Next Article in Journal
Luminescent Gold Nanoclusters for Bioimaging: Increasing the Ligand Complexity
Next Article in Special Issue
Nanoporous Hollow Carbon Spheres Derived from Fullerene Assembly as Electrode Materials for High-Performance Supercapacitors
Previous Article in Journal
Recent Progress of Energy-Storage-Device-Integrated Sensing Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Reversible Hydrogen Storage Media by g-CN Monolayer Decorated with NLi4: A First-Principles Study

1
State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China
2
School of Materials Science and Engineering, Chongqing University of Arts and Sciences, Chongqing 402160, China
3
Chongqing Key Laboratory of Precision Optics, Chongqing Institute of East China Normal University, Chongqing 401120, China
4
School of Computer Science and Technology, Northwestern Polytechnical University, Xian 710129, China
5
School of Science, Key Laboratory of High Performance Scientific Computation, Xihua University, Chengdu 610039, China
6
School of Electric and Electrical Engineering, Shangqiu Normal University, Shangqiu 476000, China
7
Material Science and Engineering Department, City University of Hongkong, Hongkong 999077, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(4), 647; https://doi.org/10.3390/nano13040647
Submission received: 28 December 2022 / Revised: 2 February 2023 / Accepted: 4 February 2023 / Published: 7 February 2023
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials II)

Abstract

:
A two-dimensional graphene-like carbon nitride (g-CN) monolayer decorated with the superatomic cluster NLi4 was studied for reversible hydrogen storage by first-principles calculations. Molecular dynamics simulations show that the g-CN monolayer has good thermal stability at room temperature. The NLi4 is firmly anchored on the g-CN monolayer with a binding energy of −6.35 eV. Electronic charges are transferred from the Li atoms of NLi4 to the g-CN monolayer, mainly due to the hybridization of Li(2s), C(2p), and N(2p) orbitals. Consequently, a spatial local electrostatic field is formed around NLi4, leading to polarization of the adsorbed hydrogen molecules and further enhancing the electrostatic interactions between the Li atoms and hydrogen. Each NLi4 can adsorb nine hydrogen molecules with average adsorption energies between −0.152 eV/H2 and −0.237 eV/H2. This range is within the reversible hydrogen storage energy window. Moreover, the highest achieved gravimetric capacity is up to 9.2 wt%, which is superior to the 5.5 wt% target set by the U.S. Department of Energy. This study shows that g-CN monolayers decorated with NLi4 are a good candidate for reversible hydrogen storage.

1. Introduction

Hydrogen energy is a promising alternative to traditional fossil fuels due to its highest energy density and the environmental friendliness of its combustion products. Therefore, interest in hydrogen storage is growing due to increasing environmental protection requirements and the trend of low-carbon development. It is expected that the demand in the global hydrogen storage market will expand at a rate of 5.8% from 2019 to 2025 [1]. However, there is a gap between this demand and existing storage technology, where materials play a critical role.
Conventional strategies such as high-pressure compression and liquefaction suffer from low safety and high cost [2]. Therefore, solid-state hydrogen storage is becoming an attractive alternative. A suitable storage medium should have good reversibility between adsorption and desorption and an adequate binding energy of about −0.1 to −0.2 eV per H2 [3]. The energy requirement for reversible adsorption just falls into the range of physical adsorption, which guarantees reversible and fast dynamics. Motivated by this consideration, researchers have focused on various nanostructured materials, including two-dimensional (2D) sheets (boron-based nanomaterials, graphene-like nanomaterials, MXenes, MoS2, CxNy, etc.) [4,5,6,7,8,9,10,11,12,13], metal atom-modified covalent organic frameworks (COFs) [14,15,16], and metal organic frameworks (MOFs) [17,18,19,20,21]. These materials can exhibit significantly improved kinetics during the adsorption process by decreasing the enthalpy of formation and hydrogenation temperature, and they exhibit many other catalytic effects [22,23,24].
Two-dimensional materials are believed to provide a pathway for the design of next-generation hydrogen storage media due to their superior physical and chemical properties, including their large surface-to-volume ratios, abundant active sites, light weight, and adjustability [25]. However, it is challenging for 2D materials to directly adsorb H2 molecules because pure 2D materials fail to provide strong electrostatic interaction. The adsorption energies of H2 on graphene and BN are only 0.07 eV [26] and 0.03 eV [27], respectively. Therefore, modification and decoration are necessary to improve their storage capacity. Some research has demonstrated that dopants such as alkali metals, alkaline earth metals, transition metals, and functional groups can improve the storage ability of pristine 2D sheets [26,27,28,29,30,31,32,33,34]. Decorating Mg [28], Ti [35], Ca [36], and Li [26] on graphene-like or carbon-based sheets can provide gravimetric capacities in the range of 7.96 wt% to 10.81 wt%, which meet the 5.5 wt% target value proposed by the U.S. Department of Energy (DOE) in 2020. In addition to single particles, superatomic clusters such as NLi4 are also an eye-catching option. The suitability of these clusters has been confirmed by time-of-flight powder neutron diffraction experiments [37,38]. Like alkali metal ions, NLi4 has low ionization potentials, which means that decorated NLi4 can easily transfer electrons to the substrate and become positively charged. Consequently, an enhanced local electrostatic field can be obtained, which is beneficial for improving hydrogen storage capacity compared to the use of single particles. However, unlike their alkali metal counterparts, the robust binding of Li-N provides superior stability for these clusters. This guarantees a stable connection between the clusters and substrate in addition to less aggregation between the clusters. Some research has provided theoretical support for the significant potential of NLi4 for hydrogen storage [33,38,39]. Xiang Wang et al. found that NLi4 can enhance the adsorption performance of boron-based 2D materials by forming covalent bonds between the N atoms of NLi4 clusters and the B atoms of an h-BN sheet. The average adsorption energy per H2 is around −0.20 eV, and the capacity can reach 9.40 wt% [40]. NLi4 clusters show better performance when decorated on graphene-like substrates with better conductivity. Hao Qi et al. reported that NLi4-decorated graphene can achieve a hydrogen storage capacity of 10.75 wt% with an average adsorption energy of −0.21 eV/H2 due to its ideal adsorption strength and abundance of anchor sites [33].
As a type of 2D material, the graphene-like g-CN monolayer structure is composed of uniform holes and aromatic benzene rings containing three N atoms and three C atoms alternatively arranged in the ring [26]. This material is a member of the CxNy family, which is a group of heteroatom-doped carbon materials. A typical and well-known CxNy is g-C3N4, whose unique structure makes it an attractive candidate for both photocatalytic and electrochemical applications [41]. Many studies have verified the availability and structure adjustability of C3N4 by various methods, including physical and chemical vapor deposition, thermal condensation, microwave-assisted processes, electrodeposition, hydrothermal and solvothermal synthesis, and sol-gel processes [42,43,44]. In addition to g-C3N4, the synthesis strategies for preparing of CxNy materials with other ratios, such as C3N5, C3N3, C2N, and CN, have been developed in the past decade [45]. The substrate materials in our designed CN can be obtained by reacting C3N3Cl3 with molten alkali Na and K, indicating its potential for practical use [46]. G-CN is a semiconductor whose bandgap can be adjusted by dopants [41], and it is expected to have superior hydrogen storage capacity due to its large surface-to-volume ratio and porous geometry structure, which provide sufficient adsorption sites for H2 molecules. Some theoretical studies have demonstrated that the decoration of Li [26], Mg [47], and Al [48] noticeably enhances gravimetric density from 6.5 wt% to 10.81 wt%, and the average adsorption energy of H2 molecules is in the range of −0.1 to −0.23 eV.
In this work, we used first-principles calculations to design a novel material consisting of NLi4 clusters decorated on a g-CN monolayer for hydrogen storage applications. The main motivations for the decoration of NLi4 are: (i) unlike general metal ions, the NLi4 superatomic cluster can alleviate the cluster effect and can be evenly distributed on the 2D substrate; (ii) NLi4 is firmly adsorbed on the surface of the g-CN monolayer (the binding energy is −6.35 eV, which is lower than the previously reported −6.178 and −2.47 eV) [33,38]; (iii) both dopants (NLi4 and g-CN) only contain superlight elements, which means that their complex has higher gravimetric capacity for hydrogen storage; (iv) g-CN is favorable for excellent hydrogen storage performance due to its flexible adjustable electronic structure and porous geometry. The average adsorption energy per H2 of the NLi4@g-CN is between −0.152 eV and −0.237 eV, and a high gravimetric density of 9.2 wt% can be achieved, which considerably surpasses the DOE target value of 5.5 wt%. In addition, the partial density of states (PDOS), charge density, Bader charge analysis, adsorption process, and storage mechanism were thoroughly investigated, and the results also support the excellent hydrogen storage performance of NLi4@g-CN. We hope our design and analysis will lay a foundation for the further practical application of advanced energy materials.

2. Computational Details

All calculations were conducted by the Vienna Ab initio Simulation Package (VASP) using the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional under periodic boundary conditions [49,50]. The core electron interactions were described via the projector augmented wave (PAW) approach [51], and the electronic states were approximated by the solution of the plane waves. To avoid coupling effects among the periodic structures, a 20 Å vacuum layer was added on the slab model along the direction perpendicular to the sheet plane. The expansion cut-off energy of the plane wave was set to 520 eV. The optimized structures were obtained by relaxation with a conjugated-gradient algorithm, where the energy convergence of the atomic position and lattice parameters was 1 × 10−5 eV. The Hellmann–Feynman forces on each atom were converged within 0.02 eV/Å. The spin effects were considered, and van der Waals corrections were also applied via the DFT-D2 method of Grimme. A Brillouin zone was set for the cell with Gamma-centered k-point grids of 3 × 3 × 1 and 20 × 20 × 1 (based on the convergence tests shown in Supplementary Materials) for structural optimization and partial density of states (PDOS) calculations, respectively [50,52]. The charge transfer between the NLi4 superatomic cluster and g-CN was described by Bader charge analysis [53,54].

3. Results and Discussion

In this work, the optimized lattice parameters of the 2 × 2 × 1 g-CN supercell are a = b = 1.424 nm, and its structure is shown in Figure 1. The g-CN supercell consists of C-N (CN) 6-membered rings and CN 18-membered rings formed by the connections among the CN 6-membered rings, which is identical to previous reports [26]. The porosity of the g-CN layer provides sufficient anchoring active sites for the subsequent decoration of NLi4 superatomic clusters, indicating great potential for doping and hydrogen storage. The PDOS for the 2s/2p orbitals of the nitrogen and carbon atoms in the g-CN monolayer were calculated, as presented in Figure 2. Hybridizations exist between the C-2p and N-2p orbitals in the energy range of −6 eV to 4 eV, and the valence bands are mainly occupied by C-2p and N-2p. The band gap of g-CN is about 1.3 eV, i.e., the g-CN is a semiconductor. The thermodynamic stability of the g-CN monolayer at room temperature (300 K) was estimated by first-principles molecular dynamics (MD) simulation with the Nose–Hoover thermostat algorithm. As shown in Figure 3, the system energy slightly oscillates around −406.5 eV. Even after exerting thermal perturbation, the original planar structure with the CN 6-membered rings and 18-membered rings is well-retained, reflecting the good thermodynamic stability of the g-CN.
The porous structure of the g-CN monolayer makes it an ideal host for various dopants, including metal atoms and nanoclusters. According to the DFT calculation, the four possible binding sites of NLi4 on the g-CN monolayer were tested, including the inner cavity of the large CN 18-membered ring in the center, and the inner cavity of the surrounding small NC rings, C-C bonds, and C-N bonds. It was determined that NLi4 can be easily doped on the cavity formed by the large CN 18-membered rings via the formation of ionic bonds between the Li atoms of the NLi4 cluster and the N atoms of the g-CN, as shown in Figure 4. The binding energy Eb was also calculated according to Formula (1) [55].
E b = E NLi 4 @ g - CN E NLi 4 E g - CN
where E NLi 4 @ g - CN , E NLi 4 , and E g - CN are the energies of the single NLi4-decorated g-CN monolayer, an isolated NLi4 cluster, and the pure g-CN monolayer, respectively. The binding energy of NLi4 and g-CN is −6.35 eV, meaning that a stable interaction exists between NLi4 and the g-CN monolayer. The top and side views of the optimized NLi4@g-CN structure (Figure 4) show that g-CN is not twisted by the decoration of NLi4. Therefore, the complex still has a high surface-to-volume ratio. The side view shows that the superatomic cluster has a “pyramid building” appearance on the g-CN surface. This is a stable connection and provides enough space for subsequent H2 adsorption.
The nature of the interaction between the NLi4 dopant and g-CN was investigated by PDOS and charge density difference analysis. NLi4 decoration can improve the conductivity of g-CN, supporting our assumption that binding formation and charge transfer occur between NLi4 and g-CN. Figure 5 shows that the bandgap disappears at the Fermi level, indicating the transformation of the complex from semiconductor to conductor due to the doping of NLi4. Clear hybridization between the Li(2s) orbital and the C(2p)/N(2p) orbitals is present at −0.2 and −0.5 eV below the Fermi level, indicating that charge transfer and the formation of ion bonds occur during the combination process. The charge density difference of NLi4@g-CN (Figure 6) also provides evidence for this mechanism. There is a clear charge transfer from NLi4 to the substrate g-CN. NLi4 loses part of its charge, displaying electronegativity. Meanwhile, the g-CN gains charge and displays electropositivity. Thus, a polarization field is formed around NLi4, paving the way for subsequent H2 adsorption. In addition, the highly charged NLi4 superatomic clusters strongly repel each other, which inhibits the aggregation of NLi4 molecules. To quantitatively investigate the charge transfer between the dopants and substrate, Bader charge analysis was performed, showing that this transformation is about 0.87 e/atom. This indicates that bonding is formed by strong ionic interaction between the NLi4 clusters and g-CN. In other words, a novel material was developed by decorating NLi4 on a g-CN monolayer with a large surface-to-volume ratio and favorable electron structure. This is highly promising for achieving excellent H2 adsorption performance.

4. Hydrogen Adsorption Performance of NLi4@g-CN

To thoroughly investigate the H2 adsorption performance of the NLi4@g-CN monolayer, hydrogen molecules were systematically added to the top and side of the Li ions of NLi4, followed by structural optimization to obtain the most stable configurations. The corresponding results are shown in the Figure 7. Each NLi4 anchored on the g-CN can accommodate a maximum of nine H2 molecules. As displayed in Figure 6, the Li ions of the NLi4 cluster partially transfer their charge to the substrate, forming a local electrostatic field around the NLi4 decoration sites. This is favorable for hydrogen adsorption. This assumption also was confirmed by charge density difference analysis of the NLi4@g-CN with adsorbed H2, as shown in Figure 8. The notable positive electronic potential around the Li of the NLi4 superatomic cluster induces the adsorbed H2 to have an unbalanced charge distribution. The center of the H-H bond is a charge-abundant area while the ends are charge-deficient, indicating the polarization of the H2 molecules. Therefore, H2 is smoothly adsorbed by NLi4@g-CN due to the polarization mechanism, and the length of the H-H bond is elongated to 0.76 Å (the bond length of free H2 is 0.75 Å [11]). To perform a more quantitative analysis, the average adsorption energy per H2 and storage capacity were calculated using Formulas (2) and (3). The average adsorption energy per H2 can be written as:
E ad = ( E NLi 4 @ g - CN - nH 2 E NLi 4 @ g - CN n E H 2 ) / n
where E ad , E NLi 4 @ g - CN - nH 2 , E NLi 4 @ g - CN , E H 2 , and n are the average adsorption energy per H2, the energy of the NLi4-decorated g-CN monolayer with adsorbed H2, the energy of the NLi4@g-CN complex, the energy of a single H2, and the number of H2 molecules, respectively. The storage capacity can be defined as:
W t = n ( H ) * M ( H ) [ n ( C ) * M ( C ) + n ( N ) * M ( N ) + n ( Li ) * M ( Li ) ]
where W t , n ( H ) , n ( C ) , n ( N ) , and n ( Li ) denote the storage capacity of the complex and the number of H, C, N, and Li atoms, respectively. M ( H ) , M ( C ) , M ( N ) , and M ( Li ) denote the molar masses of H, C, N, and Li atoms, respectively. The calculated results are listed in Table 1.
To systematically estimate the conditions for hydrogen desorption from NLi4@g-CN, the desorption temperature was calculated by the van’t Hoff equation, which can be written as:
T d = E ad K B ( Δ S R - lnp ) - 1
where T d , E ad , K B , Δ S , R and p are the desorption temperature, the average adsorption energy per H2, Boltzmann’s constant, the entropy change of H2 from gas to solid, the universal gas constant, and the equilibrium pressure, respectively [56]. This work focused on the desorption temperature Td at standard atmospheric pressure, so p = 1 atm and Δ S = 75.44 J K−1 mol−1 were used for calculation. The estimated desorption temperatures are also listed in Table 1. Table 1 shows that the RH-H value is about 0.76 Å and the Td is between 198–303 K, which is much higher than the critical temperature of H2. The average adsorption energy per H2 is in the range of −0.152 eV to −0.237 eV, which falls into the range of appropriate reversible adsorption energy. Table 1 also shows that for a single NLi4 superatomic cluster, the average adsorption energies per H2 are negatively correlated with the number of adsorbed H2 molecules, which means that the average adsorption energy per H2 decreases with increasing H2 loading. This is caused by the repulsion among the adsorbed H2, which is in agreement with the behavior reported in Ref. [11]. With an increasing number of NLi4 clusters, the adsorption energies per H2 maintain proximity as a constant, which means that the storage capacity of g-CN is largely determined by the number of NLi4 superatomic clusters. Finally, the highest gravimetric capacity is 9.2 wt%, which surpasses the DOE target of 5.5 wt%.

5. Conclusions

In summary, we proposed a promising NLi4-decorated g-CN material and estimated its performance for H2 storage by first-principles calculations. The NLi4 superatomic cluster can be firmly anchored on a g-CN monolayer with a bonding energy of −6.35 eV. The adsorption energies for H2 are between −0.152 eV and −0.237 eV, lying within the range of reversible hydrogen storage. Moreover, the storage capacity is as high as 9.2 wt%, which is much higher than the benchmark set by the U.S. DOE. This superior hydrogen storage capacity is attributed to the formation of a spatial local electrostatic field around the NLi4 caused by charge transfer from the NLi4 superatomic cluster to the g-CN. Hydrogen molecules tend to be polarized due to the formation of this electrostatic field. Thus, electrostatic interaction is enhanced between the hydrogen molecules and the substrate and the adsorption capacity is favorably improved. In addition, the spatial and electronic properties of the NLi4 cluster inhibit NLi4 aggregation and the repulsion between multiple H2 molecules. In the future, we hope that more advanced hydrogen storage materials will be developed along this direction.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano13040647/s1, Figure S1: The convergence tests of total density of states with an increasing k-point grids of N × N × N for pure g-CN monolayer and the optimized structure coordination of pure g-CN monolayer.

Author Contributions

X.C. and J.R. outlined the whole project and designed the calculations. X.C., Y.L., W.H., N.W., L.Z. and S.Y. carried out the calculations. X.C., J.R., W.H. and J.C. worked on data analysis. X.C., J.R. and F.Z. contributed to the writing of this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Research Program of Chongqing Municipal Education Commission (No. KJQN202201327), the Program of Chongqing Science & Technology Commission (No. cstc2019jcyj-msxmX0877) and the Chongqing Science and Technology Bureau (No. cstc2021jcyj-msxmX0775 and cstc2022ycjh-bgzxm0168).

Data Availability Statement

Data supporting reported results are available online at https://www.mdpi.com/article/10.3390/nano13040647/s1.

Acknowledgments

The numerical calculations in this paper have been done at the Hefei advanced computing center.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Elberry, A.M.; Thakur, J.; Santasalo-Aarnio, A.; Larmi, M. Large-scale compressed hydrogen storage as part of renewable electricity storage systems. Int. J. Hydrogen Energy 2021, 46, 15671–15690. [Google Scholar] [CrossRef]
  2. Tarhan, C.; Çil, M.A. A study on hydrogen, the clean energy of the future: Hydrogen storage methods. J. Energy Storage 2021, 40, 102676. [Google Scholar] [CrossRef]
  3. Bhatia, S.K.; Myers, A.L. Optimum conditions for adsorptive storage. Langmuir 2006, 22, 1688–1700. [Google Scholar] [CrossRef]
  4. Chen, X.-H.; Li, J.-W.; Wu, Q.; Tan, Y.; Yuan, S.; Gao, P.; Zhu, G.-Y. Reversible hydrogen storage for NLi4-Decorated honeycomb borophene oxide. Int. J. Hydrogen Energy 2022, 47, 19168–19174. [Google Scholar] [CrossRef]
  5. Hussain, T.; Mortazavi, B.; Bae, H.; Rabczuk, T.; Lee, H.; Karton, A. Enhancement in hydrogen storage capacities of light metal functionalized Boron–Graphdiyne nanosheets. Carbon 2019, 147, 199–205. [Google Scholar] [CrossRef]
  6. Kumar, P.; Singh, S.; Hashmi, S.A.R.; Kim, K.-H. MXenes: Emerging 2D materials for hydrogen storage. Nano Energy 2021, 85, 105989. [Google Scholar] [CrossRef]
  7. Alhameedi, K.; Hussain, T.; Bae, H.; Jayatilaka, D.; Lee, H.; Karton, A. Reversible hydrogen storage properties of defect-engineered C4N nanosheets under ambient conditions. Carbon 2019, 152, 344–353. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Liu, P.; Zhu, X.; Liu, Z. A reversible hydrogen storage material of Li-decorated two-dimensional (2D) C4N monolayer: First principles calculations. Int. J. Hydrogen Energy 2021, 46, 32936–32948. [Google Scholar] [CrossRef]
  9. Panigrahi, P.; Kumar, A.; Karton, A.; Ahuja, R.; Hussain, T. Remarkable improvement in hydrogen storage capacities of two-dimensional carbon nitride (g-C3N4) nanosheets under selected transition metal doping. Int. J. Hydrogen Energy 2020, 45, 3035–3045. [Google Scholar] [CrossRef]
  10. Panigrahi, P.; Desai, M.; Talari, M.K.; Bae, H.; Lee, H.; Ahuja, R.; Hussain, T. Selective decoration of nitrogenated holey graphene (C2N) with titanium clusters for enhanced hydrogen storage application. Int. J. Hydrogen Energy 2021, 46, 7371–7380. [Google Scholar] [CrossRef]
  11. Varunaa, R.; Ravindran, P. Potential hydrogen storage materials from metal decorated 2D-C2N: An ab initio study. Phys. Chem. Chem. Phys. 2019, 21, 25311–25322. [Google Scholar] [CrossRef] [PubMed]
  12. Faye, O.; Hussain, T.; Karton, A.; Szpunar, J. Tailoring the capability of carbon nitride (C3N) nanosheets toward hydrogen storage upon light transition metal decoration. Nanotechnology 2019, 30, 075404. [Google Scholar] [CrossRef] [PubMed]
  13. Guerrero-Avilés, R.; Orellana, W. Hydrogen storage on cation-decorated biphenylene carbon and nitrogenated holey graphene. Int. J. Hydrogen Energy 2018, 43, 22966–22975. [Google Scholar] [CrossRef]
  14. Tong, M.; Zhu, W.; Li, J.; Long, Z.; Zhao, S.; Chen, G.; Lan, Y. An easy way to identify high performing covalent organic frameworks for hydrogen storage. Chem. Commun. 2020, 56, 6376–6379. [Google Scholar] [CrossRef]
  15. Kopac, T. Covalent Organic Frameworks-Based Nanomaterials for Hydrogen Storage. In Covalent Organic Frameworks; CRC Press: Boca Raton, FL, USA, 2023; pp. 345–360. [Google Scholar]
  16. Nemiwal, M.; Sharma, V.; Kumar, D. Improved designs of multifunctional covalent-organic frameworks: Hydrogen storage, methane storage, and water harvesting. Mini Rev. Org. Chem. 2021, 18, 1026–1036. [Google Scholar] [CrossRef]
  17. Dinca, M.; Long, J.R. Hydrogen storage in microporous metal-organic frameworks with exposed metal sites. Angew. Chem. Int. Ed. Engl. 2008, 47, 6766–6779. [Google Scholar] [CrossRef]
  18. Wang, Y.; Lan, Z.; Huang, X.; Liu, H.; Guo, J. Study on catalytic effect and mechanism of MOF (MOF = ZIF-8, ZIF-67, MOF-74) on hydrogen storage properties of magnesium. Int. J. Hydrogen Energy 2019, 44, 28863–28873. [Google Scholar] [CrossRef]
  19. Butova, V.V.; Burachevskaya, O.A.; Podshibyakin, V.A.; Shepelenko, E.N.; Tereshchenko, A.A.; Shapovalova, S.O.; Il’in, O.I.; Bren, V.A.; Soldatov, A.V. Photoswitchable zirconium MOF for light-driven hydrogen storage. Polymers 2021, 13, 4052. [Google Scholar] [CrossRef]
  20. Hussain, T.; Hankel, M.; Searles, D.J. Computational evaluation of lithium-functionalized carbon nitride (g-C6N8) monolayer as an efficient hydrogen storage material. J. Phys. Chem. C 2016, 120, 25180–25188. [Google Scholar] [CrossRef]
  21. Wu, H.; Zhou, W.; Yildirim, T. Hydrogen storage in a prototypical zeolitic imidazolate framework-8. J. Am. Chem. Soc. 2007, 129, 5314–5315. [Google Scholar] [CrossRef]
  22. Niemann, M.U.; Srinivasan, S.S.; Phani, A.R.; Kumar, A.; Goswami, D.Y.; Stefanakos, E.K. Nanomaterials for Hydrogen Storage Applications: A Review. J. Nanomater. 2008, 2008, 950967. [Google Scholar] [CrossRef]
  23. Han, S.A.; Sohn, A.; Kim, S.-W. Recent advanced in energy harvesting and storage applications with two-dimensional layered materials. FlatChem 2017, 6, 37–47. [Google Scholar] [CrossRef]
  24. Liu, J.; Cao, H.; Jiang, B.; Xue, Y.; Fu, L. Newborn 2D materials for flexible energy conversion and storage. Sci. China Mater. 2016, 59, 459–474. [Google Scholar] [CrossRef]
  25. Khan, K.; Tareen, A.K.; Aslam, M.; Zhang, Y.; Wang, R.; Ouyang, Z.; Gou, Z.; Zhang, H. Recent advances in two-dimensional materials and their nanocomposites in sustainable energy conversion applications. Nanoscale 2019, 11, 21622–21678. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, Y.-D.; Yu, S.; Zhao, W.-H.; Li, S.-F.; Duan, X.-M. A potential material for hydrogen storage: A Li decorated graphitic-CN monolayer. Phys. Chem. Chem. Phys. 2018, 20, 13473–13477. [Google Scholar] [CrossRef] [PubMed]
  27. Zhou, J.; Wang, Q.; Sun, Q.; Jena, P.; Chen, X.S. Electric field enhanced hydrogen storage on polarizable materials substrates. Proc. Natl. Acad. Sci. USA 2010, 107, 2801–2806. [Google Scholar] [CrossRef] [PubMed]
  28. Gao, P.; Li, J.-w.; Zhang, J.; Wang, G. Computational exploration of magnesium-decorated carbon nitride (g-C3N4) monolayer as advanced energy storage materials. Int. J. Hydrogen Energy 2021, 46, 21739–21747. [Google Scholar] [CrossRef]
  29. Song, N.; Wang, Y.; Sun, Q.; Jia, Y. First-principles study of hydrogen storage on Ti (Sc)-decorated boron-carbon-nitride sheet. Appl. Surf. Sci. 2012, 263, 182–186. [Google Scholar] [CrossRef]
  30. Habibi, P.; Vlugt, T.J.H.; Dey, P.; Moultos, O.A. Reversible Hydrogen Storage in Metal-Decorated Honeycomb Borophene Oxide. ACS Appl. Mater. Interfaces 2021, 13, 43233–43240. [Google Scholar] [CrossRef]
  31. Wang, Y.S.; Wang, F.; Li, M.; Xu, B.; Sun, Q.; Jia, Y. Theoretical prediction of hydrogen storage on Li decorated planar boron sheets. Appl. Surf. Sci. 2012, 258, 8874–8879. [Google Scholar] [CrossRef]
  32. Bull, D.J.; Sorbie, N.; Baldissin, G.; Moser, D.; Telling, M.T.; Smith, R.I.; Gregory, D.H.; Ross, D.K. In situ powder neutron diffraction study of non-stoichiometric phase formation during the hydrogenation of Li3N. Faraday Discuss. 2011, 151, 163–270; discussion 285–295. [Google Scholar] [CrossRef]
  33. Qi, H.; Wang, X.; Chen, H. Superalkali NLi4 decorated graphene: A promising hydrogen storage material with high reversible capacity at ambient temperature. Int. J. Hydrogen Energy 2021, 46, 23254–23262. [Google Scholar] [CrossRef]
  34. Yartys, V.A.; Lototskyy, M.V.; Akiba, E.; Albert, R.; Antonov, V.E.; Ares, J.R.; Baricco, M.; Bourgeois, N.; Buckley, C.E.; Bellosta von Colbe, J.M.; et al. Magnesium based materials for hydrogen based energy storage: Past, present and future. Int. J. Hydrogen Energy 2019, 44, 7809–7859. [Google Scholar] [CrossRef]
  35. Zhang, W.; Zhang, Z.; Zhang, F.; Yang, W. Ti-decorated graphitic-C3N4 monolayer: A promising material for hydrogen storage. Appl. Surf. Sci. 2016, 386, 247–254. [Google Scholar] [CrossRef]
  36. Beheshti, E.; Nojeh, A.; Servati, P. A first-principles study of calcium-decorated, boron-doped graphene for high capacity hydrogen storage. Carbon 2011, 49, 1561–1567. [Google Scholar] [CrossRef]
  37. Gao, P.; Li, J.-w.; Wang, G. Computational evaluation of superalkali-decorated graphene nanoribbon as advanced hydrogen storage materials. Int. J. Hydrogen Energy 2021, 46, 24510–24516. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Liu, P. Hydrogen storage on superalkali NLi4 decorated β12-borophene: A first principles insights. Int. J. Hydrogen Energy 2022, 47, 14637–14645. [Google Scholar] [CrossRef]
  39. Zhang, Y.-F.; Liu, P.-P.; Luo, Z.-H. Superalkali NLi4 cluster decorated γ-graphyne as a promising hydrogen storage material: A density functional theory simulation. FlatChem 2022, 36, 100429. [Google Scholar] [CrossRef]
  40. Wang, X.; Qi, H.; Ma, L.; Chen, H. Superalkali NLi4 anchored on BN sheets for reversible hydrogen storage. Appl. Phys. Lett. 2021, 118, 093902. [Google Scholar] [CrossRef]
  41. Tan, L.; Nie, C.; Ao, Z.; Sun, H.; An, T.; Wang, S. Novel two-dimensional crystalline carbon nitrides beyond gC3N4: Structure and applications. J. Mater. Chem. A 2021, 9, 17–33. [Google Scholar] [CrossRef]
  42. Sharma, P.; Kumar, S.; Tomanec, O.; Petr, M.; Zhu Chen, J.; Miller, J.T.; Varma, R.S.; Gawande, M.B.; Zbořil, R. Carbon Nitride-Based Ruthenium Single Atom Photocatalyst for CO2 Reduction to Methanol. Small 2021, 17, 2006478. [Google Scholar] [CrossRef] [PubMed]
  43. Darkwah, W.K.; Ao, Y. Mini Review on the Structure and Properties (Photocatalysis), and Preparation Techniques of Graphitic Carbon Nitride Nano-Based Particle, and Its Applications. Nanoscale Res. Lett. 2018, 13, 388. [Google Scholar] [CrossRef] [PubMed]
  44. Ahmad, R.; Tripathy, N.; Khosla, A.; Khan, M.; Mishra, P.; Ansari, W.A.; Syed, M.A.; Hahn, Y.-B. Review—Recent Advances in Nanostructured Graphitic Carbon Nitride as a Sensing Material for Heavy Metal Ions. J. Electrochem. Soc. 2020, 167, 037519. [Google Scholar] [CrossRef]
  45. Besharat, F.; Ahmadpoor, F.; Nezafat, Z.; Nasrollahzadeh, M.; Manwar, N.R.; Fornasiero, P.; Gawande, M.B. Advances in Carbon Nitride-Based Materials and Their Electrocatalytic Applications. ACS Catal. 2022, 12, 5605–5660. [Google Scholar] [CrossRef]
  46. Li, J.; Cao, C.; Hao, J.; Qiu, H.; Xu, Y.; Zhu, H. Self-assembled one-dimensional carbon nitride architectures. Diam. Relat. Mater. 2006, 15, 1593–1600. [Google Scholar] [CrossRef]
  47. Chen, X.; Li, J.-w.; Dou, X.; Gao, P. Computational evaluation of Mg-decorated g-CN as clean energy gas storage media. Int. J. Hydrogen Energy 2021, 46, 35130–35136. [Google Scholar] [CrossRef]
  48. Gao, P.; Chen, X.; Li, J.; Wang, Y.; Liao, Y.; Liao, S.; Zhu, G.; Tan, Y.; Zhai, F. Computational evaluation of Al-decorated g-CN nanostructures as high-performance hydrogen-storage media. Nanomaterials 2022, 12, 2580. [Google Scholar] [CrossRef]
  49. 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]
  50. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  51. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B Condens. Matter 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  52. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  53. Sanville, E.; Kenny, S.D.; Smith, R.; Henkelman, G. Improved grid-based algorithm for Bader charge allocation. J. Comput. Chem. 2007, 28, 899–908. [Google Scholar] [CrossRef] [PubMed]
  54. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  55. Zhang, R.; Li, Z.; Yang, J. Two-Dimensional Stoichiometric Boron Oxides as a Versatile Platform for Electronic Structure Engineering. J. Phys. Chem. L 2017, 8, 4347–4353. [Google Scholar] [CrossRef] [PubMed]
  56. Tavhare, P.; Titus, E.; Chaudhari, A. Boron substitution effect on adsorption of H2 molecules on organometallic complexes. Int. J. Hydrogen Energy 2019, 44, 345–353. [Google Scholar] [CrossRef]
Figure 1. (a) Top view and (b) side view of the optimized structure of g-CN monolayer. Red and black balls represent N and C atoms, respectively.
Figure 1. (a) Top view and (b) side view of the optimized structure of g-CN monolayer. Red and black balls represent N and C atoms, respectively.
Nanomaterials 13 00647 g001
Figure 2. The PDOS of pure g-CN monolayer. The PDOS is shown for carbon and nitrogen atoms. Energy calculation was conducted with reference to Fermi energy level.
Figure 2. The PDOS of pure g-CN monolayer. The PDOS is shown for carbon and nitrogen atoms. Energy calculation was conducted with reference to Fermi energy level.
Nanomaterials 13 00647 g002
Figure 3. The first-principles MD study about g-CN monolayer. Temperature (red) and energy (blue) under room temperature (300 k) against time. The time step is 0.5 fs and the total testing is 4 ps, including 8000 steps.
Figure 3. The first-principles MD study about g-CN monolayer. Temperature (red) and energy (blue) under room temperature (300 k) against time. The time step is 0.5 fs and the total testing is 4 ps, including 8000 steps.
Nanomaterials 13 00647 g003
Figure 4. The top view (a) and the side view (b) of the optimized structure of NLi4 decorated on the g-CN monolayer. Red, black, and blue balls are the symbols for N, C, and Li atoms, respectively.
Figure 4. The top view (a) and the side view (b) of the optimized structure of NLi4 decorated on the g-CN monolayer. Red, black, and blue balls are the symbols for N, C, and Li atoms, respectively.
Nanomaterials 13 00647 g004
Figure 5. The PDOS of N, C, and Li atoms of NLi4 decorated on the g-CN monolayer. Energy calculation was conducted with reference to Fermi energy level.
Figure 5. The PDOS of N, C, and Li atoms of NLi4 decorated on the g-CN monolayer. Energy calculation was conducted with reference to Fermi energy level.
Nanomaterials 13 00647 g005
Figure 6. The top view (a) and the side view (b) of the charge density difference for the NLi4 on the g-CN monolayer. Bule and yellow region mean charge loss and gain. The isosurface of charge density is set to 0.0012 (number of charge/Bohr3).
Figure 6. The top view (a) and the side view (b) of the charge density difference for the NLi4 on the g-CN monolayer. Bule and yellow region mean charge loss and gain. The isosurface of charge density is set to 0.0012 (number of charge/Bohr3).
Nanomaterials 13 00647 g006
Figure 7. (af) The top and side views of processing of multiple H2 absorption on optimized NLi4 decorated g-CN monolayer in the increasing order.
Figure 7. (af) The top and side views of processing of multiple H2 absorption on optimized NLi4 decorated g-CN monolayer in the increasing order.
Nanomaterials 13 00647 g007
Figure 8. (a) Top view and (b) side view of charge density difference for the adsorbed H2 molecules on NLi4-decorated g-CN monolayer, where blue and yellow are charge loss and charge gain region, respectively. The isosurface is set to 0.005 (number of charge/Bohr3).
Figure 8. (a) Top view and (b) side view of charge density difference for the adsorbed H2 molecules on NLi4-decorated g-CN monolayer, where blue and yellow are charge loss and charge gain region, respectively. The isosurface is set to 0.005 (number of charge/Bohr3).
Nanomaterials 13 00647 g008
Table 1. Caption. The average adsorption energy per H2, average bond length of H-H, hydrogen storage density and desorption temperature of H2 in g-CN system decorated with NLi4.
Table 1. Caption. The average adsorption energy per H2, average bond length of H-H, hydrogen storage density and desorption temperature of H2 in g-CN system decorated with NLi4.
SystemsEad (eV)RH-H Length (A)HSC (wt%)Td (k)
1NLi4@g-CN 1H2−0.2370.760.30303
1NLi4@g-CN 8H2−0.1770.762.42227
1NLi4@g-CN 9H2−0.1700.762.70217
2NLi4@g-CN 18H2−0.1550.765.10198
3NLi4@g-CN 27H2−0.1520.767.30194
4NLi4@g-CN 36H2−0.1550.769.20198
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

Chen, X.; Hou, W.; Zhai, F.; Cheng, J.; Yuan, S.; Li, Y.; Wang, N.; Zhang, L.; Ren, J. Reversible Hydrogen Storage Media by g-CN Monolayer Decorated with NLi4: A First-Principles Study. Nanomaterials 2023, 13, 647. https://doi.org/10.3390/nano13040647

AMA Style

Chen X, Hou W, Zhai F, Cheng J, Yuan S, Li Y, Wang N, Zhang L, Ren J. Reversible Hydrogen Storage Media by g-CN Monolayer Decorated with NLi4: A First-Principles Study. Nanomaterials. 2023; 13(4):647. https://doi.org/10.3390/nano13040647

Chicago/Turabian Style

Chen, Xihao, Wenjie Hou, Fuqiang Zhai, Jiang Cheng, Shuang Yuan, Yihan Li, Ning Wang, Liang Zhang, and Jie Ren. 2023. "Reversible Hydrogen Storage Media by g-CN Monolayer Decorated with NLi4: A First-Principles Study" Nanomaterials 13, no. 4: 647. https://doi.org/10.3390/nano13040647

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