 6 mA g−1), good rate performance (from 6 to 600 mA g−1), as well as good cycling stability (265.6 and 412.9 mA h g−1 after 600 cycles for SAC and SANC, respectively). Compared to SAC, its N-doped activated derivative SANC showed more perfect properties in Li-ion battery, demonstrating the important impact of heteroatom doping. This work provides an approach to retard the environmental challenges that originate from wreaking havoc on invasive species. More importantly, it can make the waste into a profitable resource and benefit the field of energy storage.

2. Experimental

2.1. Material Synthesis

The Spartina alterniflora was reaped from the east coast of China. After thoroughly washed with distilled water and ethanol, the chopped Spartina alterniflora was grinded into powder using a high-speed multifunctional grinder. Then, the Spartina alterniflora powder was pre-carbonized in a tube furnace under an argon atmosphere. After the pre-carbonization, the product was mixed with NaOH in a 1:1 mass ratio and thoroughly ground with a mortar. 600 °C was applied to activate the material and then HCl (0.1 mol L−1) was used to regulate the Ph until neutral. After the filtration and drying process, SAC material was successfully obtained. To prepare the N-doped activated derivative SANC material, melamine was used as the nitrogen source. It was first mixed with Spartina alterniflora powder in a 1:1 mass ratio during the pre-carbonization process. The subsequent treatment steps followed the similar prepared procedure for SAC.

2.2. Characterization

XRD spectra were measured by Empyrean diffractometer (Cu Kα radiation, with an output power of 1.6 kW at a voltage of 40 kV) (Malvern Panalytical, Malvern, UK). Renishaw-2000 Raman spectrometer was used to measure Raman spectra with an Ar laser at 473 nm (Wotton-under-Edge, UK). Thermo Scientific ESCALab 250Xi apparatus using 200W mono-chromated Al Kα radiation (Boston, MA, USA) was used to measure XPS spectra. BET (ASAP2020) (Belsorp-max, Osaka, Japan) N2 adsorption–desorption tests were carried out at 77 K to measure the specific surface area and pore information of the materials. SEM images, elemental distributions and TEM images were measured by Hitachi Model S-4800 field emission scanning electron microscope and HT-7700 electron microscope (an accelerating voltage of 100 kV) (Ibaraki, Japan).

2.3. Electrochemical Testing

The prepared SAC and SANC materials (90%) were fully mixed with carbon black (5%) and polyvinylidene fluoride (PVDF, 5%) with the N-methyl-2-pyrrolidinone (NMP) into a slurry mixture. After placed on the Cu foil, the slurry mixture was dried in a vacuum oven. Before the fabrication, the foil was slowly cut into small pieces with a diameter of 12 mm. Then, the CR2032 button batteries were assembled in an argon-filled glove box (in which the O2/H2O content was below 0.01 ppm), with opposite lithium metal electrode, Celgard 2400 PP (Celgard, Charlotte, NC, USA) polypropylene separator and EC/DMC = 1:1 liquid electrolyte with 1.0 M LiPF6. The average mass of loading active SAC/SANC was 1 mg cm−2. For the battery electrochemical performance, the CT2001A (Wuhan, China) land battery system was used in the voltage range from 5 mV to 3 V. The cyclic voltammograms and the electrochemical impedance spectroscopy were measured on CHI760E (Shanghai, China) electrochemical workstation (scan rate of 0.5 mV s−1, frequency range 0.01 Hz to 100 kHz).

3. Results and Discussion

The formation of carbonaceous derivatives from Spartina alterniflora was confirmed by structural characterizations using XRD, Raman, BET and XPS. The XRD patterns of SAC and SANC in Figure 1a showed a peak at around 22° ((002) plane of graphite carbon) [22]. The broad diffraction patterns with lower intensity also demonstrated their amorphous nature and limited graphitization. The partial graphitization of carbon may benefit the electrochemical performance as an anode by effectively increasing the reversible capacity of Li-ion intercalation [23]. Increasing the spacing between the layers of carbon material can enhance Li-ion transfer efficiency, thereby improving the cycling and rate performance of Li-ion batteries. Notably, the (002) diffraction peak of the N-doped SANC material (22.4°) exhibited a higher 2θ value compared to SAC (21.8°). This was due to the active nitrogen atoms, which can efficiently promote the transportation of lithium within the pores and increase the disordered property of the biomass structure [24]. Raman spectroscopy is a useful technology to evaluate the graphitization degree [25]. As shown in Figure 1b, these two carbon materials showed similar characteristic peaks at 1380 cm−1 (D band, representing the defects of graphitic carbon) and 1595 cm−1 (G band, representing the degree of graphitization). The intensity ratio (ID/IG) is usually utilized to demonstrate the degree of amorphous structure in the prepared carbon material [26,27]. As a result, the ID/IG values were calculated as 0.919 for SAC and 0.929 for SANC, respectively, indicating that nitrogen-doping introduced more structural defects and increased the amorphous nature. Compared to the ordered carbon material, the defects, and/or disordered structure are beneficial for enhancing the active sites, thus potentially resulting in the improved electrochemical performance when employed as the anodes in Li-ion batteries.
To detect the specific surface areas and pore size distributions of SAC and SANC, N2 adsorption/desorption tests were measured. As shown in Figure 1c,d, the two carbon materials exhibited a characteristic type I adsorption–desorption isotherms and a distinct hysteresis loop at relative high pressure, demonstrating their microporous structures of the biomass carbons from Spartina alterniflora. The corresponding specific surface areas were calculated to be 569.148 m2 g−1 for SAC and 657.897 m2 g−1 for SANC, respectively. The pore size distribution curves in Figure 1c,d were used to confirm the microporous features. The microporous diameters mainly distributed between 1 and 3 nm, which is beneficial to the enhanced Li-ion storage and facilitates the permeation of electrolyte [28]. Based on the BET results, it can be predicted that the N-doped derivative SANC may have more superior performance as anode electrodes due to its higher specific surface area and abundant active defects, which can increase the storage of lithium.
The surface chemical composition of the carbon materials SAC and SANC were realized by XPS measurements. As can be seen in Figure 2a, in the XPS survey spectrum of SAC, C (85.01%) and O (14.1%) can be observed. In SANC, the contents were 80.82% for C (284.4 eV), 4.22% for N (400.6 eV) and 14.95% for O (533.2 eV), respectively. This phenomenon indicated that N was successfully doped into the activated carbon SAC. Figure 2b–f and Table 1 displayed the high-resolution spectrum of C1s, N1s and O1s, respectively. As can be seen in the C1s spectrum of SAC in Figure 2b, there were four peaks at 283.62 (C−C bond), 284.29 (C=C bond), 286.41 (C−O bond) and 288.62 eV (O=C−OH bond), respectively [29]. In the high-resolution O1s spectrum in Figure 2c, three peaks at 531.06 (C=O bond), 532.08 (C−OH/C−O−C bond) and 533.22 eV (O=C−OH bond) can be observed [30]. For the N-doped activated derivative SANC material, the C1s spectrum can be deconvoluted into five peaks at 284.29 eV (C−C bond), 283.62 eV (C=C bond), 285.27 eV (C−N bond), 286.41 eV (C−O bond) and 288.62 eV (C=O bond), respectively. There were three peaks in the high-resolution N1s spectrum at 398.42 (pyridinic nitrogen), 399.29 (pyrrolic nitrogen) and 399.83 eV (graphitic nitrogen), respectively [31]. In the high-resolution O1s spectrum, there were three characteristic peaks at 531.06 (C=O bond), 532.08 (C−OH/C−O−C bond) and 533.22 eV (O=C−OH bond), respectively. According to the reported literature [32,33], the doping of nitrogen can promote the formation of functional groups and increase active defects for Li-ion storage, thus benefiting to the enhancement of the electrochemical performance.
Figure 3, Figures S1 and S2 showed the morphologies of the two carbon materials derived from Spartina alterniflora, including SEM, TEM and EDS. The activated carbon materials SAC and its N-doped derivative SANC all present irregular carbon blocks (Figure 3a,c), demonstrating their amorphous nature, consistent with the XRD results. As shown in SEM images, there were many pore structures in the bulk carbons, which also can be confirmed by the BET results. The TEM images of SAC and SANC were exhibited in Figure 3b,d. It was obvious that there were no visible lattice fringes, which was attributed to their limited graphitization degree as well as the amorphous characteristics. Compared to the SAC, after the successful N-doping by melamine, the N was uniformly distributed in SANC (Figure S1). Heteroatom doping in carbon materials is an efficient approach to induce plenty of active sites and defects [34,35]. Consequently, the N-doping would have a positive effect to regulate the electronic and chemical structure and enhance the electrochemical reactivity and performance.
To evaluate their potential in energy storage, the two materials SAC and SANC from Spartina alterniflora were used as the lithium storage material in the Li-ion battery. The electrochemical properties of SAC and SANC were first measured by cyclic voltammetry (CV) (0.01-3.0 V, 0.5 mV s−1). Figure 4a showed the CV of the SAC electrode in the initial five cycles. Firstly, there were two small cathode peaks, which disappeared in the subsequent cycles. This phenomenon can be attributed to some irreversible reactions related to the electrolyte decomposition on the surface of the electrode and the formation of the solid electrolyte interphase (SEI) film [36,37]. This is also the main reason for the low Coulomb efficiency (44%) in the initial process. In the following measurements, CV curves almost overlapped, indicating the good structural stability of the electrode. For the SANC electrode, the CV just presented similar curves compared to the SAC material, indicating that there was no significant difference between the structure of SAC and its N-doped derivative SANC.
Figure 4c,d showed the constant charge–discharge processes of the two materials at a current density of 6 mA g−1 from 0.01 to 3 V. The first charge specific capacity of the SANC sample (Figure 4d) can reach 776.3 mA h g−1, and the initial Coulomb efficiency is 41%, lower than that of SAC. During the first cycle near 1.6 V, there is a small platform, which was attributed to the formation of SEI and the degradation of the electrolyte [38]. The specific capacity can maintain 399.5 mA h g−1 even after 300 repeated charge–discharge cycles, which was larger than the carbon material SAC (290.5 mA h g−1) before N-doping. Rate capabilities of the two electrodes were evaluated at different current densities ranging from 6 to 600 mA g−1. In Figure 4e, the N-doped SANC electrode showed reversible capacities of 780.8, 580.6, 459.7, 372.5, 305.2, 260.1 and 212.3 mA h g−1 at 6, 15, 30, 60, 150, 300 and 600 mA g−1, respectively, higher than that of SAC electrode (456.7, 421.2, 326.5, 273.5, 229.9, 158.3 and 117.8 mA h g−1 at the same current densities), demonstrating more excellent Li-ion storage ability and rate performance of SANC than SAC. When the current density returned to 6 mA g−1, 445.0 and 588.4 mA h g−1 were obtained for SAC and SANC. After N-doping, the cycling performance was also improved. At the current density of 6 mA g−1 in Figure 4f and Figure S3, the SANC electrode exhibited excellent cycling stability. Even after 600 repeated charge–discharge cycles, the battery can retain a capacity of 412.9 mA h g−1. In contrast, the SAC electrode achieved a reversible capacity of 265.6 mA h g−1 under the same condition, lower than that of SANC electrode. The SANC electrode also showed superior cycle stability with a reversible capacity at a higher current density of 600 mA g−1 (Figure S4). Owing to the environment interference and the internal polarization during long-term tests, for the rate performance and the cycle stability, there were slight fluctuations in the capacities.
To determine the Li-ion storage processes of the two materials and the electrochemical kinetic behaviors in Li-ion batteries, CV curves at different scanning rates (v) ranging from 0.5 to 5 mV s−1 were measured for SAC and SANC (Figure 5). As shown in Figure 5a,c, with the increase in scanning speed, the current density enhanced and the peak potential shifted, demonstrating that the intercalation and de-intercalation processes of lithium in SAC and SANC were reversible. According to the relationship between the measured current (I) and the scanning rate (v), I = avb [39], the b value of the SAC electrode was 0.59, more approximately to 0.5. This indicated that the electrochemical reaction of SAC anode in the Li-ion battery was mainly a diffusion-controlled process. Diffusion control implies a slower Li-ion transmission rate, which becomes the limiting step in the electrochemical reaction. When the b value is closer to 1, it suggests a mainly capacitance-controlled electrochemical reaction. Capacitance control indicates a faster charge transfer rate on the electrode surface, which is the primary factor influencing the electrochemical reaction rate. As a result, after N-doping, the faster surface charge transfer rate of the SANC electrode (with b value of 0.77), along with rapid adsorption and desorption processes, contributed to the improved rate performance.
To further explore the impact of N-doping in SAC on the electrochemical process of Li-ion storage, the electrochemical impedance spectroscopy (EIS) was investigated. As shown in Figure S5, the initial electrochemical impedance spectra of SAC and SANC showed that in the high-frequency region, the Nyquist plots of SAC and SANC electrodes exhibited semi-circular shapes, while in the low-frequency region, they all displayed sloping lines. In Figure S6, after 100 charge–discharge repeated cycles, the interface resistance was almost the same to the initial electrochemical impedance, indicating that during the charging and discharging process, a stable SEI film had been formed, which could maintain the structure of the material. In general, the semi-circle in the high-frequency region represents the charge transfer resistance, while the sloping line in the low-frequency region represents the impedance related to ion diffusion processes in anode material [40,41]. Compared to the SAC electrode, the diameter of the semi-circle in the SANC electrode was reduced, indicating that after N-doping into the SAC material, a decreased charge transfer resistance at the electrode/electrolyte interface was obtained, which can contribute to the improved electrochemical performance of the SANC.

4. Conclusions

In conclusion, activated carbon derived from Spartina alterniflora as well as its N-doped material were prepared using a simple strategy. Benefiting from their porous structures, the storage of lithium and the permeation of electrolytes would be promoted and the diffusion distance of lithium ions would be reduced. Especially, in the N-doped material, more defects were produced and can be considered as active sites. Consequently, compared to the activated carbon from Spartina alterniflora, its N-doped carbon derivative showed superior electrochemical performance when utilized as anode material of Li-ion battery. Even after 600 repeated charge-discharge cycles, the N-doped material SANC can still provide a reversible high capacity of 412.9 mA h g−1. The exceptional electrochemical performance of the two materials in Li-ion batteries presented here offers a cost-effective and simple solution to address the encroachment issue posed by Spartina alterniflora species and can effectively turn waste into a valuable resource and benefit energy storage and conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15090658/s1. Scheme S1. The schematic diagram of the synthesis procedure of SANC. Figure S1. Elemental distribution of SANC material. Figure S2. Elemental distribution of SAC material. Figure S3. Coulombic efficiency of SAC and of SANC electrodes. Figure S4. Cycling performance of SAC and SANC electrode at 600 mA g−1. Figure S5. The initial electrochemical impedance spectra and corresponding equivalent circuit of SAC and SANC. Figure S6. Electrochemical impedance spectra of SAC and SANC after 100 charge-discharge repeated cycles. Table S1. Performance comparison of biochar carbon as anodes in LIBs [42,43,44,45,46].

Author Contributions

The first draft of the manuscript was written by H.S. and X.H., and all authors commented on the previous versions of the manuscript. Materials preparation, data collection and analysis were performed by H.S., X.H., Y.Z., J.P., L.G. and H.L. Resources and supervision were provided by H.S. and B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Beijing Young Elite Scientist Sponsorship Program by Bast (BYESS24370) and the National Natural Science Foundation of China (22001240).

Data Availability Statement

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

We thank Peng Ju and Jianchao Sun for providing the biomass Spartina alterniflora. The large instruments and equipment sharing platform of China University of Geosciences (Beijing) is gratefully acknowledged.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Choi, S.; Wang, G. Advanced lithium-ion batteries for practical applications: Technology, development, and future perspectives. Adv. Mater. Technol. 2018, 3, 1870034. [Google Scholar] [CrossRef]
  2. Kim, T.; Song, W.; Son, D.; Ono, L.K.; Qi, Y. Lithium-ion batteries: Outlook on present, future, and hybridized technologies. J. Mater. Chem. A 2019, 7, 2942–2964. [Google Scholar] [CrossRef]
  3. Li, X.; Deng, C.; Liu, M.; Xiong, J.; Zhang, X.; Yan, Q.; Lin, J.; Chen, C.; Wu, F.; Zhao, Y.; et al. Reutilization and upcycling of spent graphite for sustainable lithium-ion batteries: Progress and perspectives. eScience 2025, 100394. [Google Scholar] [CrossRef]
  4. Liu, W.; Jiang, H.; Yu, H. Emerging applications of biochar-based materials for energy storage and conversion. Energy Environ. Sci. 2019, 12, 1751–1779. [Google Scholar] [CrossRef]
  5. Mishra, R.K.; Mohanty, K. A review of the next-generation biochar production from waste biomass for material applications. Sci. Total Environ. 2023, 15, 167171. [Google Scholar] [CrossRef] [PubMed]
  6. Zhu, Z.; Men, Y.; Zhang, W.; Yang, W.; Wang, F.; Zhang, Y.; Zhang, Y.; Zeng, X.; Xiao, J.; Tang, C.; et al. Versatile carbon-based materials from biomass for advanced electrochemical energy storage systems. eScience 2024, 4, 100249. [Google Scholar] [CrossRef]
  7. Jin, C.; Nai, J.; Sheng, O.; Yuan, H.; Zhang, W.; Tao, X.; Lou, X.W. Biomass-based materials for green lithium secondary batteries. Energy Environ. Sci. 2021, 14, 1326. [Google Scholar] [CrossRef]
  8. Mohanty, A.K.; Vivekanandhan, S.; Das, O.; Millán, L.M.R.; Klinghoffer, N.B.; Nzihou, A.; Misra, M. Biocarbon materials. Nat. Rev. Methods Primers 2024, 4, 19. [Google Scholar] [CrossRef]
  9. Liedel, C. Sustainable battery materials from biomass. ChemSusChem 2020, 13, 2110–2141. [Google Scholar] [CrossRef]
  10. Sekhon, S.S.; Park, J. Biomass-derived N-doped porous carbon nanosheets for energy technologies. Chem. Eng. J. 2021, 425, 129017. [Google Scholar] [CrossRef]
  11. Ariharan, A.; Ramesh, K.; Vinayagamoorthi, R.; Rani, M.S.; Viswanathan, B.; Ramaprabhu, S.; Nandhakumar, V. Biomass derived phosphorous containing porous carbon material for hydrogen storage and high-performance supercapacitor applications. J. Energy Storage 2021, 35, 102185. [Google Scholar] [CrossRef]
  12. Ma, L.; Hu, X.; Liu, W.; Li, H.; Lam, P.K.S.; Zeng, R.J.; Yu, H. Constructing N, P-dually doped biochar materials from biomass wastes for high-performance bifunctional oxygen electrocatalysts. Chemosphere 2021, 278, 130508. [Google Scholar] [CrossRef]
  13. Aristote, N.T.; Song, Z.; Deng, W.; Hou, H.; Zou, G.; Ji, X. Effect of double and triple-doping of sulfur, nitrogen and phosphorus on the initial coulombic efficiency and rate performance of the biomass derived hard carbon as anode for sodium-ion batteries. J. Power Sources 2023, 558, 232517. [Google Scholar] [CrossRef]
  14. Li, J.; Deng, W.; Li, H.; Chen, L.; Zhang, Y.; Li, J.; Song, Y.; Duan, H. Biomass-derived N-P double-doped porous carbon spheres and their lithium storage mechanism. Int. J. Hydrogen Energy 2024, 56, 828–836. [Google Scholar] [CrossRef]
  15. Wan, H.; Hu, X. Nitrogen doped biomass-derived porous carbon as anode materials of lithium ion batteries. Solid State Ionics 2019, 341, 115030. [Google Scholar] [CrossRef]
  16. Xiong, J.; Pan, Q.; Zheng, F.; Xiong, X.; Yang, C.; Hu, D.; Huang, C. N/S Co-doped carbon derived from cotton as high performance anode materials for lithium ion batteries. Front. Chem. 2018, 6, 78. [Google Scholar] [CrossRef]
  17. Butcha, S.; Rajrujithong, C.; Sattayarut, V.; Youngjan, S.; Nakajima, H.; Supruangnet, R.; Wittayakun, J.; Prayoonpokarach, S.; Faungnawakij, K.; Chanthad, C.; et al. Sustainable production of multifunctional hierarchical carbon from weed water hyacinth: Assessment for lithium-ion battery and supercapacitor. J. Energy Storage 2023, 72, 108578. [Google Scholar] [CrossRef]
  18. Yu, K.; Wang, J.; Wang, X.; Liang, J.; Liang, C. Sustainable application of biomass by-products: Corn straw-derived porous carbon nanospheres using as anode materials for lithium ion batteries. Mater. Chem. Phys. 2020, 243, 122644. [Google Scholar] [CrossRef]
  19. Zheng, S.; Luo, Y.; Zhang, K.; Liu, H.; Hu, G.; Qin, A. Nitrogen and phosphorus Co-doped mesoporous carbon nanosheets derived from bagasse for lithium-ion batteries. Mater. Lett. 2021, 290, 129459. [Google Scholar] [CrossRef]
  20. Cheng, H.; Tang, Z.; Luo, X.; Zheng, Z. Spartina alterniflora-derived porous carbon using as anode material for sodium-ion battery. Sci. Total Environ. 2021, 777, 146120. [Google Scholar] [CrossRef]
  21. Xia, L.; Yang, W.; Geng, Q.; Jeelani, N.; An, S. Research on spartina alterniflora using molecular biological techniques: An overview. Mar. Freshwater Res. 2020, 71, 1564–1571. [Google Scholar] [CrossRef]
  22. Xu, X.; Wu, F.; Yang, W.; Dai, X.; Wang, T.; Zhou, J.; Wang, J.; Guo, D. Silicon@natural nitrogen-doped biomass carbon composites derived from “silicon Tofu” as green and efficient anode materials for lithium-ion batteries. ACS Sustain. Chem. Eng. 2021, 9, 13215–13224. [Google Scholar] [CrossRef]
  23. Chen, X.; Zhang, Z.; Xu, R.; Gao, X.; Zhou, D.; Yuan, T.; Chen, Y.; Cui, L. In-situ embedding of ultrasmall MnO nanoparticles into pollen biomass-derived hollow porous N heteroatoms enriched carbon microspheres as high-performance anode for lithium-ion batteries. Chem. Eng. J. 2023, 473, 145117. [Google Scholar] [CrossRef]
  24. Bartoli, M.; Piovano, A.; Elia, G.A.; Meligrana, G.; Pedraza, R.; Pianta, N.; Tealdi, C.; Pagot, G.; Negro, E.; Triolo, C.; et al. Pristine and engineered biochar as Na-ion batteries anode material: A comprehensive overview. Renew. Sustain. Energy Rev. 2024, 194, 114304. [Google Scholar] [CrossRef]
  25. Yang, L.; Lei, Y.; Liang, X.; Qu, L.; Xu, K.; Hua, Y.; Feng, J. SnO2 nanoparticles composited with biomass N-doped carbon microspheres as low cost, environmentally friendly and high-performance anode material for sodium-ion and lithium-ion batteries. J. Power Sources 2022, 541, 232032. [Google Scholar] [CrossRef]
  26. Zeng, D.; Xiong, H.; Qi, K.; Guo, X.; Qiu, Y. Constructing N-doping biomass-derived carbon with hierarchically porous architecture to boost fast reaction kinetics for high-performance lithium storage. J. Colloid Interface Sci. 2021, 605, 741–751. [Google Scholar] [CrossRef]
  27. Yang, J.; Chen, Z.; Lin, H.; Feng, P.; Xie, Y.; Liang, Y.; Zheng, M.; Liu, X.; Liu, Y.; Xiao, Y. Exfoliating waste biomass into porous carbon with multi-structural levels for dual energy storage. ACS Appl. Energy Mater. 2022, 5, 12090–12098. [Google Scholar] [CrossRef]
  28. Duan, Y.; Li, Z.; Zhang, S.; Su, T.; Ma, Y.; Jiao, A.; Fu, Z. Metal-organic frameworks (MOFs)-derived Mn2SnO4@C anode based on dual lithium storage mechanism for high-performance lithium-ion capacitors. Chem. Eng. J. 2023, 477, 146914. [Google Scholar] [CrossRef]
  29. Sobaszek, M.; Brzhezinskaya, M.; Olejnik, A.; Mortet, V.; Alam, M.; Sawczak, M.; Ficek, M.; Gazda, M.; Weiss, Z.; Bogdanowicz, R. Highly occupied surface states at deuterium-grown boron-doped diamond interfaces for efficient photoelectrochemistry. Small 2023, 19, 2208265. [Google Scholar] [CrossRef]
  30. Plavniece, A.; Kaare, K.; Simkunaite, D.; Balciunaite, A.; Jasulaitiene, V.; Niaura, G.; Volperts, A.; Dobele, G.; Colmenares-Rausseo, L.; Kruusenberg, I.; et al. Manganese- and nitrogen-doped biomass-based carbons as catalysts for the oxygen reduction reaction. Catalysts 2024, 14, 92. [Google Scholar] [CrossRef]
  31. Brzhezinskaya, M.; Mishakov, I.V.; Bauman, Y.I.; Shubin, Y.V.; Maksimova, T.A.; Stoyanovskii, V.O.; Gerasimov, E.Y.; Vedyagin, A.A. One-pot functionalization of catalytically derived carbon nanostructures with heteroatoms for toxic-free environment. Appl. Surf. Sci. 2022, 590, 153055. [Google Scholar] [CrossRef]
  32. Chen, Y.; Wu, Y.; Liao, Y.; Zhang, Z.; Luo, S.; Li, L.; Wu, Y.; Qing, Y. Tuning carbonized wood fiber via sacrificial template-assisted hydrothermal synthesis for high-performance lithium/sodium-ion batteries. J. Power Sources 2022, 546, 231993. [Google Scholar] [CrossRef]
  33. Sun, C.; Pan, J.; Zhao, X.; Jiao, C.; Yao, W.; Wang, C.; Fu, X.; Ma, D.; Xue, H.; Liu, J.; et al. Effective coating of Si@NiO nanoflowers with nitrogen-doped wheat protein-derived biochar for efficient lithium-ion and lithium-sulfur batteries anode materials. J. Alloys Compd. 2023, 968, 171922. [Google Scholar] [CrossRef]
  34. Li, H.; Gong, Y.; Fu, C.; Zhou, H.; Yang, W.; Guo, M.; Li, M.; Kuang, Y. A novel method to prepare a nanotubes@mesoporous carbon composite material based on waste biomass and its electrochemical performance. J. Mater. Chem. A 2017, 5, 3875–3887. [Google Scholar] [CrossRef]
  35. Schneidermann, C.; Kensy, C.; Otto, P.; Oswald, S.; Giebeler, L.; Leistenschneider, D.; Grätz, S.; Dörfler, S.; Kaskel, S.; Borchardt, L. Nitrogen-doped biomass-derived carbon formed by mechanochemical synthesis for lithium-sulfur batteries. ChemSusChem 2018, 12, 310–319. [Google Scholar] [CrossRef] [PubMed]
  36. Sun, X.; Gao, X.; Li, Z.; Zhang, X.; Zhai, X.; Zhang, Q.; Li, L.; Gao, N.; He, G.; Li, H. Nanowires framework supported porous lotus-carbon anode boosts lithium-ion and sodium-ion batteries. Small Methods 2023, 8, 2300746. [Google Scholar] [CrossRef] [PubMed]
  37. Sheng, O.; Jin, C.; Yang, T.; Ju, Z.; Luo, J.; Tao, X. Designing biomass-integrated solid polymer electrolytes for safe and energy-dense lithium metal batteries. Energy Environ. Sci. 2023, 16, 2804–2824. [Google Scholar] [CrossRef]
  38. Sun, D.; Yang, J.; Yan, X. Hierarchically porous and nitrogen, sulfur-codoped graphene-like microspheres as a high capacity anode for lithium ion batteries. Chem. Commun. 2015, 51, 2134–2137. [Google Scholar] [CrossRef]
  39. Tao, L.; Liu, L.; Chang, R.F.; He, H.B.; Zhao, P.; Liu, J. Structural and interface design of hierarchical porous carbon derived from soybeans as anode materials for potassium-ion batteries. J. Power Sources 2020, 463, 228172. [Google Scholar] [CrossRef]
  40. Jiang, S.; Wang, R.; Pang, M.; Wang, H.; Zeng, S.; Yue, X.; Ni, L.; Qiu, S.; Zhang, Z. Hierarchical composites of ultrathin carbon self-coated TiO2 nanosheets on reduced graphene oxide with enhanced lithium storage capability. Chem. Eng. J. 2015, 280, 614–622. [Google Scholar] [CrossRef]
  41. Zhu, H.; Li, J.; Wu, D.; Zhang, G.; Sun, Y.; Wang, A.; Sun, K. Construction of the hierarchical porous biochar with an ultrahigh specific surface area for application in high-performance lithium-ion capacitor cathode. Biochar 2023, 5, 46. [Google Scholar] [CrossRef]
  42. Bhatia, D.; Saroha, A.K. Rice straw-derived biochar as anode material for lithium-ion battery: A step toward sustainable development. Energy Technol. 2024, 12, 6. [Google Scholar] [CrossRef]
  43. Shi, Z.Y.; Jin, Y.H.; Han, T.; Yang, H.M.; Gond, R.; Subasi, Y.; Asfaw, H.D.; Younesi, R.; Jönsson, P.G.; Yang, W.H. Bio-based anode material production for lithium-ion batteries through catalytic graphitization of biochar: The deployment of hybrid catalysts. Sci. Rep. 2024, 14, 3966. [Google Scholar] [CrossRef] [PubMed]
  44. Belmesov, A.A.; Glukhov, A.A.; Kayumov, R.R.; Podlesniy, D.N.; Latkovskaya, E.M.; Repina, M.A.; Ivanov, N.P.; Tsvetkov, M.V.; Shichalin, O.O. Using aquatic plant-derived biochars as carbon materials for the negative electrodes of Li-ion batteries. Coatings 2023, 13, 2075. [Google Scholar] [CrossRef]
  45. Wang, Y.; Chang, H.Y.; Ma, T.; Deng, H.; Zha, Z.H. Effect of cotton stalk particle size on the structure of biochar and the performance of anode for lithium-ion battery. J. Phys. Chem. Solids 2022, 169, 110845. [Google Scholar] [CrossRef]
  46. Jiang, R.T.; Yang, G. Biomass-derived N-doping porous carbon spheres by N2 plasma for lithium-ion batteries. Biomass Convers. Biorefin. 2024. [Google Scholar] [CrossRef]
Scheme 1. The schematic diagram of the preparation of SAC originated from Spartina alterniflora.
Scheme 1. The schematic diagram of the preparation of SAC originated from Spartina alterniflora.
Nanomaterials 15 00658 sch001
Figure 1. (a) XRD patterns; (b) Raman spectra; (c,d) N2 adsorption–desorption isotherms of SAC and SANC (inset: pore size distribution and pore volume).
Figure 1. (a) XRD patterns; (b) Raman spectra; (c,d) N2 adsorption–desorption isotherms of SAC and SANC (inset: pore size distribution and pore volume).
Nanomaterials 15 00658 g001
Figure 2. (a) XPS spectra; (b,c) high-resolution C 1 s and O 1 s spectra of SAC; (df) high-resolution C 1 s, N1s and O 1 s spectra of SANC.
Figure 2. (a) XPS spectra; (b,c) high-resolution C 1 s and O 1 s spectra of SAC; (df) high-resolution C 1 s, N1s and O 1 s spectra of SANC.
Nanomaterials 15 00658 g002
Figure 3. SEM images and TEM images of SAC (a,b) and SANC (c,d).
Figure 3. SEM images and TEM images of SAC (a,b) and SANC (c,d).
Nanomaterials 15 00658 g003
Figure 4. (a) Cyclic voltammetry curves of SAC electrode (0.5 mV·s−1); (b) cyclic voltammetry curves of SANC electrode (0.5 mV·s−1); (c,d) voltage-capacity curves of SAC and SANC electrodes at 6 mA g−1; (e) rate performance of SAC and SANC electrodes at different current densities; (f) cycling performance of SAC and SANC electrodes (1C = 300 mA g−1).
Figure 4. (a) Cyclic voltammetry curves of SAC electrode (0.5 mV·s−1); (b) cyclic voltammetry curves of SANC electrode (0.5 mV·s−1); (c,d) voltage-capacity curves of SAC and SANC electrodes at 6 mA g−1; (e) rate performance of SAC and SANC electrodes at different current densities; (f) cycling performance of SAC and SANC electrodes (1C = 300 mA g−1).
Nanomaterials 15 00658 g004
Figure 5. (a) CV curves at different scanning rates for SAC; (b) determination of the b-value using the relationship between peak current and scan rate for SAC; (c) CV curves at different scanning rates for SANC; (d) determination of the b-value using the relationship between peak current and scan rate for SANC.
Figure 5. (a) CV curves at different scanning rates for SAC; (b) determination of the b-value using the relationship between peak current and scan rate for SAC; (c) CV curves at different scanning rates for SANC; (d) determination of the b-value using the relationship between peak current and scan rate for SANC.
Nanomaterials 15 00658 g005
Table 1. XPS C1s analysis of the elemental composition of SANC and SAC.
Table 1. XPS C1s analysis of the elemental composition of SANC and SAC.
SamplePeaks and Their Positions (eV)
C=CC−CC−NC−OO=C−OH
283.62284.29285.27286.41288.62
SAC18.1%45.0%14.9%12.8%9.2%
SANC17.6%57.9% 13.14%11.34%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shang, H.; Hao, X.; Zhou, Y.; Peng, J.; Guo, L.; Li, H.; Sun, B. Activated Carbon from Spartina alterniflora and Its N-Doped Material for Li-Ion Battery Anode. Nanomaterials 2025, 15, 658. https://doi.org/10.3390/nano15090658

AMA Style

Shang H, Hao X, Zhou Y, Peng J, Guo L, Li H, Sun B. Activated Carbon from Spartina alterniflora and Its N-Doped Material for Li-Ion Battery Anode. Nanomaterials. 2025; 15(9):658. https://doi.org/10.3390/nano15090658

Chicago/Turabian Style

Shang, Hong, Xinmeng Hao, Yougui Zhou, Jia Peng, Lihua Guo, Huipeng Li, and Bing Sun. 2025. "Activated Carbon from Spartina alterniflora and Its N-Doped Material for Li-Ion Battery Anode" Nanomaterials 15, no. 9: 658. https://doi.org/10.3390/nano15090658

APA Style

Shang, H., Hao, X., Zhou, Y., Peng, J., Guo, L., Li, H., & Sun, B. (2025). Activated Carbon from Spartina alterniflora and Its N-Doped Material for Li-Ion Battery Anode. Nanomaterials, 15(9), 658. https://doi.org/10.3390/nano15090658

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop