Next Article in Journal
In Vitro and In Vivo Antibiofilm Activity of Red Onion Scales: An Agro-Food Waste
Next Article in Special Issue
F2BMF (M = B and Al) Molecules: A Matrix Infrared Spectra and Theoretical Calculations Investigation
Previous Article in Journal
Metal-Free Aerobic C–N Bond Formation of Styrene and Arylamines via Photoactivated Electron Donor–Acceptor Complexation
Previous Article in Special Issue
Planar Elongated B12 Structure in M3B12 Clusters (M = Cu-Au)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Evolution and Electronic Properties of Selenium-Doped Boron Clusters SeBn0/− (n = 3–16)

1
School of Physics and Electronic Science, Guizhou Education University, Guiyang 550018, China
2
College of Physics, Guizhou University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(1), 357; https://doi.org/10.3390/molecules28010357
Submission received: 9 November 2022 / Revised: 21 December 2022 / Accepted: 27 December 2022 / Published: 1 January 2023
(This article belongs to the Special Issue New Boron Chemistry: Current Advances and Future Prospects)

Abstract

:
A theoretical research of structural evolution, electronic properties, and photoelectron spectra of selenium-doped boron clusters SeBn0/− (n = 3–16) is performed using particle swarm optimization (CALYPSO) software in combination with density functional theory calculations. The lowest energy structures of SeBn0/− (n = 3–16) clusters tend to form quasi-planar or planar structures. Some selenium-doped boron clusters keep a skeleton of the corresponding pure boron clusters; however, the addition of a Se atom modified and improved some of the pure boron cluster structures. In particular, the Se atoms of SeB7, SeB8, SeB10, and SeB12 are connected to the pure quasi-planar B7, B8, B10, and B12 clusters, which leads to planar SeB7, SeB8, SeB10, and SeB12, respectively. Interestingly, the lowest energy structure of SeB9 is a three-dimensional mushroom-shaped structure, and the SeB9 cluster displays the largest HOMO–LUMO gap of 5.08 eV, which shows the superior chemical stability. Adaptive natural density partitioning (AdNDP) bonding analysis reveals that SeB8 is doubly aromatic, with 6 delocalized π electrons and 6 delocalized σ electrons, whereas SeB9 is doubly antiaromatic, with 4 delocalized π electrons and 12 delocalized σ electrons. Similarly, quasi-planar SeB12 is doubly aromatic, with 6 delocalized π electrons and 14 delocalized σ electrons. The electron localization function (ELF) analysis shows that SeBn0/− (n = 3–16) clusters have different local electron delocalization and whole electron delocalization effects. The simulated photoelectron spectra of SeBn (n = 3–16) have different characteristic bands that can identify and confirm SeBn (n = 3–16) combined with future experimental photoelectron spectra. Our research enriches the geometrical structures of small doped boron clusters and can offer insight for boron-based nanomaterials.

1. Introduction

Clusters are composed of several to thousands of atoms or molecules whose properties depend on their size and shape. Clusters are ideal model systems for correlating microscopic structure and macroscopic properties of substances, and cluster research is of great significance to deeply understand the laws of matter transformation. Research of structure and properties of clusters can offer insight for the design and manufacture of new materials and new devices at the atomic level. Boron clusters can induce polycentric chemical bonds and adopt several interesting structures with meaningful properties [1,2,3,4,5,6]. Experimental and theoretical research has shown that Bn (n < 38) have quasi-planar or planar structures [7,8], and neutral Bn have quasi-planar, planar, wheel shaped, tubular structures or other structures [1,5,9,10,11]. In 2014, the experimental finding of cage-type all-boron cluster (borospherene) [12] B40 has given rise to a lot of attention on boron clusters [13,14,15,16,17,18,19,20,21]. In 2015, researchers synthesized borophene on the Ag (111) base [22], and the structural unit of borophene is a B7 cluster. In 2021, researchers synthesized borophene crystal, which was hydrogenated with hydrogen atoms [23], and it is very stable and comparable to graphene. Interestingly, the basic unit of hydrogenated borophene happens to be a hydrogenated B7 cluster also. The experimental findings of B40 and borophene offer insight for the development of new boron nanomaterials and nanodevices. Research on small boron clusters is promising as a way to provide new ideas for new nanomaterials and nanodevices.
In recent years, researchers have studied abundant doped boron clusters, and they mainly focused on doping a single metal atom in boron clusters of different sizes. Metal-atom-doped boron clusters can induce new geometrical configuration and properties [15,17,18,19,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38]. For example, anionic B20 and B22 have a quasi-planar structure [7,39]. However, single alkali-metal-atom-doped LiB20, NaB22, and KB22 display a double-ring configuration [25,40]. Co and Rh atom-doped boron clusters MB12 (M = Co and Rh) can lead to the quasi-planar B12 form, a semi-sandwich configuration [3,24]. Neutral B24 has a double-ring configuration [39], while TiB24 and ScB24 have a cage configuration and three-ring tubular structure, respectively, after adding one Sc or Ti atom [41,42]. Anionic B24 has a quasi-planar structure [7], while TiB24 and VB24 have a cage configuration after adding one Ti or V atom [43]. In addition, the LiB40, NaB40, or KB40 cage is promising for application to the field of nonlinear optics [18]; the ScB40 or TiB40 cage is promising for application to the field of hydrogen storage [15,17,19]; Co atom-doped CoB12 and Rh atom-doped RhB12 can improve chemical activity [27]; Co atom-doped CoB40 is promising for application to molecular devices [26]; and the metal-atom-doped boron clusters ReBn(n = 3–4, 6, 8–9), MnBn(n = 6, 16), BiBn(n = 6–8), CoB16, La2Bn(n = 10–11), La3B18, MB8(M = Be, Mg), and M2B6(M = Mg, Ca, Sr) have various unique structures [29,30,31,32,33,34,35,36,37,44,45]. However, nonmetallic-atom-doped boron clusters have been poorly studied. In particular, the structural evolution of boron clusters after addition of a single nonmetallic atom is rarely studied. Similar to B7, small boron clusters with doping are promising as the structural units of borophene and other boron nanomaterials. Selenium is one of the essential microelements in the human body and has obvious inhibitory effects on tumors. The Se atom can combine with metal atoms, such as Cd and Zn, to form semiconductor clusters or quantum dots [46,47]. These materials exhibit a variety of unique optical and electronic properties, which can be further applied in imaging and diagnosis of biological systems [48,49]. Se doping of boron clusters is promising to be a useful strategy to further increase the diversity of structural forms and to affect the properties of boron clusters. Therefore, the theoretical research of Se atom-doped small boron clusters can enrich new structures and new properties of boron clusters, and can also provide theoretical guidance for the synthesis of nanomaterials, such as borophene. Herein, to demonstrate the structural evolution of Se-doped SeBn0/− (n = 3–16), extensive geometric configurations were generated and predicted, using the particle swarm optimization (CALYPSO) approach [50] in combination with the density functional theory method PBE0 [51].

2. Results and Discussion

2.1. Structures and Electronic Properties

Five low-energy structures of SeBn0/− (n = 3–16) are shown in Figures S1–S28, and the lowest energy structures of SeBn0/− (n = 3–16) are displayed in Figure 1 and Figure 2. The results indicate that the low-energy structures of SeBn0/− (n = 3–16) trend to form planar or quasi-planar structures. Interestingly, the ground state configuration of SeB9 is three-dimensional mushroom shaped. Early theoretical and experimental research found that most of the small neutral boron clusters are quasi-planar or planar, and all small anionic boron clusters are quasi-planar or planar structures. Figure 1 and Figure 2 and the research results indicate that, after adding a Se atom, some of the lowest energy configurations of SeBn0/− (n = 3–16) have a skeleton of pure boron clusters, such as SeBn (n = 3–5, 7–8, 10–14, 16) and SeBn (n = 3–5, 7–8, 10–14, 16) [52,53]. As can be seen in Figure 1 and Figure 2, except for SeBn (n = 5, 6, 9), each Se atom is connected to two boron atoms to form a three-ring. The lowest energy configurations of SeBn0/− (n = 3–5) have planar structure, and the Se atom is attached to the boron atoms of the pure boron cluster Bn0/− (n = 3–5) [52]. The lowest energy configurations of SeB60/− and SeB130/− have quasi-planar structure, and they are different from the ground-state structures of corresponding pure boron clusters B60/− and B130/− [52]. For SeB70/−, SeB100/−, and SeB120/−, the lowest energy configurations of neutral clusters and corresponding anionic clusters have similar structure, and the Se atoms of SeB7, SeB10, and SeB12 are connected to the pure quasi-planar B7, B10, and B12 clusters, respectively. However, the Se atoms of SeB7, SeB10, and SeB12 are connected to the pure quasi-planar B7, B10, and B12 clusters, which leads to the planar SeB7, SeB10, and SeB12, respectively [52]. The lowest energy structures of SeB80/− have same planar structure, and the Se atom of SeB8 is connected to the pure planar B8 cluster. However, the Se atom of SeB8 is connected to the pure quasi-planar B8 cluster, which leads to the planar SeB8. The pure B90/− have same planar wheel-shape structure [52]. However, doping of Se atom causes the anionic SeB9 to become a three-dimensional mushroom-shaped structure (with C6V symmetry) and causes the neutral SeB9 to become a boat-shaped structure. The lowest energy structures of SeB110/− have similar structure, and the Se atom of SeB11 is connected to the pure planar B11 cluster, which leads to the slight structural change. However, the Se atom of SeB11 is connected to the pure planar B11 cluster [52]. The lowest energy configurations of SeB140/− have same planar structure, the lowest energy configurations of SeB160/− have same quasi-planar structure, and the Se atoms are connected to the pure planar B140/− clusters and pure quasi-planar B160/− clusters, respectively [52,53]. The lowest energy configurations of SeB150/− have quasi-planar structure and exhibit two axially chiral isomers, and they are different from the ground-state configurations of pure boron cluster B150/−. Similar to pure B7 clusters, planar and quasi-planar Se-doped boron clusters are promising as the structural units of boron nanomaterials, which can be synthesized further into borophene.
The harmonic frequency analysis confirmed that these lowest-energy structures are actually stable (no imaginary frequency). For closed-shell clusters, the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) energy gaps (HOMO–LUMO energy gaps) of SeB3, SeB4, SeB5, SeB6, SeB7, SeB8, SeB9, SeB10, SeB11, SeB12, SeB13, SeB14, SeB15, and SeB16 are 2.48, 3.52, 2.14, 3.92, 2.90, 2.33, 5.08, 2.93, 2.89, 3.29, 2.28, 2.00, 2.91, and 2.27 eV, respectively. For open-shell clusters, α-HOMO–LUMO and β-HOMO–LUMO, energy gaps vary within the range of 1.92-4.39 eV. Meanwhile, the calculated HOMO–LUMO energy gaps of SeBn0/− (n = 3–16) clusters reveal that the SeB9 cluster possesses the largest HOMO-LUMO gap of 5.08 eV, which shows the superior chemical stability.
To further understand the stability of typical structures of SeBn0/− (n = 3–16), we analyzed the chemical bonding of closed-shell planar SeB8, three-dimensional SeB9, and quasi-planar SeB12 using the adaptive natural density partitioning (AdNDP) approach. Figure 3 displays the bonding patterns of planar SeB8. For SeB8, AdNDP analyses reveal that one lone pair (Figure 3a) is found on the Se atom and eight 2c–2e σ bonds (Figure 3b) on the peripheral ring. The remaining six bonds contain three σ bonds and three π bond, which are classified into four sets (Figure 3): one 3c–2e σ bond covers one B3 triangle, in which two of the boron atoms are connected to the Se atoms; two 4c–2e σ bonds cover two B4 rings; one 3c–2e π bond is distributed around the B-Se-B triangle; and two 5c–2e π bonds cover two B5 rings. Overall, the eight 2c–2e σ bonds, one delocalized 3c–2e σ bond, and two delocalized 4c–2e σ bonds cover the planar molecule, which renders stability to the SeB8, and the delocalized 3c–2e π bond and 5c–2e π bonds further stabilize the SeB8 cluster. Figure 4 displays the bonding patterns of SeB9, and there are five categories. First, there is one lone pair (see Figure 4a) on the Se atom. Then, there are two 2c–2e σ bonds on the Se-B and adjacent B-B. Third, two 2c–2e π bonds cover the Se-B symmetrically. Fourth, the peripheral B7 ring at the top of the mushroom is characterized by six localized B−B 2c–2e σ bonds. The last six delocalized 3c–2e σ bonds cover the inner B3 triangles at the top of the mushroom. The six localized B−B 2c-2e σ bonds and six delocalized 3c–2e σ bonds are responsible for the connection between the outer B7 ring and the inner B atom at the top of the mushroom, which enhances the stability of SeB9. Figure 5 displays the bonding patterns of quasi-planar SeB12. For SeB12, AdNDP analyses reveal that one lone pair (see Figure 5a) is found on the Se atom and ten 2c–2e σ bonds (see Figure 5b) on the peripheral ring. The remaining ten bonds contain seven σ bonds and three π bonds, which are classified into three sets (Figure 5): one 3c–2e π bond covers B-Se-B triangle, seven 3c–2e σ bonds cover the seven inner B3 triangles, and two 5c–2e π bonds are distributed symmetrically around the two B5 rings. Similar to the SeB8, the ten 2c–2e σ bonds and seven delocalized 3c–2e σ bonds cover the quasi-planar molecule, which renders stability to the SeB12 cluster, and the delocalized 3c–2e π bond and 5c–2e π bonds further stabilize the SeB12 cluster. AdNDP bonding analyses revealed that the SeB8 and SeB12 possess three delocalized π bonds, which, quite surprisingly, satisfy the 4m + 2 Hückel rule for π aromaticity. Furthermore, SeB8 and SeB12 possess three delocalized σ bonds and seven delocalized σ bonds, which satisfy the 4m + 2 Hückel rule for σ aromaticity. However, the three-dimensional SeB9 cluster possesses two delocalized π bonds and six delocalized σ bonds, which satisfy the 4m Hückel rule for π and σ antiaromaticity.
To describe the electron localization or delocalization of electrons, the electron localization function (ELF) [54] of the valence electrons was analyzed, as shown in Figures S29–S31. At the isosurface value of 0.60, the isosurface maps of most of the clusters are connected on the surface of the whole molecule. Yet, the isosurface diagrams of SeB5, SeB6, SeB9, and SeB11 are disconnected on the surface of the whole molecule, indicating that the delocalization of the whole molecule is weaker than that of the other clusters. Figure S30 displays the ELF with the isosurface value of 0.70. The isosurface diagram of SeB3, SeB3, SeB4, SeB9, SeB12, and SeB15 is still connected on the surface of the whole molecule, while the isosurface diagrams of other clusters are broken on the partial regions of the molecule, indicating that the delocalization of SeB3, SeB3, SeB4, SeB9, SeB12, and SeB15 is stronger than that of the other clusters. Figure S31 displays the ELF with the isosurface value of 0.80, in which the isosurface maps of some clusters are disconnected and there are no connected regions. The isosurface maps of SeB3, SeB3, SeB6, SeB7, SeB7, SeB9, SeB12, and SeB15 show that there is still some connected area on the surface of molecule, indicating that the local delocalization of these clusters is stronger than that of other clusters. Quite specially, the isosurface diagram of SeB9 is still connected on the peripheral B7 ring at the top of the mushroom. ELF analyses further confirm these observations based on the AdNDP analyses, such as the contributions from the valence electrons of the SeB12 were partitioned in Figure S31. Isosurface maps of the SeB12 (Figure S31) cover eight peripheral B-B bonds and two B-Se bonds that correspond to ten peripheral 2c–2e σ bonds, and they cover seven B3 triangles that correspond to seven 3c–2e σ bonds. Isosurface maps on the two B5 ring are fatter due to another two 5c–2e π bonds.
Figure S32 shows the isosurface diagram of the spin density of the open-shell clusters, and spin density can reveal the distribution of unpaired electrons. Figure S32 shows the spin density diagram with an isosurface value of 0.002, in which green represents alpha electrons and blue represents beta electrons. Figure S32 shows that the unpaired single electrons are mostly alpha electrons, and there are a small number of beta electrons on B atoms. Most of the unpaired alpha electrons are distributed on the B atoms; only a small portion of the unpaired alpha electrons are distributed on the Se atom. The spin density can reflect chemical reactions or adsorption to a certain extent. The single electrons of these clusters are mostly alpha electrons and are basically on the B atoms. The B or Se atoms with single alpha electrons can pair with free radicals or small molecules with beta single electrons to form new covalent bonds. In addition, these spin features are expected to produce interesting magnetic properties, which will further lead to potential applications in molecular devices.

2.2. Photoelectron Spectra

Photoelectron spectroscopy in combination with theoretical calculations was used to identify the structures of size-selected boron clusters [3,12,55]. To assist with future identifications of SeBn (n = 3–16), vertical detachment energies (VDEs) were calculated and photoelectron spectra of SeBn (n = 3–16) were simulated with the time-dependent DFT (TD-DFT) method [12,55,56].
Figure 6 presents the photoelectron spectra of SeBn (n = 3–16). The results indicate that SeB3 has the lowest first VDE, and SeB12 has the largest energy gap (about 1.48 eV) between the first and second peaks. The first several peaks were used to identify boron clusters [3,12]; the peaks on the low binding energy side are of great significance. The first peaks of these photoelectron spectra (except for SeB9) come from the calculated ground-state VDEs of SeB3, SeB4, SeB5, SeB6, SeB7, SeB8, SeB10, SeB11, SeB12, SeB13, SeB14, SeB15, and SeB16 at 2.52, 2.62, 2.89, 2.89, 2.89, 2.76, 3.14, 3.51, 2.66, 3.43, 3.15, 3.59, and 3.36 eV, respectively. The calculated ground-state VDEs of these closed-shell clusters originate from the detachment of the electron from the molecular orbital HOMO. However, for open-shell clusters, the calculated ground-state VDE of each cluster originates from the detachment of the electron from the singly occupied molecular orbital α-SOMO. The first peak of SeB9 comes from the second VDE at 4.01 eV, which is smaller than the ground-state VDE of 4.25 eV (second peak). The second peaks of SeB3, SeB5, SeB7, SeB11, SeB13, and SeB15 come from the second calculated VDEs at 3.23, 4.07, 4.16, 3.81, 4.30, and 4.12 eV, respectively, which originate from detaching the electrons from HOMO-1. The second peaks of SeB4, SeB6, SeB8, SeB10, SeB12, SeB14, and SeB16 come from the second VDEs at 3.36, 4.02, 3.07, 3.48, 4.14, 3.64, and 3.63 eV, respectively, which originate from detaching the electrons from the singly occupied molecular orbital β -HOMO-1. In addition, the peaks with higher binding energy originate from detaching the electrons from lower molecular orbitals. It is noted that some of the doped anionic boron clusters have a similar skeleton as the corresponding anionic pure boron clusters. Comparing their photoelectron spectra, the addition of the Se atom results in a great change in the photoelectron spectra [52,53]. However, the photoelectron spectra of some doped boron clusters are similar to those of the corresponding anionic pure boron clusters [52,53]. For example, compared with the photoelectron spectra of pure boron clusters, the addition of Se atoms causes the first two peaks of SeB3- to move 0.30 eV towards the low binding energy side and causes the first two peaks of B5- to move 0.46 eV towards the high binding energy side [52]. For SeB4, SeB8, and SeB13-, compared with the photoelectron spectra of pure boron clusters, the addition of Se atoms causes different band characteristics [52]. Compared to the photoelectron spectra of pure boron clusters, planar SeB7 and quasi-planar B7- have almost the same first VDE (2.89 eV for SeB7-, 2.85 ± 0.02 eV for B7) [52], planar SeB10- and quasi-planar B10- have almost the same first VDE (3.14 eV for SeB10, 3.06 ± 0.03 eV for B10) [52], and quasi-planar SeB14 and quasi-planar B14- have almost the same first VDE (3.14 eV for SeB14, 3.10 ± 0.01 eV for B14) [52]. For SeB11-, compared to the photoelectron spectra of pure boron clusters B11, the addition of the Se atom causes the first peak to move 0.08 eV towards the high binding energy side and causes the second peak to move 0.25 eV towards the low binding energy side. For SeB12 and SeB16, planar SeB12- and quasi-planar B12 have similar band characteristics [52], and quasi-planar SeB16 and quasi-planar B16- have similar band characteristics [53]. Figure 6 indicates that SeBn (n = 3–16) has different spectral features; especially the peaks at the low binding energy side can identify the SeBn (n = 3–16). As with the discovery of other anionic boron clusters, if the photoelectron spectra of SeBn (n = 3–16) are obtained in experiments, these simulated values can be used for the identification of SeBn (n = 3–16).

3. Computation Details

Configuration searches of Se-doped boron clusters SeBn0/− (n = 3–16) were performed with CALYPSO 5.0 software in combination with Gaussian 16 software. CALYPSO is a reliable cluster configuration prediction software, and it has successfully predicted boron or doped boron clusters [11,25,40,41,45,57,58,59,60,61]. The initial structures were generated by the CALYPSO software, and then these initial structures were optimized using Gaussian 16 software at the PBE0/3-21G level for the preliminary structural search. In each generation produced by the CALYPSO software, 70% of the structures were produced by particle swarm optimization (PSO) operations, while the others were randomly generated. When cluster sizes vary from n = 3 to n = 10, nearly 100–900 isomers are initially predicted for each boron cluster of a different size. When cluster sizes vary from n = 11 to n = 16, nearly 2000 isomers are initially predicted for each boron cluster of a different size.
After the preliminary structural search, low-energy structures were then fully optimized at the PBE0/6-311+G(d) level [51,62]. After the optimizations, frequency analyses and electronic structures were studied at the PBE0/6-311+G(d) level. PBE0/6-311+G(d) is a reliable level for boron cluster [12,60,61,63,64,65]; in particular, theoretical simulated values with PBE0/6-311+G(d) are the same as the experimental values [12]. Therefore, all calculations in this article used the method PBE0/6-311+G(d) and were performed using Gaussian 16 software [66]. The analyses and isosurface map drawings were performed using Multiwfn 3.7 code [67] and the visual molecular dynamics (VMD) program [68].

4. Conclusions

DFT combined with CALYPSO software is employed to demonstrate the structural evolution of SeBn0/− (n = 3–16) clusters. The conclusions are summarized as follows. (1) The global minima of SeBn0/− (n = 3–16) clusters tend to form quasi-planar or planar structures. (2) The ground-state structure of SeB9 is a three-dimensional, mushroom-shaped, ground-state structure, and it possesses the largest HOMO–LUMO gap of 5.08 eV, which shows the superior chemical stability. (3) AdNDP bonding analyses reveal that SeB8 is doubly aromatic, with six delocalized σ and six delocalized π electrons, whereas SeB9 is doubly antiaromatic, with twelve delocalized σ and four delocalized π electrons. Similarly, SeB12 is doubly aromatic, with fourteen delocalized σ and six delocalized π electrons. (4) ELF analysis shows that SeBn0/− (n = 3–16) clusters have different local electron delocalization and whole-electron delocalization effects. (5) SeBn (n = 3–16) have different photoelectron spectra, and especially the first several peaks can be used for the identification of SeBn (n = 3–16). This research has enriched the structures of doped boron clusters.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28010357/s1, Figures S1–28: Low-lying isomers of doped boron clusters SeBn0/− (n = 3–16); Figures S29–31: Electron localization function (ELF) of SeBn0/− (n = 3–16); Figure S32: Spin density of open-shell doped boron clusters.

Author Contributions

Conceptualization, S.-X.L. and Z.-W.L.; methodology, Y.-J.Y. and S.-X.L.; software, S.-X.L.; formal analysis, Z.-W.L.; investigation, S.-X.L.; data processing, Y.-J.Y.; writing—original draft preparation, Y.-J.Y.; writing—review and editing, S.-X.L. and Y.-J.Y.; supervision, D.-L.C.; funding acquisition, D.-L.C. and Y.-J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Central Guiding Local Science and Technology Development Foundation of China (Grant No. QK ZYD [2019]4012) and the Growth Foundation for Young Scientists of the Education Department of Guizhou Province (Grant No: QJH KY [2022]310), China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and Supplementary Materials.

Conflicts of Interest

There are no conflicts of interest to declare.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Boustani, I. Systematic ab initio investigation of bare boron clusters: Determination of the geometryand electronic structures of Bn (n = 2 − 14). Phys. Rev. B 1997, 55, 16426–16438. [Google Scholar] [CrossRef]
  2. Zhai, H.-J.; Wang, L.-S.; Alexandrova, A.N.; Boldyrev, A.I. Electronic structure and chemical bonding of B5 and B5 by photoelectron spectroscopy andab initiocalculations. J. Chem. Phys. 2002, 117, 7917–7924. [Google Scholar] [CrossRef]
  3. Zhai, H.J.; Kiran, B.; Li, J.; Wang, L.S. Hydrocarbon analogues of boron clusters--planarity, aromaticity and antiaromaticity. Nat. Mater. 2003, 2, 827–833. [Google Scholar] [CrossRef]
  4. Kiran, B.; Bulusu, S.; Zhai, H.J.; Yoo, S.; Zeng, X.C.; Wang, L.S. Planar-to-tubular structural transition in boron clusters: B20 as the embryo of single-walled boron nanotubes. Proc. Natl. Acad. Sci. USA 2005, 102, 961–964. [Google Scholar] [CrossRef] [Green Version]
  5. Bean, D.E.; Fowler, P.W. Double Aromaticity in “Boron Toroids”. J. Phys. Chem. C 2009, 113, 15569–15575. [Google Scholar] [CrossRef]
  6. Chen, Q.; Wei, G.F.; Tian, W.J.; Bai, H.; Liu, Z.P.; Zhai, H.J.; Li, S.D. Quasi-planar aromatic B36 and B36- clusters: All-boron analogues of coronene. Phys. Chem. Chem. Phys. 2014, 16, 18282–18287. [Google Scholar] [CrossRef] [PubMed]
  7. Sergeeva, A.P.; Popov, I.A.; Piazza, Z.A.; Li, W.L.; Romanescu, C.; Wang, L.S.; Boldyrev, A.I. Understanding boron through size-selected clusters: Structure, chemical bonding, and fluxionality. Acc. Chem. Res. 2014, 47, 1349–1358. [Google Scholar] [CrossRef]
  8. Jian, T.; Chen, X.; Li, S.D.; Boldyrev, A.I.; Li, J.; Wang, L.S. Probing the structures and bonding of size-selected boron and doped-boron clusters. Chem. Soc. Rev. 2019, 48, 3550–3591. [Google Scholar] [CrossRef] [PubMed]
  9. Casillas, R.; Baruah, T.; Zope, R.R. Geometry and electronic structure of neutral and charged B21 clusters. Chem. Phys. Lett. 2013, 557, 15–18. [Google Scholar] [CrossRef]
  10. Pham, H.T.; Duong, L.V.; Pham, B.Q.; Nguyen, M.T. The 2D-to-3D geometry hopping in small boron clusters: The charge effect. Chem. Phys. Lett. 2013, 577, 32–37. [Google Scholar] [CrossRef]
  11. Lv, J.; Wang, Y.; Zhu, L.; Ma, Y. B38: An all-boron fullerene analogue. Nanoscale 2014, 6, 11692–11696. [Google Scholar] [CrossRef] [PubMed]
  12. Zhai, H.J.; Zhao, Y.F.; Li, W.L.; Chen, Q.; Bai, H.; Hu, H.S.; Piazza, Z.A.; Tian, W.J.; Lu, H.G.; Wu, Y.B.; et al. Observation of an all-boron fullerene. Nat. Chem. 2014, 6, 727–731. [Google Scholar] [CrossRef]
  13. Bai, H.; Chen, Q.; Zhai, H.J.; Li, S.D. Endohedral and exohedral metalloborospherenes: M@B40 (M=Ca, Sr) and M&B40 (M=Be, Mg). Angew. Chem. Int. Ed. 2015, 54, 941–945. [Google Scholar]
  14. Li, S.-X.; Zhang, Z.-P.; Long, Z.-W.; Qin, S.-J. Structures, stabilities and spectral properties of metalloborospherenes MB400/− (M = Cu, Ag, and Au). RSC Adv. 2017, 7, 38526–38537. [Google Scholar] [CrossRef] [Green Version]
  15. Dong, H.; Hou, T.; Lee, S.T.; Li, Y. New Ti-decorated B40 fullerene as a promising hydrogen storage material. Sci. Rep. 2015, 5, 9952. [Google Scholar] [CrossRef] [Green Version]
  16. An, Y.; Zhang, M.; Wu, D.; Fu, Z.; Wang, T.; Xia, C. Electronic transport properties of the first all-boron fullerene B40 and its metallofullerene Sr@B40. Phys. Chem. Chem. Phys. 2016, 18, 12024–12028. [Google Scholar] [CrossRef]
  17. Bai, H.; Bai, B.; Zhang, L.; Huang, W.; Mu, Y.W.; Zhai, H.J.; Li, S.D. Lithium-Decorated Borospherene B40: A Promising Hydrogen Storage Medium. Sci. Rep. 2016, 6, 35518. [Google Scholar] [CrossRef]
  18. Shakerzadeh, E.; Biglari, Z.; Tahmasebi, E. M@B40 (M = Li, Na, K) serving as a potential promising novel NLO nanomaterial. Chem. Phys. Lett. 2016, 654, 76–80. [Google Scholar] [CrossRef]
  19. Tang, C.; Zhang, X. The hydrogen storage capacity of Sc atoms decorated porous boron fullerene B40: A DFT study. Int. J. Hydrogen Energ. 2016, 41, 16992–16999. [Google Scholar] [CrossRef]
  20. Li, S.; Zhang, Z.; Long, Z.; Chen, D. Structures, Stabilities, and Spectral Properties of Endohedral Borospherenes M@B40 0/- (M = H2, HF, and H2O). ACS Omega 2019, 4, 5705–5713. [Google Scholar] [CrossRef] [Green Version]
  21. Li, S.X.; Zhang, Z.P.; Long, Z.W.; Chen, D.L. Structures, Electronic, and Spectral Properties of Doped Boron Clusters MB120/− (M = Li, Na, and K). ACS Omega 2020, 5, 20525–20534. [Google Scholar] [CrossRef]
  22. Mannix, A.J.; Zhou, X.F.; Kiraly, B.; Wood, J.D.; Alducin, D.; Myers, B.D.; Liu, X.; Fisher, B.L.; Santiago, U.; Guest, J.R.; et al. Synthesis of borophenes: Anisotropic, two-dimensional boron polymorphs. Science 2015, 350, 1513–1516. [Google Scholar] [CrossRef] [Green Version]
  23. Li, Q.; Kolluru, V.S.C.; Rahn, M.S.; Schwenker, E.; Li, S.; Hennig, R.G.; Darancet, P.; Chan, M.K.Y.; Hersam, M.C. Synthesis of borophane polymorphs through hydrogenation of borophene. Science 2021, 371, 1143–1148. [Google Scholar] [CrossRef] [PubMed]
  24. Popov, I.A.; Li, W.L.; Piazza, Z.A.; Boldyrev, A.I.; Wang, L.S. Complexes between planar boron clusters and transition metals: A photoelectron spectroscopy and ab initio study of CoB12 and RhB12. J. Phys. Chem. A 2014, 118, 8098–8105. [Google Scholar] [CrossRef] [PubMed]
  25. Liang, W.Y.; Das, A.; Dong, X.; Cui, Z.H. Lithium doped tubular structure in LiB20 and LiB20-: A viable global minimum. Phys. Chem. Chem. Phys. 2018, 20, 16202–16208. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, W.; Guo, Y.-D.; Yan, X.-H. The spin-dependent transport of transition metal encapsulated B40 fullerene. RSC Adv. 2016, 6, 40155–40161. [Google Scholar] [CrossRef]
  27. Saha, R.; Kar, S.; Pan, S.; Martinez-Guajardo, G.; Merino, G.; Chattaraj, P.K. A Spinning Umbrella: Carbon Monoxide and Dinitrogen Bound MB12- Clusters (M = Co, Rh, Ir). J. Phys. Chem. A 2017, 121, 2971–2979. [Google Scholar] [CrossRef]
  28. Li, S.-X.; Chen, D.-L.; Zhang, Z.-P.; Long, Z.-W. Ground state structures and properties of Be atom doped boron clusters BeBn (n = 10 − 15). Acta. Physica. Sinica. 2020, 69, 193101. [Google Scholar] [CrossRef]
  29. Cheung, L.F.; Kocheril, G.S.; Czekner, J.; Wang, L.S. Observation of Mobius Aromatic Planar Metallaborocycles. J. Am. Chem. Soc. 2020, 142, 3356–3360. [Google Scholar] [CrossRef]
  30. Popov, I.A.; Jian, T.; Lopez, G.V.; Boldyrev, A.I.; Wang, L.S. Cobalt-centred boron molecular drums with the highest coordination number in the CoB16 cluster. Nat. Commun. 2015, 6, 8654. [Google Scholar] [CrossRef] [Green Version]
  31. Jian, T.; Li, W.L.; Popov, I.A.; Lopez, G.V.; Chen, X.; Boldyrev, A.I.; Li, J.; Wang, L.S. Manganese-centered tubular boron cluster—MnB16 (-): A new class of transition-metal molecules. J. Chem. Phys. 2016, 144, 154310. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, T.T.; Li, W.L.; Bai, H.; Chen, W.J.; Dong, X.R.; Li, J.; Wang, L.S. ReB8- and ReB9-: New Members of the Transition-Metal-Centered Borometallic Molecular Wheel Family. J. Phys. Chem. A 2019, 123, 5317–5324. [Google Scholar] [CrossRef] [PubMed]
  33. Cheung, L.F.; Czekner, J.; Kocheril, G.S.; Wang, L.S. ReB6-: A Metallaboron Analog of Metallabenzenes. J. Am. Chem. Soc. 2019, 141, 17854–17860. [Google Scholar] [CrossRef]
  34. Chen, T.T.; Li, W.L.; Chen, W.J.; Yu, X.H.; Dong, X.R.; Li, J.; Wang, L.S. Spherical trihedral metallo-borospherenes. Nat. Commun. 2020, 11, 2766. [Google Scholar] [CrossRef] [PubMed]
  35. Cheung, L.F.; Kocheril, G.S.; Czekner, J.; Wang, L.S. MnB6-: An Open-Shell Metallaboron Analog of 3d Metallabenzenes. J. Phys. Chem. A 2020, 124, 2820–2825. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, W.J.; Kulichenko, M.; Choi, H.W.; Cavanagh, J.; Yuan, D.F.; Boldyrev, A.I.; Wang, L.S. Photoelectron Spectroscopy of Size-Selected Bismuth-Boron Clusters: BiBn (n = 6 − 8). J. Phys. Chem. A 2021, 125, 6751–6760. [Google Scholar] [CrossRef]
  37. Jiang, Z.Y.; Chen, T.T.; Chen, W.J.; Li, W.L.; Li, J.; Wang, L.S. Expanded Inverse-Sandwich Complexes of Lanthanum Borides: La2B10 and La2B11. J. Phys. Chem. A 2021, 125, 2622–2630. [Google Scholar] [CrossRef]
  38. Barroso, J.; Pan, S.; Merino, G. Structural transformations in boron clusters induced by metal doping. Chem. Soc. Rev. 2022, 51, 1098–1123. [Google Scholar] [CrossRef] [PubMed]
  39. Chacko, S.; Kanhere, D.G.; Boustani, I. Ab initiodensity functional investigation of B24 clusters: Rings, tubes, planes, and cages. Phys. Rev. B 2003, 68, 035414. [Google Scholar] [CrossRef] [Green Version]
  40. Liang, W.; Das, A.; Dong, X.; Wang, M.; Cui, Z. Structural and electronic properties of MB22 (M = Na, K) clusters: Tubular boron versus quasi-planar boron forms. New J. Chem. 2019, 43, 6507–6512. [Google Scholar] [CrossRef]
  41. Lv, J.; Wang, Y.; Zhang, L.; Lin, H.; Zhao, J.; Ma, Y. Stabilization of fullerene-like boron cages by transition metal encapsulation. Nanoscale 2015, 7, 10482–10489. [Google Scholar] [CrossRef] [PubMed]
  42. Yang, Y.J.; Li, S.X.; Chen, D.L.; Long, Z.W. Structural and Electronic Properties of Single-Atom Transition Metal-Doped Boron Clusters MB24 (M = Sc, V, and Mn). ACS Omega 2021, 6, 30442–30450. [Google Scholar] [CrossRef]
  43. Li, S.X.; Yang, Y.J.; Chen, D.L.; Long, Z.W. Structures, and electronic and spectral properties of single-atom transition metal-doped boron clusters MB24- (M = Sc, Ti, V, Cr, Mn, Fe, Co, and Ni). RSC Adv. 2022, 12, 16706–16716. [Google Scholar] [CrossRef]
  44. Dordevic, S.; Radenkovic, S. Electronic structure, stability, and aromaticity of M2B6 (M = Mg, Ca, Sr, and Ba): An interplay between spin pairing and electron delocalization. Phys. Chem. Chem. Phys. 2022, 24, 5833–5841. [Google Scholar] [CrossRef] [PubMed]
  45. Dordevic, S.; Radenkovic, S. Spatial and Electronic Structures of BeB8 and MgB8: How far Does the Analogy Go? Chemphyschem 2022, 23, e202200070. [Google Scholar] [CrossRef]
  46. Alivisatos, A.P. Semiconductor Clusters, Nanocrystals, and Quantum Dots. Science 1996, 271, 933–937. [Google Scholar] [CrossRef] [Green Version]
  47. Kasuya, A.; Sivamohan, R.; Barnakov, Y.A.; Dmitruk, I.M.; Nirasawa, T.; Romanyuk, V.R.; Kumar, V.; Mamykin, S.V.; Tohji, K.; Jeyadevan, B.; et al. Ultra-stable nanoparticles of CdSe revealed from mass spectrometry. Nat. Mater. 2004, 3, 99–102. [Google Scholar] [CrossRef]
  48. Michalet, X.; Pinaud, F.F.; Bentolila, L.A.; Tsay, J.M.; Doose, S.; Li, J.J.; Sundaresan, G.; Wu, A.M.; Gambhir, S.S.; Weiss, S. Quantum dots for live cells, in vivo imaging, and diagnostics. Science 2005, 307, 538–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Medintz, I.L.; Uyeda, H.T.; Goldman, E.R.; Mattoussi, H. Quantum dot bioconjugates for imaging, labelling and sensing. Nat. Mater. 2005, 4, 435–446. [Google Scholar] [CrossRef]
  50. Lv, J.; Wang, Y.; Zhu, L.; Ma, Y. Particle-swarm structure prediction on clusters. J. Chem. Phys. 2012, 137, 084104. [Google Scholar] [CrossRef]
  51. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  52. Alexandrova, A.N.; Boldyrev, A.I.; Zhai, H.-J.; Wang, L.-S. All-boron aromatic clusters as potential new inorganic ligands and building blocks in chemistry. Coord. Chem. Rev. 2006, 250, 2811–2866. [Google Scholar] [CrossRef]
  53. Sergeeva, A.P.; Zubarev, D.Y.; Zhai, H.-J.; Boldyrev, A.I.; Wang, L.-S. A Photoelectron Spectroscopic and Theoretical Study of B16 and B162−: An All-Boron Naphthalene. J. Am. Chem. Soc. 2008, 130, 7244–7246. [Google Scholar] [CrossRef] [PubMed]
  54. Becke, A.D.; Edgecombe, K.E. A simple measure of electron localization in atomic and molecular systems. J. Chem. Phys. 1990, 92, 5397–5403. [Google Scholar] [CrossRef]
  55. Chen, Q.; Li, W.L.; Zhao, Y.F.; Zhang, S.Y.; Hu, H.S.; Bai, H.; Li, H.R.; Tian, W.J.; Lu, H.G.; Zhai, H.J.; et al. Experimental and theoretical evidence of an axially chiral borospherene. ACS Nano 2015, 9, 754–760. [Google Scholar] [CrossRef]
  56. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
  57. Dong, X.; Jalife, S.; Vasquez-Espinal, A.; Barroso, J.; Orozco-Ic, M.; Ravell, E.; Cabellos, J.L.; Liang, W.Y.; Cui, Z.H.; Merino, G. Li2B24: The simplest combination for a three-ring boron tube. Nanoscale 2019, 11, 2143–2147. [Google Scholar] [CrossRef]
  58. Dong, X.; Liu, Y.Q.; Liu, X.B.; Pan, S.; Cui, Z.H.; Merino, G. Be4B12+: A Covalently Bonded Archimedean Beryllo-Borospherene. Angew. Chem. Int. Ed. 2022, 61, e202208152. [Google Scholar] [CrossRef]
  59. Wei, D.; Ren, M.; Lu, C.; Bi, J.; Maroulis, G. A quasi-plane IrB18- cluster with high stability. Phys. Chem. Chem. Phys. 2020, 22, 5942–5948. [Google Scholar] [CrossRef]
  60. Ren, M.; Jin, S.; Wei, D.; Jin, Y.; Tian, Y.; Lu, C.; Gutsev, G.L. NbB12-: A new member of half-sandwich type doped boron clusters with high stability. Phys. Chem. Chem. Phys. 2019, 21, 21746–21752. [Google Scholar] [CrossRef]
  61. Tian, Y.; Wei, D.; Jin, Y.; Barroso, J.; Lu, C.; Merino, G. Exhaustive exploration of MgBn (n = 10–20) clusters and their anions. Phys. Chem. Chem. Phys. 2019, 21, 6935–6941. [Google Scholar] [CrossRef] [PubMed]
  62. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  63. Le Chen, B.; Sun, W.G.; Kuang, X.Y.; Lu, C.; Xia, X.X.; Shi, H.X.; Maroulis, G. Structural Stability and Evolution of Medium-Sized Tantalum-Doped Boron Clusters: A Half-Sandwich-Structured TaB12- Cluster. Inorg. Chem. 2018, 57, 343–350. [Google Scholar] [CrossRef]
  64. Li, P.; Du, X.; Wang, J.J.; Lu, C.; Chen, H. Probing the Structural Evolution and Stabilities of Medium-Sized MoBn0/– Clusters. J. Phys. Chem. C 2018, 122, 20000–20005. [Google Scholar] [CrossRef]
  65. Jin, S.; Chen, B.; Kuang, X.; Lu, C.; Sun, W.; Xia, X.; Gutsev, G.L. Structural and Electronic Properties of Medium-Sized Aluminum-Doped Boron Clusters AlBn and Their Anions. J. Phys. Chem. C 2019, 123, 6276–6283. [Google Scholar] [CrossRef]
  66. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  67. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  68. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of SeBn. (a) SeB3 C2V; (b) SeB4 Cs; (c) SeB5 Cs; (d) SeB6 Cs; (e) SeB7 Cs; (f) SeB8 C2V; (g) SeB9 Cs; (h) SeB10; (i) SeB11 Cs; (j) SeB12 Cs; (k) SeB13 C1; (l) SeB14 Cs; (m) SeB15 I C1; (n) SeB15 II C1; (o) SeB16 C1.
Figure 1. Structures of SeBn. (a) SeB3 C2V; (b) SeB4 Cs; (c) SeB5 Cs; (d) SeB6 Cs; (e) SeB7 Cs; (f) SeB8 C2V; (g) SeB9 Cs; (h) SeB10; (i) SeB11 Cs; (j) SeB12 Cs; (k) SeB13 C1; (l) SeB14 Cs; (m) SeB15 I C1; (n) SeB15 II C1; (o) SeB16 C1.
Molecules 28 00357 g001
Figure 2. Structures of SeBn. (a) SeB3 C2V; (b) SeB4 Cs; (c) SeB5 Cs; (d) SeB6 C1; (e) SeB7 C2V; (f) SeB8 C2V; (g) SeB9 C6V; (h) SeB10 C1; (i) SeB11 Cs; (j) SeB12 C2V; (k) SeB13 C1; (l) SeB14 Cs; (m) SeB15 I C1; (n) SeB15 II C1; (o) SeB16 C1.
Figure 2. Structures of SeBn. (a) SeB3 C2V; (b) SeB4 Cs; (c) SeB5 Cs; (d) SeB6 C1; (e) SeB7 C2V; (f) SeB8 C2V; (g) SeB9 C6V; (h) SeB10 C1; (i) SeB11 Cs; (j) SeB12 C2V; (k) SeB13 C1; (l) SeB14 Cs; (m) SeB15 I C1; (n) SeB15 II C1; (o) SeB16 C1.
Molecules 28 00357 g002
Figure 3. Bonding patterns of SeB8. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9599 |e|; (b) 8 × 2c − 2e σ bonds, ON = 1.9277 − 1.9475 |e|; (c) 1 × 3c − 2e π bond, ON = 1.9653 |e|; (d) 1 × 3c − 2e σ bond, ON = 1.7792 |e|; (e) 2 × 4c − 2e σ bonds, ON = 1.8980 |e|; (f) 1 × 5c − 2e π bond, ON = 1.9445 |e|; (g) 1 × 5c − 2e π bond, ON = 1.9445 |e|.
Figure 3. Bonding patterns of SeB8. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9599 |e|; (b) 8 × 2c − 2e σ bonds, ON = 1.9277 − 1.9475 |e|; (c) 1 × 3c − 2e π bond, ON = 1.9653 |e|; (d) 1 × 3c − 2e σ bond, ON = 1.7792 |e|; (e) 2 × 4c − 2e σ bonds, ON = 1.8980 |e|; (f) 1 × 5c − 2e π bond, ON = 1.9445 |e|; (g) 1 × 5c − 2e π bond, ON = 1.9445 |e|.
Molecules 28 00357 g003
Figure 4. Bonding patterns of SeB9. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9644 |e|; (b) 1 × 2c − 2e σ bond, ON = 1.9965 |e|; (c) 1 × 2c − 2e σ bond, ON = 1.9523 |e|; (d) 1 × 2c − 2e π bond, ON = 1.9884 |e|; (e) 1 × 2c − 2e π bond, ON = 1.9884 |e|; (f) 6 × 2c − 2e σ bonds, ON = 1.9247 |e|; (g) 6 × 3c − 2e σ bonds, ON = 1.7547 |e|.
Figure 4. Bonding patterns of SeB9. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9644 |e|; (b) 1 × 2c − 2e σ bond, ON = 1.9965 |e|; (c) 1 × 2c − 2e σ bond, ON = 1.9523 |e|; (d) 1 × 2c − 2e π bond, ON = 1.9884 |e|; (e) 1 × 2c − 2e π bond, ON = 1.9884 |e|; (f) 6 × 2c − 2e σ bonds, ON = 1.9247 |e|; (g) 6 × 3c − 2e σ bonds, ON = 1.7547 |e|.
Molecules 28 00357 g004
Figure 5. Bonding patterns of SeB12. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9731 |e|; (b) 10 × 2c − 2e σ bonds, ON = 1.8571 − 1.9418 |e|; (c) 1 × 3c − 2e π bond, ON = 1.9823 |e|; (d) 7 × 3c − 2e σ bonds, ON = 1.6403 − 1.9246 |e|; (e) 2 × 5c − 2e π bonds, ON = 1.8623 |e|.
Figure 5. Bonding patterns of SeB12. ON is occupation number and the orange ball is Se atom. (a) 1 × 1c − 2e, ON = 1.9731 |e|; (b) 10 × 2c − 2e σ bonds, ON = 1.8571 − 1.9418 |e|; (c) 1 × 3c − 2e π bond, ON = 1.9823 |e|; (d) 7 × 3c − 2e σ bonds, ON = 1.6403 − 1.9246 |e|; (e) 2 × 5c − 2e π bonds, ON = 1.8623 |e|.
Molecules 28 00357 g005
Figure 6. Photoelectron spectra at the PBE0/6-311 + G * level. (a) SeB3; (b) SeB4; (c) SeB5; (d) SeB6; (e) SeB7; (f) SeB8; (g) SeB9; (h) SeB10; (i) SeB11; (j) SeB12; (k) SeB13; (l) SeB14; (m) SeB15; (n) SeB16.
Figure 6. Photoelectron spectra at the PBE0/6-311 + G * level. (a) SeB3; (b) SeB4; (c) SeB5; (d) SeB6; (e) SeB7; (f) SeB8; (g) SeB9; (h) SeB10; (i) SeB11; (j) SeB12; (k) SeB13; (l) SeB14; (m) SeB15; (n) SeB16.
Molecules 28 00357 g006
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

Yang, Y.-J.; Li, S.-X.; Chen, D.-L.; Long, Z.-W. Structural Evolution and Electronic Properties of Selenium-Doped Boron Clusters SeBn0/− (n = 3–16). Molecules 2023, 28, 357. https://doi.org/10.3390/molecules28010357

AMA Style

Yang Y-J, Li S-X, Chen D-L, Long Z-W. Structural Evolution and Electronic Properties of Selenium-Doped Boron Clusters SeBn0/− (n = 3–16). Molecules. 2023; 28(1):357. https://doi.org/10.3390/molecules28010357

Chicago/Turabian Style

Yang, Yue-Ju, Shi-Xiong Li, De-Liang Chen, and Zheng-Wen Long. 2023. "Structural Evolution and Electronic Properties of Selenium-Doped Boron Clusters SeBn0/− (n = 3–16)" Molecules 28, no. 1: 357. https://doi.org/10.3390/molecules28010357

Article Metrics

Back to TopTop