Next Article in Journal
A Novel HDAC6 Inhibitor Ameliorates Imiquimod-Induced Psoriasis-Like Inflammation in Mice
Previous Article in Journal
Reactivity of Ammonia in 1,2-Addition to Group 13 Imine Analogues with G13–P–Ga Linkages: The Electronic Role of Group 13 Elements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lithium-Decorated C26 Fullerene in DFT Investigation: Tuning Electronic Structures for Enhanced Hydrogen Storage

State Key Laboratory of Exterme Environment Optoelectronic Dynamic Measyrement Technology and Instrument, North University of China, Taiyuan 030051, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(15), 3223; https://doi.org/10.3390/molecules30153223 (registering DOI)
Submission received: 8 July 2025 / Revised: 23 July 2025 / Accepted: 28 July 2025 / Published: 31 July 2025

Abstract

Hydrogen energy holds immense potential to address the global energy crisis and environmental challenges. However, its large-scale application is severely hindered by the lack of efficient hydrogen storage materials. This study systematically investigates the H2 adsorption properties of intrinsic C26 fullerene and Li-decorated C26 fullerene using density functional theory (DFT) calculations. The results reveal that Li atoms preferentially bind to the H5-5 site of C26, driven by significant electron transfer (0.90 |e|) from Li to C26. This electron redistribution modulates the electronic structure of C26, as evidenced by projected density of states (PDOS) analysis, where the p orbitals of C atoms near the Fermi level undergo hybridization with Li orbitals, enhancing the electrostatic environment for H2 adsorption. For Li-decorated C26, the average adsorption energy and consecutive adsorption energy decrease as more H2 molecules are adsorbed, indicating a gradual weakening of adsorption strength and signifying a saturation limit of three H2 molecules. Charge density difference and PDOS analyses further demonstrate that H2 adsorption induces synergistic electron transfer from both Li (0.89 |e| loss) and H2 (0.01 |e| loss) to C26 (0.90 |e| gain), with orbital hybridization between H s orbitals, C p orbitals, and Li orbitals stabilizing the adsorbed system. This study aimed to provide a comprehensive understanding of the microscopic mechanism underlying Li-enhanced H2 adsorption on C26 fullerene and offer insights into the rational design of metal-decorated fullerene-based systems for efficient hydrogen storage.

Graphical Abstract

1. Introduction

As the global energy demand soars incessantly and the alarm over environmental challenges intensifies, the quest for energy storage solutions that are both efficacious and enduring has ascended to a critical imperative [1,2]. Hydrogen, endowed with high energy density, renewability, and friendliness toward the environment, emerges as a promising clean energy carrier, making it a potential substitute for fossil fuels in the future energy landscape [3,4]. Yet, the large-scale application of hydrogen energy is limited by the difficulties in its storage and transportation, particularly the lack of efficient and reversible hydrogen storage materials at room temperature [5,6]. The efficient storage of hydrogen is a critical challenge that must be addressed to facilitate the widespread adoption of hydrogen as an energy carrier [7].
Among various hydrogen storage methods, physical adsorption on solid materials has emerged as a particularly attractive option, offering the potential for high storage capacity, rapid adsorption kinetics, and low energy consumption [8,9]. In recent years, carbon-based nanomaterials, such as carbon nanotubes, graphene, and fullerenes, have garnered significant interest for hydrogen storage because of their unique structures and properties [10,11]. Fullerenes with their hollow and symmetrical cage-like structures present a sufficient space for the physisorption of hydrogen molecules [12,13]. Compared to common fullerenes such as C60, the larger curvature of C26 may result in a more specific electron cloud distribution, thereby influencing its interaction patterns with other substances [14,15]. Meanwhile, the smaller molecular size could endow it with unique advantages in certain application scenarios, such as the fabrication of nanoscale devices or targeted delivery in the biomedical field [16,17]. Owing to its unique structural and electronic properties, C26 fullerene is expected to serve as a potential high-efficiency hydrogen storage material [18].
The storage capacity of pristine fullerenes, however, is limited by their weak van der Waals interactions with hydrogen molecules [19]. To enhance the hydrogen storage performance of fullerenes, various strategies have been explored, including the modification of fullerenes with metal atoms [20,21]. Metal decoration has been identified as an effective means to strengthen the interaction between fullerenes and hydrogen molecules, thereby increasing the storage capacity and reversibility of the system [22,23].
Lithium (Li), an alkali metal with a high affinity for hydrogen, stands out as a potential modifier for modifying fullerenes to enhance hydrogen storage [24]. The interaction between Li and hydrogen is strong enough to provide a significant binding energy, while the low weight of Li contributes to a high gravimetric storage capacity [25]. Moreover, the electronic structure of Li-decorated fullerenes can be tuned to optimize the hydrogen adsorption and desorption processes under mild conditions, which is essential for practical applications [26].
Therefore, it is necessary to investigate the H2 adsorption properties of C26 fullerene to understand its hydrogen storage mechanism. In this study, the adsorption sites, interaction energies, and electronic structures of H2 adsorbed on intrinsic C26 fullerene and Li-modified C26 fullerene were studied using density functional theory (DFT) methods to evaluate their potential as hydrogen storage media. Our work aims to contribute to the development of efficient hydrogen storage materials, which are crucial for the practical implementation of hydrogen fuel cell technologies and the transition to a sustainable energy economy.

2. Results and Discussion

2.1. Hydrogen Adsorption on C26 Fullerene

From the top view and side view of the C26 fullerene in Figure 1, we can gain insights into its structural characteristics, such as atomic arrangement and symmetry. In terms of the atomic arrangement, the C26 molecule consists of 26 carbon atoms, which are interconnected via covalent bonds to form a closed cage-like structure with D3h symmetry. The surface of the C26 molecule contains three hexagonal rings and twelve pentagonal rings.
To further identify the most stable adsorption site of H2 on C26 fullerene, the different sites were considered to perform energy calculations. It was found that H2 exhibits the lowest adsorption energy (−0.25 eV) at the H5-5 adsorption site, indicating that this is the most stable adsorption site for H2. Upon further adsorption of hydrogen, the calculated stepwise adsorption energy is −0.01 eV, which is close to zero. This indicates that C26 fullerene can only adsorb one hydrogen molecule, implying that the intrinsic C26 fullerene is not suitable as a hydrogen carrier.

2.2. Li-Decorated C26 Fullerene

The binding characteristics of Li atoms on different sites of C26 fullerene were investigated, as presented in Table 1. The results reveal that the H5-5 site exhibits the highest binding energy (−2.68 eV) among all considered positions, indicating the most thermodynamically favorable adsorption configuration for Li decoration. This strong binding is accompanied by a relatively short C-Li distance of 2.22 Å. In contrast, the Li atom bonded at the T site is dynamically unstable, prompting Li atoms to migrate spontaneously to the H5-5 site, highlighting the intrinsic preference for this specific adsorption geometry. The H6-6 and B5-5 sites also demonstrate substantial binding energies (−2.59 eV and −2.42 eV, respectively), with corresponding distances of 2.28 Å and 2.11 Å, reflecting moderate interaction strengths. The B6-5 site, while still stable, exhibits the lowest binding energy (−2.38 eV) among the viable configurations, which is attributed to its intermediate coordination environment.
The cohesive energy of the solid bulk phase of the Li atom is 1.13 eV/atom [27], which is less than the binding energy of the Li atom on C26 fullerene, suggesting that the Li atom tends to bind to C26 fullerene rather than aggregating with each other. These findings suggest that Li-decorated C26 is most effective at the H5-5 site, which may significantly modulate the electronic structure of the C26 fullerene, thereby enhancing its potential as a hydrogen storage medium.

2.3. Interaction Between Li and C26 Fullerene

To further explore the intrinsic relationship between the electronic structure of Li and C26 fullerene, the charge density difference for the Li-decorated C26 fullerene was studied. As shown in Figure 2b, a region of electron depletion (red) is observed around the Li atom, with a region of electron accumulation (blue) around the C26 fullerene. This indicates that electron transfer occurs from the Co surface to the C atom. Bader charge analysis [28,29] showed that 0.90 electrons (0.90 |e|) are transferred from Li atoms to C26 fullerene after the interaction between them. This charge transfer process is attributed to the electronegativity difference between Li atoms and C26 fullerene. Li atoms exhibited low electronegativity and weak electron-binding ability, whereas C26 fullerene had relatively higher electronegativity and stronger electron-attracting capacity. The increased electron cloud density of C26 fullerene after gaining electrons may enhance its ability to adsorb hydrogen.
To gain an in-depth understanding of the interaction between Li and C26 fullerene, a detailed analysis was performed on the projected density of states (PDOS) of C26 fullerene with Li decoration. As observed in Figure 3a, the density of states C26 molecule is mainly composed of the s and p orbitals of the C atoms. Among these, a distinct distribution of the density of states of the p orbitals near the Fermi level was observed. This is due to the formation of a conjugated system via covalent bonding between carbon atoms in the C26 molecule, which allows p-orbital electrons to be delocalized within the molecule, thereby influencing the electronic properties of C26 fullerene.
After Li decoration, as shown in Figure 3b, the s orbital of C atoms becomes broader and more dispersed compared to pristine C26 fullerene. The C s orbital exhibits broadened and less intense peaks compared to pristine C26 fullerene, indicating electron redistribution induced by Li decoration. This effect is attributed to charge transfer from Li to C26 fullerene, altering the electrostatic environment around C atoms. Near the Fermi level, the C p orbital shows reduced peak intensities and increased peak widths, suggesting weakened electron delocalization. This reduction is caused by electron injection from Li, which modifies the hybridization state of C atoms. From Figure 3c, it can be seen that the s and p orbitals of the Li atom show sparse low-intensity peaks concentrated at higher energies (0 eV to 5 eV), consistent with the alkali-metal character of Li.
The PDOS analysis demonstrates that Li decoration modifies the electronic structure of C26 through electron transfer and weak orbital hybridization. These changes enhance the electron density on C26, particularly around the Fermi level, which may improve its ability to adsorb hydrogen by providing favorable electrostatic interactions and increased orbital overlap with H2 molecules.

2.4. Hydrogen Adsorption on Li-Decorated C26 Fullerene

To investigate the role of Li in enhancing the hydrogen adsorption performance of C26 fullerene, the optimized structures of Li-decorated C26 fullerene with different numbers of adsorbed nH2 molecules were investigated, as depicted in Figure 4. For the structure with one H2 adsorption (Figure 4a), the single H2 molecule is positioned near the Li-C26 with the average adsorption energy of −0.15 eV. When two H2 molecules are adsorbed (Figure 4c), the second H2 molecule attaches to the system in a distinct orientation relative to the first one, suggesting that the surface of the Li-decorated C26 has additional available sites to accommodate more H2. As for the structure with three H2 adsorptions (Figure 4d), the third H2 molecule and the first two H2 molecules present a triangular symmetric position.
The average adsorption energy, consecutive adsorption energy, and Li-H2 distance for increasing numbers of H2 molecules adsorbed onto Li-decorated C26 fullerene are shown in Table 2. As the number of adsorbed H2 molecules (n) increases, the average adsorption energy shows a gradual decreasing trend, dropping from −0.15 eV when n = 1 to −0.09 eV when n = 4. This indicates that the adsorption capacity of C26 for H2 is gradually weakened.
The consecutive adsorption energy reflects the energy change during the adsorption process when one more hydrogen molecule is added. It can be observed from Table 2 that the consecutive adsorption energy gradually decreases, falling from −0.15 eV when n = 1 to −0.02 eV when n = 4. This means that as the number of already adsorbed hydrogen molecules increases, it becomes increasingly difficult to adsorb one more hydrogen molecule. When n reaches 4, the value of consecutive adsorption energy is −0.02 eV, close to zero, indicating that the maximum saturation number of hydrogen molecules adsorbed by the C26 structure is 3, which is more than one H2 adsorption of intrinsic C26 fullerene. The distance increases from 2.09 Å (1H2) to 2.44 Å (4H2), implying a weakening interaction between Li and H2 as more H2 molecules are adsorbed.

2.5. Interaction of Hydrogen with Li-Decorated C26 Fullerene

To elucidate the electronic effects of H2 adsorption, the charge density difference of H2 adsorption on Li-decorated C26 fullerene was studied, as shown in Figure 4b. The electron transfer was quantified by Bader charge analysis: after hydrogen absorption, Li loses 0.89 |e|, H2 loses 0.01 |e|, and C26 gains 0.90 |e|. Upon H2 adsorption, the charge depletion around Li becomes more pronounced in the H2 adsorbed system (broader red–green regions), indicating a strengthened electron-donating role of Li under H2 interaction. Conversely, the charge accumulation on C26 expands and intensifies (more extensive blue regions) compared to the Li-decorated C26 system, which is attributed to the combined electron transfer from both Li and H2. The decoration of Li to C26 leads to the spatial charge redistribution, which enhances H2 adsorption stability to its saturation at three H2 molecules, consistent with the trends in adsorption energy, providing a microscopic storage mechanism for C26 fullerene.
The PDOS of Li-decorated C26 fullerene after H2 adsorption was calculated to explore the interaction of H2 with Li-decorated C26 fullerene, as shown in Figure 5. For the isolated H2 molecule in Figure 5a, a sharp and intense peak is observed in the H s orbital at −17 eV. After H2 adsorption onto Li-decorated C26, as shown in Figure 5b, the peak of the H s orbital was shifted to a higher energy (−5 eV) and broadened. This change is attributed to electron transfer and orbital hybridization between H2 and the C26-Li system. The interaction disrupts the localized H2 s orbital, redistributing electron density and altering the energy distribution of H electrons.
As shown in Figure 5d, upon H2 adsorption, the peak intensities in s and p orbitals of the Li atom were decreased, and distributions became more dispersed. This reduction in electron density is caused by electron transfer from Li to H2, modifying the local electronic environment of the Li atom. The synergistic effects of electron transfer and orbital hybridization resulted in preferential H2 adsorption at Li-decorated C26, exhibiting selectivity. This selective adsorption is of great significance for the development of efficient hydrogen storage materials, as it can improve both the storage capacity and selectivity of C26 fullerene.

2.6. Kubas-Type Interaction

In our calculations, the H-H bond length of H2 molecules adsorbed on Li-decorated C26 (for one to four H2 molecules) was consistently 0.77 Å, which is slightly elongated compared to the 0.75 Å bond length of isolated H2. This slight elongation (0.02 Å) aligns with the key characteristic of Kubas-type interactions, where H-H bond length changes moderately (neither drastic enough for chemisorption nor negligible for pure physisorption), as observed in the Y-decorated C24 system [30].
Additionally, our charge density difference and PDOS analyses reveal electron transfer from Li and H2 to C26, with orbital hybridization between H s orbitals, C p orbitals, and Li orbitals. This orbital interaction, coupled with the observed H-H bond elongation, suggests a mechanism analogous to the Kubas interaction described for Y-decorated systems, where charge transfer and weak orbital hybridization contribute to an interaction strength between physisorption and chemisorption.
Notably, the H-H bond length remains nearly constant as more H2 molecules are adsorbed, while the distance between Li and H centers increases slightly. This stability in bond elongation further supports the presence of Kubas-type interactions, as the interaction mechanism remains consistent across different H2 coverage levels. These observations indicate that Kubas-type interactions play a significant role in H2 adsorption on Li-decorated C26, contributing to the enhanced adsorption capacity compared to intrinsic C26 fullerene. In future research, we will further elaborate on the orbital-level charge transfer details to better quantify this interaction.

3. Computational Details

Computational Methods

Density functional theory (DFT) computations were employed by the Vienna ab initio simulation package (VASP) [31,32] 6.1.0. The interactions between electrons and ions were represented by the projector augmented wave (PAW) method [33,34]. The exchange-correlation functional was determined within the framework of the generalized gradient approximation (GGA) with Perdew–Burke–Ernzerhof (PBE) functional [35,36]. Spin polarization was calculated during the process of geometric optimization, and van der Waals (vdW) interactions were considered by the DFT-D3 method with Becke–Johnson damping [27]. A plane wave basis set cutoff energy [37] of 450 eV was applied. The convergence criteria of geometric optimization process for energy and force thresholds were set to 10−5 eV and 0.02 eV/Å, respectively. To separate the interactions between adjacent slabs, a 30 Å vacuum was established. The k-point integration was sampled using the Monkhorst–Pack scheme [38] with a (1 × 1 × 1) grid. The vacuum layer was set to 20 Å to avoid the interaction of periodically repeated slabs.
The binding energy (Eb) of the Li atom to C26 was calculated as follows:
E b = E Li + C 26 E C 26 E Li
where ELi+C26, EC26, and ELi are the total energies of Li-decorated C26 fullerene, clean C26 fullerene, and isolated Li atoms, respectively. Hence, a negative binding energy suggests that the metal atom can be attached to the C26.
The average adsorption energy (Ead) and consecutive adsorption energy (Ec) of the C26 fullerene to the hydrogen molecule were calculated as follows:
E ad = E nH 2 + Li + C 26 E Li + C 26 n E H 2 / n
E c = E nH 2 + Li + C 26 E n 1 H 2 + Li + C 26 E H 2
where EnH2+Li+C26, E(n−1)H2+Li+C26, and EH2 are the total energies of nH2 and (n − 1)H2 adsorbed on the Li-decorated C26 fullerene and the total energy of an isolated hydrogen molecule, respectively; n is the number of adsorbed hydrogen molecules.
The larger absolute value of adsorption energy indicates a stronger adsorption strength. The larger the negative value, the stronger the adsorption capacity. The consecutive adsorption energy is the adsorption energy when the nth hydrogen molecule adsorbed onto C26, which already has (n − 1) hydrogen molecules adsorbed. The small value of consecutive adsorption energy implies that the adsorption process is not energetically favored, which is typically not desired for efficient hydrogen storage materials.

4. Conclusions

In summary, this study systematically investigates the structural, electronic, and H2 adsorption properties of Li-modified C26 fullerenes through DFT calculations, unraveling the critical role of Li decoration in modulating H2 adsorption behavior at the atomic and electronic levels. Intrinsic C26 fullerene exhibits limited hydrogen storage capacity, with a maximum adsorption of only one H2 molecule due to weak van der Waals interaction. Li decoration emerges as an effective strategy to overcome this limitation. The strong binding of Li atoms to the H5-5 site of C26 triggers substantial electron transfer (0.90 |e|) from Li to C26. PDOS reveals that Li decoration reconstructs the s and p orbitals of C atoms, altering electron distribution near the Fermi level and strengthening interactions with H2. For Li-decorated C26, the average adsorption energy and consecutive adsorption energy both decrease with increasing H2 coverage, with a saturation capacity of three H2 molecules. This saturation is attributed to the weakening Li-H2 interaction and reduced electron donation from Li, as observed in the charge density. Bader charge analysis further reveals that H2 adsorption induces synergistic electron transfer from Li and H2 to C26, with orbital hybridization between H s orbitals, C p orbitals, and Li orbitals stabilizing the adsorbed system, as evidenced by PDOS shifts in H s orbitals. These findings demonstrate that Li decoration enhances the hydrogen storage capacity of C26 by modulating its electronic structure through electron transfer and orbital hybridization, providing a theoretical basis for designing C26-based hydrogen storage materials.

Author Contributions

Investigation, Q.L., J.Y. and Z.X.; Data curation, J.Y. and Y.S.; Writing—original draft, J.Y., Y.S. and L.L.; Writing—review and editing, L.L. and C.L.; Project administration, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 22203080), the Key Research and Development Program of Shanxi Province (Grant NO. 202302030201001), and the Science and Technology Major Program of Shanxi Province (Grant NO. 202301030201003).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Zhou, J.; Ji, W.; Cao, X.; He, W.; Fan, J.; Yuan, Y. A current perspective on the renewable energy hydrogen production process. J. Therm. Sci. 2023, 32, 542–596. [Google Scholar] [CrossRef]
  2. Yang, J.; Lam, T.Y.; Luo, Z.; Cheng, Q.; Wang, G.; Yao, H. Renewable energy driven electrolysis of water for hydrogen production, storage, and transportation. Renew. Sustain. Energy Rev. 2025, 218, 115804. [Google Scholar] [CrossRef]
  3. Dai, S.; Shen, P.; Deng, W.; Yu, Q. Hydrogen energy in electrical power systems: A review and future outlook. Electronics 2024, 13, 3370. [Google Scholar] [CrossRef]
  4. Ge, L.; Zhang, B.; Huang, W.; Li, Y.; Hou, L.; Xiao, J.; Mao, Z.; Li, X. A review of hydrogen generation, storage, and applications in power system. J. Energy Storage 2024, 75, 109307. [Google Scholar] [CrossRef]
  5. Wang, Z.; Lao, G. Hydrogen production from renewable sources: Bridging the gap to sustainable energy and economic viability. Int. J. Hydrogen Energy 2025, 117, 121–134. [Google Scholar] [CrossRef]
  6. Buckley, C.E.; Chen, P.; van Hassel, B.A.; Hirscher, M. Hydrogen-based Energy Storage (IEA-HIA Task 32). Appl. Phys. A-Mater. Sci. Process. 2016, 122, 141. [Google Scholar] [CrossRef]
  7. Li, C.; Li, H.; Yue, H.; Lv, J.; Zhang, J. Flexibility Value of Multimodal Hydrogen Energy Utilization in Electric-Hydrogen-Thermal Systems. Sustainability 2024, 16, 4939. [Google Scholar] [CrossRef]
  8. Liu, P.; Zhang, H.; Cheng, X.; Tang, Y. Ti-decorated B38 fullerene: A high capacity hydrogen storage material. Int. J. Hydrogen Energy 2016, 41, 19123–19128. [Google Scholar] [CrossRef]
  9. Li, C.; Li, J.; Wu, F.; Li, S.-S.; Xia, J.-B.; Wang, L.-W. High capacity hydrogen storage in Ca decorated graphyne: A first-principles study. J. Phys. Chem. C 2011, 115, 23221–23225. [Google Scholar] [CrossRef]
  10. Srinivasu, K.; Ghosh, S.K. Transition metal decorated porphyrin-like porous fullerene: Promising materials for molecular hydrogen adsorption. J. Phys. Chem. C 2012, 116, 25184–25189. [Google Scholar] [CrossRef]
  11. Salem, M.A.; Katin, K.P.; Kaya, S.; Kochaev, A.I.; Maslov, M.M. Interaction of dopants and functional groups adsorbed on the carbon fullerenes: Computational study. Phys. E Low-Dimens. Syst. Nanostructures 2020, 124, 114319. [Google Scholar] [CrossRef]
  12. Dai, W.; Xiao, M.; Chen, M.-Q.; Xu, J.-J.; Tang, Y.-J. A simulation study on the hydrogen storage properties of fullerene family molecules Cx (x = 56, 60, 70) and their hydrides. Mod. Phys. Lett. B 2016, 30, 1650303. [Google Scholar] [CrossRef]
  13. Wang, Y.; Alcami, M.; Martin, F. Understanding the Supramolecular Self-Assembly of the Fullerene Derivative PCBM on Gold Surfaces. Chemphyschem 2008, 9, 1030–1035. [Google Scholar] [CrossRef]
  14. Chandrakumar, K.R.S.; Ghosh, S.K. Alkali-metal-induced enhancement of hydrogen adsorption in C60 fullerene: An ab initio study. Nano Lett. 2008, 8, 13–19. [Google Scholar] [CrossRef] [PubMed]
  15. Yadav, P.; Rai, P.K.; Kumar, P. How an electric field makes endohedral fullerene an improved catalyst for hydrogen evolution reaction. Comput. Theor. Chem. 2023, 1221, 114026. [Google Scholar] [CrossRef]
  16. Chen, Z.; Jiao, H.; Bühl, M.; Hirsch, A.; Thiel, W. Theoretical Investigation into Structures and Magnetic Properties of Smaller Fullerenes and their Heteroanalogues. Theor. Chem. Acc. Theory Comput. Model. (Theor. Chim. Acta) 2001, 106, 352–363. [Google Scholar] [CrossRef]
  17. Dunk, P.W.; Kaiser, N.K.; Mulet-Gas, M.; Rodriguez-Fortea, A.; Poblet, J.M.; Shinohara, H.; Hendrickson, C.L.; Marshall, A.G.; Kroto, H.W. The smallest stable fullerene, M@C28 (M = Ti, Zr, U): Stabilization and growth from carbon vapor. J. Am. Chem. Soc. 2012, 134, 9380–9389. [Google Scholar] [CrossRef] [PubMed]
  18. Dogan, M.; Kalafat, M.Y.; Kizilduman, B.K.; Bicil, Z.; Turhan, Y.; Yanmaz, E.; Duman, B. Hydrogen storage analysis of fullerene and defective fullerenes: The first experimental study. Fuel 2025, 390, 134705. [Google Scholar] [CrossRef]
  19. Lee, H.; Park, D.G.; Park, J.; Kim, Y.-H.; Kang, J.K. Amorphized Defective Fullerene with a Single-Atom Platinum for Room-Temperature Hydrogen Storage. Adv. Energy Mater. 2023, 13, 2300041. [Google Scholar] [CrossRef]
  20. Sarfaraz, S.; Yar, M.; Lakhani, A.; Sheikh, N.S.; Bayach, I.; Ayub, K. Probing the catalytic activity of first-row transition metal doped C20 fullerene as remarkable HER electrocatalysts: A DFT study. Mater. Sci. Semicond. Process. 2023, 167, 107785. [Google Scholar] [CrossRef]
  21. Chai, Z.; Liu, L.; Liang, C.; Liu, Y.; Wang, Q. Insight into the Reversible Hydrogen Storage of Titanium-Decorated Boron-Doped C20 Fullerene: A Theoretical Prediction. Mol. Catal. 2024, 29, 4728. [Google Scholar] [CrossRef] [PubMed]
  22. Paul, D.; Mane, P.; Sarkar, U.; Chakraborty, B. Yttrium decorated fullerene C30 as potential hydrogen storage material: Perspectives from DFT simulations. Theor. Chem. Acc. 2023, 142, 94. [Google Scholar] [CrossRef]
  23. Mahamiya, V.; Shukla, A.; Chakraborty, B. Applied Surface Science Scandium decorated C24 fullerene as high capacity reversible hydrogen storage material: Insights from density functional theory simulations. Appl. Surf. Sci. 2022, 573, 151389. [Google Scholar] [CrossRef]
  24. Pandey, A.K.; Singh, V.; Dubey, D.D.; Pandey, K.K.; Avaish, M.; Dwivedi, A. Alkali Metal-Doped Fullerenes as Hydrogen Storage-A Quantum Chemical Investigation. In Proceedings of the 4th National Conference on Material and Device (NCMD), Moradabad, India, 28–29 December 2023. [Google Scholar]
  25. Sahoo, R.K.; Chakraborty, B.; Sahu, S. Reversible hydrogen storage on alkali metal (Li and Na) decorated C20 fullerene: A density functional study. Int. J. Hydrogen Energy 2021, 46, 40251–40261. [Google Scholar] [CrossRef]
  26. Deniz, C.U.; Mert, H.; Baykasoglu, C. Li-doped fullerene pillared graphene nanocomposites for enhancing hydrogen storage: A computational study. Comp. Mater. Sci. 2021, 186, 110023. [Google Scholar] [CrossRef]
  27. Dong, S.; Lv, E.; Wang, J.; Li, C.; Ma, K.; Gao, Z.; Yang, W.; Ding, Z.; Wu, C.; Gates, I.D. Construction of transition metal-decorated boron doped twin-graphene for hydrogen storage: A theoretical prediction. Fuel 2021, 304, 121351. [Google Scholar] [CrossRef]
  28. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comp. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  29. Bader, R. Atoms in molecules: A quantum theory. Int. Ser. Monogr. Chem. 1990, 22, 9–15. [Google Scholar]
  30. Mahamiya, V.; Shukla, A.; Chakraborty, B. Exploring yttrium doped C24 fullerene as a high-capacity reversible hydrogen storage material: DFT investigations. J. Alloys Compd. 2022, 897, 162797. [Google Scholar] [CrossRef]
  31. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  32. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef] [PubMed]
  33. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  34. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  36. Wang, Y.; Burke, K.; Perdew, J. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar]
  37. VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105. [Google Scholar] [CrossRef]
  38. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
Figure 1. The (a) top and (b) side views of the C26 fullerene structure. The C atoms are shown in grey.
Figure 1. The (a) top and (b) side views of the C26 fullerene structure. The C atoms are shown in grey.
Molecules 30 03223 g001
Figure 2. (a) Optimized structure and (b) charge density difference of Li−decorated C26 fullerene. The C atoms and Li atom are shown in grey and purple, respectively. Red regions indicate charge depletion, while green and blue represent regions of reduced and increased charge density, respectively. The isosurfaces are set to 0.16 eÅ−3.
Figure 2. (a) Optimized structure and (b) charge density difference of Li−decorated C26 fullerene. The C atoms and Li atom are shown in grey and purple, respectively. Red regions indicate charge depletion, while green and blue represent regions of reduced and increased charge density, respectively. The isosurfaces are set to 0.16 eÅ−3.
Molecules 30 03223 g002
Figure 3. The projected density of states (PDOS) onto (a) s and p orbitals of the C atom of C26 fullerene; (b) s and p orbitals of the C atom of Li−decorated C26 fullerene; (c) s and p orbitals of the Li atom of Li−decorated C26 fullerene;. The Fermi energy is set to zero. The s and p orbitals are labeled with red and blue lines, respectively.
Figure 3. The projected density of states (PDOS) onto (a) s and p orbitals of the C atom of C26 fullerene; (b) s and p orbitals of the C atom of Li−decorated C26 fullerene; (c) s and p orbitals of the Li atom of Li−decorated C26 fullerene;. The Fermi energy is set to zero. The s and p orbitals are labeled with red and blue lines, respectively.
Molecules 30 03223 g003
Figure 4. The optimized structure of Li−decorated C26 fullerene with (a) 1 H2 adsorption, (c) 2 H2 adsorption, and (d) 3 H2 adsorption. (b) Charge density difference of Li−decorated C26 fullerene with H2 adsorption. The H, C, and Li atoms are shown in white, grey, and purple, respectively. Red regions indicate charge depletion, while green and blue represent regions of reduced and increased charge density, respectively. The isosurfaces are set to 0.16 eÅ−3.
Figure 4. The optimized structure of Li−decorated C26 fullerene with (a) 1 H2 adsorption, (c) 2 H2 adsorption, and (d) 3 H2 adsorption. (b) Charge density difference of Li−decorated C26 fullerene with H2 adsorption. The H, C, and Li atoms are shown in white, grey, and purple, respectively. Red regions indicate charge depletion, while green and blue represent regions of reduced and increased charge density, respectively. The isosurfaces are set to 0.16 eÅ−3.
Molecules 30 03223 g004
Figure 5. The projected density of states (PDOS) onto (a) s orbital of the H atom of isolated H2 molecule; (b) s orbital of the H atom of Li−decorated C26 fullerene after H2 adsorption; (c) s and p orbitals of the Li atom of Li−decorated C26 fullerene; (d) s and p orbitals of the Li atom of Li−decorated C26 fullerene after H2 adsorption. The Fermi energy is set to zero. The s and p orbitals are labeled with red and blue lines, respectively.
Figure 5. The projected density of states (PDOS) onto (a) s orbital of the H atom of isolated H2 molecule; (b) s orbital of the H atom of Li−decorated C26 fullerene after H2 adsorption; (c) s and p orbitals of the Li atom of Li−decorated C26 fullerene; (d) s and p orbitals of the Li atom of Li−decorated C26 fullerene after H2 adsorption. The Fermi energy is set to zero. The s and p orbitals are labeled with red and blue lines, respectively.
Molecules 30 03223 g005
Table 1. The binding energy (Eb) of the Li atom on C26 fullerene and the distance (dC-Li, Å) between Li and C26 at different sites of Li-decorated C26 fullerene.
Table 1. The binding energy (Eb) of the Li atom on C26 fullerene and the distance (dC-Li, Å) between Li and C26 at different sites of Li-decorated C26 fullerene.
SiteEb (eV)dC-Li (Å)
H5-5−2.682.22
H6-6−2.592.28
B5-5−2.422.11
B6-5−2.382.12
Tunstable and move to H5-5 site
Table 2. The average adsorption energy (Eads), consecutive adsorption energy (Ec) of the nH2/Li-C26 system, and corresponding distance (dC-Li, Å) between Li and H2.
Table 2. The average adsorption energy (Eads), consecutive adsorption energy (Ec) of the nH2/Li-C26 system, and corresponding distance (dC-Li, Å) between Li and H2.
nH2Eads (eV)Ec (eV)dLi-H2 (Å)
1H2−0.15−0.152.09
2H2−0.13−0.112.13
3H2−0.11−0.072.25
4H2−0.09−0.022.44
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

Yu, J.; Liu, L.; Li, Q.; Xu, Z.; Shi, Y.; Lei, C. Lithium-Decorated C26 Fullerene in DFT Investigation: Tuning Electronic Structures for Enhanced Hydrogen Storage. Molecules 2025, 30, 3223. https://doi.org/10.3390/molecules30153223

AMA Style

Yu J, Liu L, Li Q, Xu Z, Shi Y, Lei C. Lithium-Decorated C26 Fullerene in DFT Investigation: Tuning Electronic Structures for Enhanced Hydrogen Storage. Molecules. 2025; 30(15):3223. https://doi.org/10.3390/molecules30153223

Chicago/Turabian Style

Yu, Jiangang, Lili Liu, Quansheng Li, Zhidong Xu, Yujia Shi, and Cheng Lei. 2025. "Lithium-Decorated C26 Fullerene in DFT Investigation: Tuning Electronic Structures for Enhanced Hydrogen Storage" Molecules 30, no. 15: 3223. https://doi.org/10.3390/molecules30153223

APA Style

Yu, J., Liu, L., Li, Q., Xu, Z., Shi, Y., & Lei, C. (2025). Lithium-Decorated C26 Fullerene in DFT Investigation: Tuning Electronic Structures for Enhanced Hydrogen Storage. Molecules, 30(15), 3223. https://doi.org/10.3390/molecules30153223

Article Metrics

Back to TopTop