Yi-Sha
Chen
,
Jing-Jing
Guo
,
Peng-Bo
Liu
,
Hui-Yan
Zhao
*,
Jing
Wang
and
Ying
Liu
*
Department of Physics and Hebei Advanced Thin Film Laboratory, Hebei Normal University, Shijiazhuang 050024, Hebei, China. E-mail: hyzhao@hebtu.edu.cn; yliu@hebtu.edu.cn
First published on 23rd October 2025
Metallo-borospherenes, a rapidly evolving class of boron-based nanostructures, exhibit intriguing diversity in both geometric architectures and electronic properties. Recently, the experimental identification of the D3h-symmetric Ln3B18− (Ln = La, Tb) clusters marked a significant milestone in this emerging field. In this study, a new family of metallo-borospherenes M12B60 (M = Y, Lu) was predicted using first-principles calculations. Ab initio molecular dynamics simulations and vibrational frequency analyses confirm their thermodynamic and kinetic stability. Detailed bonding analysis reveals the presence of 108 delocalized multi-center two-electron bonds within the cage, accompanied by pronounced aromatic character, which collectively account for their exceptional stability. Notably, the Y12B60 cage exhibits potential as a promising hydrogen storage material. Its curved surface provides multiple favorable adsorption sites, enabling H2 molecules to form layered distributions. The Y12B60 cage can adsorb up to 89 H2 molecules with an average adsorption energy of −0.201 eV per H2, corresponding to a gravimetric density of 9.44 wt%, meeting practical adsorption requirements. Ab initio molecular dynamics simulations further validate the structural robustness and adsorption capacity of Y12B60 under near-ambient conditions, underscoring its promise for reversible hydrogen storage applications.
Metal atom doping has emerged as a powerful strategy to compensate for boron's electron deficiency, significantly altering geometries and introducing novel chemical bonding patterns and properties.5–7 In 2013, inspired by the perfect molecular wheels B8− and B9−,8,9 a design principle for metal-centered boron rings (M©Bn) was proposed, where the central boron atom was replaced with a transition metal. This led to the prediction of several dually aromatic species, including D8h Co©B8−, D9h Ru©B9− and D10h Ta©B10−.10 More recently, this concept has been extended to M(m)©Bnk− clusters, culminating in the discovery of the D11h symmetry YB112− structure.11 Representative examples of single-metal doping included dually anti-aromatic compound BiB6− with planar configuration, dually aromatic compound BiB7− and BiB8−,12 half-sandwich configuration LaBn−/0 (n = 14–17)13 and twenty-coordinate boron molecular drums An@B20 (An = U, Np, and Pu).14
By doping two metal atoms, the icosahedral boron-based clusters B10M2 (M = Rh, Ir)15 were first predicted. Additionally, it has been observed that Bn− (n = 10, 11) doped by 2 La atoms could form the anti-sandwich complex La2Bn− (n = 10, 11), offering a novel pathway for the structural evolution of larger lanthanide boron clusters.16 A series of bimetal-doped inverse sandwich boron clusters, including Ta2B6−/0, Pr2B8−, and La2Bn− (n = 7–9), as well as B10La2− and B11La2− extended inverse-sandwich complexes, have also been experimentally observed.16–18
As the number of doped metal and boron atoms increased, the geometric configuration gradually evolved to metallo-borospherenes. The experimental discovery of trimetal-doped metallo-borospherenes D3h La3B18− marked a breakthrough advancement in the field of metallo-borospherenes.19 Subsequent studies revealed its potential as a hydrogen storage material, capable of adsorbing 18 H2 molecules with a gravimetric density of 5.6 wt%.20 Inspired by the D3h La3B18− structure, a core–shell metallo-borospherene, D3h La3&[B2@B18]−, sharing the same shell configuration was proposed.21 Further exploration with increased metal doping led to the proposal of tetrahedral cage Td La4B24 (exhibiting spherical aromaticity) and core–shell structures Td La4B290/+/−, and La@[La5&B30]0/−/2−, enriching the family of metallo-borospherenes.22,23 Most recently, spherical trihedral metallo-borospherenes B20TMn (TM = Sc, Y; n = 3, 4) were reported. Their B20 framework consisted of three B2 units bridging two equivalent B7 triangles, providing a B7 triangular motif for the B-framework.24
Metal-doped boron nanomaterials have been extensively studied for hydrogen storage. Commonly investigated metals included alkali (Li, Na, and K),25–28 alkaline earth (Be, Mg, and Ca),29–32 transition metals (Sc, Ti, Y, Fe, Co, and Ni),33–38 and rare earth elements (La, Ce, Nd, Gd, and Er).20,39–41 Notably, Ti-decorated B38 cages can adsorb up to 6 H2 molecules per Ti atom, achieving a gravimetric capacity of 7.44 wt%.42 Recently, Y3C3B9H12 was found to bind 14 H2 molecules, yielding a storage capacity of 6.4 wt%.37 Theoretical predictions indicated that C48B12 decorated with 12 Y atoms is an excellent hydrogen storage material. Each Y atom adsorbed 6 H2 molecules primarily through Kubas interactions, with an average adsorption energy of −0.64 eV per H2 and a corresponding gravimetric density of 7.51 wt%.43 The hydrogen storage behavior of multi-yttrium-doped B38 systems has been explored using first-principles calculations. These studies revealed that Y atoms preferentially occupied the hexagonal cavities of B38 rather than forming clusters, effectively preventing Y atoms from diffusion. These results confirmed that Y4@B38 is a promising hydrogen storage material, with each Y atom adsorbing up to 6 H2 molecules, corresponding to a gravimetric density of 4.96 wt%.38
While prior work on metallo-borospherenes has largely concentrated on endohedral and exohedral motifs, investigations of hollow, cage-like structures remained relatively scarce. Notably, the experimentally characterized La3B18− cage-like motif—conceived from three shared La©B10 units—provided an important proof of concept for constructing larger cage architectures.19 In addition, previous theoretical studies of yttrium-decorated boron nanostructures (e.g., Y©B112− molecular wheel,11 B20Y3/Y4 metallo-borospherenes,24 and Y4@B38 hydrogen-storage candidates38) demonstrated that Y can stabilize a variety of boron frameworks and exhibit favorable properties for hydrogen adsorption. These findings motivated us to explore whether a Y©B10 building block can be extended to form a highly symmetric D2h Y12B60 cage assembled from twelve shared Y©B10 units; subsequent calculations indicated that this structure was energetically and kinetically viable and possessed promising hydrogen storage characteristics.
To probe the role of the metal center, we selected Lu as a comparative element. Y and Lu shared similar valence electron counts (both commonly trivalent), but Lu had a filled 4f shell and experienced a pronounced lanthanide contraction, resulting in a smaller atomic radius. This combination—similar valence electronic structure but different ionic/atomic sizes—provided a controlled way to disentangle size (steric) effects from purely electronic effects on the cage geometry and stability. Consequently, comparing Y12B60 and Lu12B60 yields direct insight into how metal-center size influences structural parameters and hydrogen adsorption behavior.
The hydrogen storage properties were evaluated employing the WB97XD functional47 in the Gaussian 16 package,48 which included empirical dispersion corrections and provided a more reliable description of weak interactions.
Thermodynamic stability was further assessed by performing ab initio molecular dynamics (AIMD) simulations within the canonical (NVT) ensemble at 300 K, as implemented in the DMol3 package.49 The simulations were run for a total of 5.0 ps using a time step of 1.0 fs.
Additional property calculations, including Raman spectra, photoelectron spectroscopy (PES), adaptive natural density partitioning (AdNDP) calculations, and interaction region indicator (IRI) were conducted using the Gaussian 16 package.48 Data processing and visualization were conducted using Multiwfn50 and VMD,51 respectively.
![]() | ||
| Fig. 1 Optimized structures (a) Y12B60 and (b) Lu12B60. Boron atoms were shown in pink; yttrium atoms, in cyan; and lutetium atoms, in green. | ||
Following geometry optimization, average bond lengths for the Y12B60 structure were determined. The average Y–B bond length was 2.68 Å. The average B–B bond length was 1.63 Å, intermediate between typical B
B double bonds (about 1.56 Å) and B–B single bonds (about 1.70 Å). For Lu12B60, the average Lu–B bond length was 2.63 Å, while the average B–B bond length was 1.63 Å. Additionally, the average binding energy per atom of Lu12B60 was −5.48 eV per atom, slightly lower than that of Y12B60 (−5.58 eV per atom), indicating lower stability in Lu12B60. This originated from Lu's larger atomic radius, which led to a less optimal match with the B10 rings in the B60 cage. Consequently, Lu12B60 adopted a lower Cs symmetry compared to the D2h-symmetric Y12B60. Given its superior stability and symmetry, the D2h Y12B60 structure was the focus of subsequent analysis.
![]() | ||
| Fig. 2 Representative snapshots from AIMD simulations of the Y12B60 structure at 800 K, 1000 K and 1200 K under the NVT ensemble. The total simulation time was 5.0 ps with a time step of 1.0 fs. | ||
Subsequently, to further assess their relative stability, a set of low-energy isomers extracted from high-temperature AIMD trajectories were fully reoptimized. The energy differences (ΔE) between these isomers and the primary structures are summarized in Fig. S4 and S5 of the SI. For Y12B60, isomers 1 and 2 exhibited slightly lower total energies than the primary D2h structure, accompanied by reduced symmetry and partial disruption of surface Y©B10 units, which evolved into B6 nanoribbon-like motifs. In contrast, isomers 3–6, which retained intact Y©B10 units, were slightly higher in energy. A similar energetic ordering was observed for Lu12B60. Although the structures discussed in the manuscript were not the absolute global minima, they represented low-lying metastable isomers characterized by intact Y©B10/Lu©B10 motifs, high symmetry, and pronounced kinetic stability—features that were crucial for potential experimental realization and functional applications.
The Raman spectra and infrared (IR) of the Y12B60 and Lu12B60 structures were simulated at the PBE0 level. While the frequency calculation provided all normal vibrational modes, only those that are IR- or Raman-active appeared as peaks in the corresponding spectra. The simulated spectra therefore served not only as a theoretical reference for future experimental characterization but also as a complementary verification of the stability of the predicted structures. Raman spectra (Fig. S6) revealed the frequencies of Y12B60 and Lu12B60 structures spanning 39.7 to 1405.1 cm−1 and 39.6 to 1427.1 cm−1, respectively, with no imaginary frequencies presented, confirming the good kinetic stability. For the Y12B60 structure, major vibration modes corresponding to specific spectral features are detailed in Fig. S7. The dominant peak at 61.4 cm−1 corresponded to the global breathing mode of the entire cage structure, while prominent bands observed at 279.6, 603.4, 937.9, and 1346.4 cm−1 originated from localized stretching vibrations of B3 triangles. For infrared (IR) spectra (Fig. S6), the Y12B60 structure exhibited several major asymmetric infrared activity peaks. Among them, the first infrared peak at 104.0 cm−1 and the highest peak at 845.6 cm−1 mainly corresponded to localized stretching vibrations of B3 triangles.
Photoelectron spectroscopy (PES) is a powerful analytical technique that probes the elemental composition, chemical state, and electronic structure of matter by measuring the kinetic energies of electrons ejected upon irradiation with photons, with particular sensitivity to surface properties. The simulated PES offers valuable predictive insights for subsequent experimental studies. The simulated photoelectron spectrum of the Y12B60− structure (Fig. S6) displayed distinct features that were expected to be readily detectable in experimental measurements. The first vertical detachment energy (VDE) was calculated to be 1.56 eV, which corresponded to the energy difference between the anionic ground state and the neutral species at the anion geometry. The first adiabatic detachment energy (ADE) was found to be 1.44 eV, representing the energy difference between the lowest potential energy surfaces of the anion and the neutral molecule. The small difference between the VDE and ADE (0.12 eV) suggested that the geometric structures of the neutral molecule and the anion in their ground states were likely similar. Therefore, it was inferred that the first ionization band may exhibit a relatively sharp profile, possibly with vibrational fine structure. Furthermore, the noticeable intervals (>0.2 eV) among the computed VDE values, including 1.82, 2.10, 2.70, 3.02, 3.32, 3.67, 4.11, 4.87, 5.14, 5.42, 5.65 and 6.06 eV, indicated that the corresponding electronic states were expected to be resolved as distinct bands in the experimental spectrum. These predicted spectral features may provide useful theoretical guidance for future characterization of the Y12B60− structure using photoelectron spectroscopy in experimental studies.
Recent studies have demonstrated that the spherical trihedral metallo-borospherene La3B18− has been successfully realized through a combination of theoretical prediction and experimental photoelectron spectroscopy, where three La©B10 structural units formed the stable cage.19 In an analogous manner, the Y12B60 (Lu12B60) cluster can be regarded as being constructed from twelve such Y©B10 (Lu©B10) subunits, thereby exhibiting both good thermodynamic stability and strong structural similarity to the experimentally confirmed La3B18−.19
Based on structural analogies to previously reported metallo-borospherenes, it was conceivable that M12B60 (M = Y, Lu) could potentially be generated experimentally through laser ablation of a Y/B (Lu/B) mixed target, followed by supersonic expansion in a helium carrier gas—the same method employed in the synthesis of La3B18−.19
![]() | ||
| Fig. 3 Total density of states (TDOS) and projected density of states (PDOS) of the Y12B60 structure. The HOMO and LUMO energy levels were indicated by pink and orange dashed lines, respectively. | ||
The deformation electron density was analyzed to clarify the electronic properties of the Y12B60 structure (Fig. 4). Regions of charge accumulation and depletion were colored in blue and yellow, respectively. The results revealed significant charge depletion around the metal atoms and B atoms. Charge accumulation occurred between the Y and B atoms, as well as among the B atoms, indicating the presence of multi-center two-electron bonds. This finding was consistent with the subsequent AdNDP analysis. The Hirshfeld charge analysis further confirmed an average transfer of 0.69e from the Y atom to the B atom.
![]() | ||
| Fig. 4 Deformation electron density for the Y12B60 structure, with an isosurface of 0.025 e Å−3. The blue and yellow represented the accumulation and depletion of electrons, respectively. | ||
Adaptive natural density partitioning (AdNDP) was a method of chemical bonding analysis that described the electronic structure of a molecule by decomposing its electron density (typically derived from computations) into a set of chemically meaningful units, such as lone pairs, two-center two-electron bonds, and multi-center bonds. Therefore, to further investigate the bonding properties of the M12B60 (M = Y, Lu) structures, AdNDP bonding analysis was conducted. A total of 108 multi-center two-electron bonds were identified within the Y12B60 framework. As shown in Fig. 5, Y12B60 possessed thirty localized 2c–2e σ bonds on the B–B bonds shared by adjacent B10 rings. Twenty 3c–2e σ bonds were located on the B3 triangular units, and another forty 3c–2e σ bonds on the two B atoms and one Y atom, where the B atoms originated from the two B atoms in the B3 triangles. Fourteen 8c–2e σ bonds are located between the six B atoms and the two adjacent Y atoms. At the drum-face positions of the Y12B60 structure, there were four 9c–2e σ bonds between the six B atoms and the three Y atoms. The existence of these covalent bonds enabled the system to maintain higher stability. Previous studies have indicated that the Lu 4f orbitals were fully occupied and their energy levels were significantly lower than those of the 5d orbitals, rendering them largely non-participatory in bonding interactions.53 Therefore, the Lu12B60 structure has similar bonding characteristics, as shown in Fig. S12.
![]() | ||
| Fig. 5 AdNDP analysis of the Y12B60 structure. The occupation numbers were indicated for each bonding pattern, and the isosurface was set to 0.05 e Å−3. | ||
The nucleus-independent chemical shifts (NICSs) provided a reliable assessment of molecular aromaticity. The NICS(0) value of the Y12B60 structure, calculated at the PBE0 level, was −18.19 ppm, confirming the Y12B60 structure's aromatic character. The calculated ICSS_ZZ of the Y12B60 structure reflected the isosurface of the ZZ component of the NICS, providing a better understanding of aromaticity. The isochemical shielding surface (ICSS) method can be considered as a three-dimensional extension of the nucleus-independent chemical shift (NICS) concept. By calculating the magnetic shielding values at grid points in space and visualizing the corresponding isosurfaces, this approach offered a more intuitive depiction of the spatial distribution of magnetic response. In particular, ICSS_ZZ denoted the isosurface associated with the ZZ component of the NICS. A principal advantage of the ZZ component lies in its enhanced capacity to differentiate between long-range magnetic effects contributed by delocalized π-electrons and local magnetic anisotropy arising from the σ-framework. This distinction facilitated the identification of magnetically shielded regions that were indicative of aromatic character. As shown in Fig. 6, distinct green regions represented the chemical shielding area, while the yellow regions indicated the chemical de-shielding area. The analysis revealed a chemical shielding region along the radial z direction and a chemical de-shielding area in the horizontal direction surrounding the structure. These results collectively confirmed the aromaticity of the Y12B60 structure, which represents a key factor underpinning the good stability.
![]() | ||
| Fig. 6 Side view (a) and top view (b) of the ICSS_ZZ for the Y12B60 structure. Green and yellow represented chemical shielding and de-shielding areas, respectively. | ||
To determine the hydrogen storage capacity of the Y12B60 structure, 150 H2 molecules were placed around it and structural optimization was conducted. The results revealed that the H2 molecules formed distinct adsorption layers (Fig. 7 and 8). The first adsorption layer, located 6.60–7.00 Å from the structure center, contained about 38 H2 molecules with an average adsorption energy of −0.253 eV per H2. The second adsorption layer was positioned 7.30–7.81 Å from the Y12B60 structure center (corresponding to an interlayer distance of 0.70–0.81 Å relative to the first layer), which was significantly shorter than the 2.7 Å equilibrium distance between free H2 molecules, thereby confirming the adsorption of H2 molecules by the Y12B60 structure. Two layers accommodated about 89 H2 molecules with an adsorption energy of −0.201 eV per H2, achieving a gravimetric hydrogen storage density of 9.44 wt% and a corresponding desorption temperature of 256 K based on the Van’t Hoff equation. Furthermore, the structural schematic diagram of the Y12B60-38 H2 (the first layer) and the Y12B60-89 H2 systems has been provided to more intuitively display the molecular arrangements within each layer (Fig. S15).
![]() | ||
| Fig. 7 The distributions of the number density of B, Y, and H atoms, and the total number of H2 molecules, as a function of distance from the Y12B60 structure center. | ||
Quantum theory of atoms in molecules (QTAIM) provided a robust framework for analysing chemical bonding. To investigate hydrogen adsorption between H2 molecules and the Y12B60 structure, we performed QTAIM topological analysis at four adsorption sites listed in Table S3. For non-covalent interactions, the Laplacian of electron density ∇2ρ(r) exhibited positive values. Higher electron density values ρ(r) correlated with more negative potential energy density V(r), signifying stronger interactions. Obviously, positive ∇2ρ(r) values were characteristic of non-covalent interactions between H2 and the Y12B60 framework. The adsorption site above the Y–B bond showed the strongest binding, consistent with the calculated adsorption energies.
The interaction region indicator (IRI) was a real-space function derived from the electron density and its derivatives. It was employed to visualize and qualitatively analyse diverse chemical interactions in molecular systems, encompassing covalent bonds, hydrogen bonds, van der Waals forces, and steric repulsions. Accordingly, to further investigate the hydrogen adsorption mechanism between H2 molecules and the Y12B60 structure, we conducted interaction region indicator (IRI) and partial density of states (PDOS) analysis for the Y12B60-89 H2 system (Fig. 9 and Fig. S16, S17). Weak interactions in Fig. 9 regions were coloured in green. Between the Y12B60 structure and the adsorbed 89 H2 molecules were surrounded by green regions, confirming that hydrogen adsorption was primarily mediated by weak interactions. This finding was supported by PDOS analysis (Fig. S16), which revealed weak hybridization between the H 1s orbitals and Y 5d orbitals within the −14.5 to −11 eV energy range. Furthermore, no significant hybridization peaks (Fig. S17) were observed between the Y 5d orbitals and H 1s orbitals near the Fermi level, providing additional evidence that the adsorption between Y12B60 and H2 molecules occurred via van der Waals interactions. Most importantly, the H–H bond lengths for all adsorbed H2 molecules in both layers remained in the range of 0.75–0.76 Å. Although slightly elongated compared to that of the free H2 molecule (0.74 Å), no significant bond cleavage occurred, providing key evidence for a mechanism dominated by physical adsorption.
![]() | ||
| Fig. 9 The geometric structure (a) and IRI analysis (b) of the Y12B60-89 H2 complex. The weak interaction regions were coloured green. | ||
To elucidate the desorption mechanism of hydrogen molecules, we performed ab initio molecular dynamics simulations on the Y12B60-89 H2 system at different temperatures (250 K, 300 K, 350 K, 400 K, 500 K, 600 K), with corresponding statistical results presented in Fig. 10. The simulation duration was set to 5.0 ps with a time step of 0.5 fs to accurately capture the high frequency vibrations of hydrogen molecules. Besides, the analysis of the energy and temperature evolution confirmed that the system reached a well-equilibrated state (Fig. S18), indicating the reliability of the computational results.
![]() | ||
| Fig. 10 The number of adsorbed H2 molecules as a function of AIMD simulation temperature for the Y12B60-89 H2 complex. | ||
The simulations revealed a clearly temperature-dependent adsorption/desorption behavior. At 250 K, the system exhibited a low desorption rate, retaining ∼84 H2 molecules (8.97 wt%). Increasing the temperature to 300 K reduced the adsorption to ∼66 H2 molecules (7.19 wt%), while at 350 K the number decreased further to ∼53 H2 molecules (5.85 wt%). At 600 K, nearly complete desorption occurred, with only ∼8 H2 molecules remaining adsorbed. These results indicated that the desorption process was effectively activated with increasing temperature.
Importantly, according to the U.S. Department of Energy (DOE) target, a practical hydrogen storage material should achieve a gravimetric capacity of ≥5.5 wt% near ambient temperature (∼300 K). The Y12B60 cluster maintains a capacity above this threshold within the 250–350 K range, demonstrating its potential for practical hydrogen storage applications. The temperature-activated desorption further supported the material's applicability.
| This journal is © the Owner Societies 2025 |