Next Article in Journal
The TRPA1 Ion Channel Mediates Oxidative Stress-Related Migraine Pathogenesis
Next Article in Special Issue
A Nickel/Organoboron-Catalyzed Coupling of Aryl Bromides with Sodium Sulfinates: The Synthesis of Sulfones under Visible Light
Previous Article in Journal
Evaluation of Anticancer Activity of Nucleoside–Nitric Oxide Photo-Donor Hybrids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Evolution of Small-Sized Phosphorus-Doped Boron Clusters: A Half-Sandwich-Structured PB15 Cluster

School of Physics and Electronic Science, Guizhou Education University, Guiyang 550018, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(14), 3384; https://doi.org/10.3390/molecules29143384
Submission received: 30 May 2024 / Revised: 12 July 2024 / Accepted: 16 July 2024 / Published: 18 July 2024
(This article belongs to the Special Issue Boron Chemistry and Applications)

Abstract

:
The present study is a theoretical investigation into the structural evolution, electronic properties, and photoelectron spectra of phosphorus-doped boron clusters PBn0/− (n = 3–17). The results of this study revealed that the lowest energy structures of PBn (n = 3–17) clusters, except for PB17, exhibit planar or quasi-planar structures. The lowest energy structures of PBn (n = 3–17), with the exceptions of PB7, PB9, and PB15, are planar or quasi-planar. The ground state of PB7 has an umbrella-shaped structure, with C6V symmetry. Interestingly, the neutral cluster PB15 has a half-sandwich-like structure, in which the P atom is attached to three B atoms at one end of the sandwich, exhibiting excellent relative and chemical stability due to its higher second-order energy difference and larger HOMO–LUMO energy gap of 4.31 eV. Subsequently, adaptive natural density partitioning (AdNDP) and electron localization function (ELF) analyses demonstrate the bonding characteristics of PB7 and PB15, providing support for the validity of their stability. The calculated photoelectron spectra show distinct characteristic peaks of PBn (n = 3–17) clusters, thus providing theoretical evidence for the future identification of doped boron clusters. In summary, our work has significant implications for understanding the structural evolution of doped boron clusters PBn0/− (n = 3–17), motivating further experiments regarding doped boron clusters.

1. Introduction

It is well known that boron is a naturally abundant element, and scientists have been exploring the properties and applications of boron, boron compounds, and their derivatives [1,2,3,4,5]. With the successive discoveries of fullerene C60, carbon nanotubes, and graphene [6,7,8], carbon nanomaterials have attracted considerable attention and become a dynamic field of research; similarly, as a neighbor of carbon, boron has also received significant attention from researchers. Boron compounds exhibit intrinsic electronic defects, as well as the ability to form an outstanding number of multicenter bonds, which endow them with rich structural features and chemical properties [9,10,11,12,13,14]. The groundbreaking discovery of borospherene [15] in 2014 sparked extensive research on boron clusters [16,17,18,19,20,21,22]. Over the past two decades, a great deal of theoretical and experimental research has focused on the geometric structure and electronic characteristics of bare boron clusters [23]. It is generally believed that bare boron clusters exhibit planar or quasi-planar geometric shapes over a large range of sizes. One of the most important discoveries is the planar B36 cluster [24], a highly stable cluster with central hexagonal vacancies that can be synthesized on a metal substrate to form “borophene”. The experimental discoveries of borospherene and borophene provide new insights into novel boron nanomaterials and nanodevices.
Over the past decade, scientists have investigated richly doped boron clusters, with a primary focus on doping single metal atoms into boron clusters of various sizes. Metal doping is an effective method for adjusting the chemical bonds and occupied energy levels of boron clusters through the addition of metal elements, thereby changing their physicochemical properties. Doping metal atoms into boron clusters can lead to the formation of new geometric configurations and chemical properties, such as ring-like, sandwich-like, tube-like, and cage-like structures [17,19,20,25,26,27,28,29,30,31,32,33,34]. Following the doping of single alkali metal atoms into the quasi-planar structures of B20 and B22 [35], species including LiB20, NaB22, and KB22 exhibit bi-ring structures [33,36]. Quasi-planar B12 clusters, upon doping with single transition metal atoms of Co, Rh, or Ta, exhibit half-sandwich structures [10,37,38]. Doping with single transition metal atoms can modify the bi-ring tubular B24 into cage-like boron clusters (TiB24, VB24, and MnB24) and tri-ring tubular doped boron clusters (ScB24) [39,40,41], or convert quasi-planar B24 into cage-like boron clusters (TiB24, CrB24, and VB24) [35,42].
Compared with metal-doped boron clusters, there is relatively little research on the doping of boron clusters with non-metal atoms [43,44,45,46,47]. In particular, there have been few studies on the structural evolution of boron clusters with the addition of phosphorus atoms. The recently reported P-doped boron cluster P2B12+/0/− has the same cage structure, with D3h symmetry [46]. The theoretical study of phosphorus-doped small boron clusters is of great significance for the discovery of new structures and properties of boron clusters. In this study, the effect of P-atom doping on the structure and electronic properties of boron clusters Bn0/− (n = 3–17) was investigated through employment of the particle swarm optimization (CALYPSO) method [48] and the density-functional theory method PBE0 [49].

2. Results and Discussion

2.1. Geometric Configurations

Early theoretical and experimental studies found that anionic boron clusters with an atomic numbers less than 37 always maintain quasi-planar or planar structures; however, while some neutral boron clusters do have quasi-planar or planar structures, others have tube or cage-like structures. The structures of boron clusters change after atom doping. In order to facilitate understanding via visualization, the low-lying isomers of PBn0/− (n = 3–17), along with their corresponding relative energy values, are displayed in Figures S1–S30, and the lowest energy structures of PBn0/− (n = 3–17) are depicted in Figure 1 and Figure 2. Clear structural diagrams show that the lowest-lying energy structures of PBn (n = 3–17) clusters have planar or quasi-planar structures, except for PB17. The lowest energy structures of PBn (n = 3–17) are planar or quasi-planar, with the exceptions of PB7, PB9, and PB15. As can be observed in Figure 1 and Figure 2, after the addition of the P atom, most boron clusters remain virtually unchanged from their corresponding bare boron cluster forms, such as PBn (n = 3–6, 8, 10, 12) and PBn (n = 3, 5–8, 10–16) [23,50,51]. In order to further discuss the structural changes of boron clusters, we also compared their structures after replacing a boron atom with a phosphorus atom, and without replacing the boron atom. Compared with the bare boron clusters Bn0/− (n = 4–18), the structures of most PBn−10/− (n = 4–18) clusters changed, except for those of PB3, PB3, PB4, and PB5 (which remained almost similar to those of B4, B4, B5, and B6, respectively); that is, the structures of the bare boron clusters changed due to the replacement of one atom with a P atom. The lowest energy configurations of PBn0/− (n = 3, 5, 6, 8) are planar structures, with the P atom attached to the boron atoms of the bare boron clusters Bn0/− (n = 3, 5, 6, 8) [50]. The lowest-lying energy structures of PBn0/− (n =10, 12) are quasi-planar, similar to the ground-state structures of their corresponding bare boron clusters, B100/− and B120/−, with the addition of a phosphorus atom bonded to the boron atoms. PB10 has a chiral symmetrical structure. Likewise, the planar clusters PB4, PB7, PB11, PB13, and PB14, as well as the quasi-planar cluster PB16, have similar structures to those of their corresponding bare boron clusters [23,50]. However, the ground-state structures of the planar PB4 and quasi-planar PB11, PB13, and PB16 clusters are different from their corresponding bare boron clusters due to the action of the P atom.
The planar wheel-shaped structure of B9 becomes the three-dimensional structure of PB9 with the doping of a P atom; the anion cluster B9, with a similar structure, changes from the planar wheel to the double-chain planar structure of PB9. The lowest energy configurations of PB14 and PB17 have quasi-two-dimensional structures, and PB17 has a three-dimensional cage-shaped structure. It is important to note the specific behaviors of PB7, PB9, and PB13. For the PB9 cluster, at the PBE0/6-311+G(d) level, the double-chain planar structure (Figure S14a) is more stable than the planar structure (Figure S14b), with an energy gap of 0.17 eV. At the higher level, CCSD(T)/6-311+G(d)//PBE0/6-311+G(d)+ZPE (zero-point energy corrections), both structures are almost degenerate in regards to energy, each with a tiny energy gap (that is, the planar structure (Figure S14b) is more stable than the double-chain planar structure (Figure S14a), with an energy gap of 0.01 eV). We consider the double-chain structure as the lowest energy structure of PB9; the same is true for the PB7 and PB13 clusters. It is worth mentioning that the umbrella-like structure found in PB7 is also the lowest energy structure in many metal-doped boron clusters, such as LiB7, BeB70/−, BeB80/−, and MgB8 [52,53,54]. Research shows that the low-valent actinide(III) boron clusters AnB7 (An = Pa, U, Np, and Pu), with umbrella-shaped structures, exhibit high electronic stability and can be obtained in the gas phase at room temperature [55]. Similarly, the neutral cluster PB7 also has an umbrella-shaped structure, with C6v symmetry. Therefore, in this study, we will highlight the umbrella-like structure of PB7, which can provide a theoretical basis for the design of highly stable boron-based nanomaterials. In addition, the quasi-planar structure of B15 becomes the three-dimensional half-sandwich-shaped structure of PB15 after P atom doping; thus, studying the structures and characteristics of PB15 is crucial to understanding the structural laws of doped nonmetallic boron clusters. In summary, the study of PB7 and PB15 is of great significance for the discovery of new stable boron-based nanomaterials.

2.2. Relative Stabilities

To analyze the relative stabilities of the lowest energy level of PBn0/− (n = 3−17), we calculated the average binding energy (Eb) and the second-order energy differences (∆2E) of the clusters using the following formulas, where n represents the number of boron atoms, and E is the total energy of the corresponding atom or cluster. The calculation results obtained using Equations (1)–(3) are plotted in Figure 3. The formulas are as follows:
Eb(PBn) = [nE(B) + E(P) − E(PBn)]/(n + 1)
Eb(PBn) = [(n − 1)E(B) + E(P) + E(B) − E(PBn)]/(n + 1)
Δ2E(PBn0/−) = E(PBn−10/−) + E(PBn+10/−) − 2E(PBn0/−)
Eb represents the inherent stability of the cluster, in that a larger value of Eb denotes higher relative stability. As evidenced in Figure 3, the Eb values of PBn0/− (n = 3−17) are gradually increasing with the increase in the boron atom number n, indicating that the cluster becomes more and more stable. Furthermore, the Eb values of the anions are overall more sizable than those of the corresponding neutral ions; thus, we can infer that the anion clusters of PBn (n = 3−17) are more stable than their corresponding neutral clusters, and that the excess electrons enhance the stability of the P-doped boron clusters. ∆2E is an important indicator of relative stability and can provide valuable insights into the stability of the clusters. In Figure 3, the second-order difference of the peaks found in n = 4, 8, 10, 12, 13, 14, 15, and 16 indicate that PB4, PB8, PB8, PB10, PB12, PB13, PB14, PB15, and PB16 are relatively more stable than their adjacent clusters.
The energy difference Egap between HOMO and LUMO is an indicator of chemical stability. A larger Egap value for a cluster implies that the electrons are more difficult to excite from HOMO to LUMO, which indicates that the corresponding structures are chemically inert. As can be seen in Figure 3, two obvious Egap peaks are found in PB7 (about 5.11 eV) and PB15 (about 4.31 eV), implying that the neutral clusters PB7 and PB15 exhibit better chemical stability than that of the other clusters. Therefore, combining ∆2E and Egap analyses proved that PB15 has a high relative and chemical stability. Additionally, due to the PB7 cluster having the largest values of Egap, as well as a special umbrella structure, we chose these clusters as examples through which to analyze chemical bonding in P-doped boron clusters.

2.3. Chemical Bonding Analysis

Through a comprehensive stability analysis of all of the clusters detailed in the previous section, we inferred that PB7 and PB15 show chemical stability, due to their high Egap values. We analyzed the charge populations of PB7 and PB15 (see Figure S33). Different from the metal elements in metal-doped boron clusters, which can undergo one or more electron transfers, the phosphorus atom in PB7 or PB15 transfers less than one electron. To further understand the electronic properties and stability of the closed-shell clusters PB7 and PB15, the AdNDP method was used to analyze their chemical bonding characteristics. The AdNDP method is an extension of the popular natural bond orbital (NBO) analysis, which can be used to analyze localized and delocalized multicenter bonds. We quantitatively analyzed the bonding properties of PB7 and PB15 using the AdNDP method, and the results are depicted in Figure 4 and Figure 5. In Figure 4, we clearly observe six two-center two-electron (2c–2e) σ bonds around the peripheral B atoms of the PB7 cluster. In addition, there are three 8c–2e σ bonds and three 8c–2e π bonds in PB7, which together support the plane of the cluster and enhance the stability of PB7. As can be seen in the Figure 4, the contribution of the P atom to the 8c–2e π bonds is greater than its contribution to the 8c–2e σ bonds. The delocalized σ and π bonds give rise to double aromaticity and fulfill the Hückel 4N + 2 rule with N = 1.
For PB15 (Figure 5), there is one lone pair on the P atom, and three 2c–2e σ bonds cover the B–P bond attached to the P atom. These three 2c–2e σ bonds are formed by three electrons on the P atom combining with an electron of each B atom on the B3 ring. In addition, nine 2c–2e σ bonds are distributed over the peripheral B–B bond. The seven 3c–2e σ bonds are divided into the four following groups: one 3c–2e σ bond is distributed on the inner ring triangle at the back side of the P atom, two 3c–2e σ bonds are distributed on two peripheral B3 rings, three 3c–2e σ bonds cover the three peripheral B3 rings behind the P atom, and the last 3c–2e σ bond is distributed on a ternary boron ring. Moreover, there are three 4c–2e σ bonds over three B4 rings, one 6c–2e σ bond attached to an internal B6 ring, and one 7c–2e σ bond over a B7 ring. On the whole, there are strong interactions between B15 and the P atom through delocalized bonds that stabilize the PB15 cluster. According to AdNDP investigations, the PB15 cluster possesses 12 delocalized σ bonds that do not satisfy spherical aromaticity [2(n + 1)2 rule].
To further confirm the above AdNDP analysis results, we analyzed PB7 and PB15 using the ELF method, a function that can be used to describe the localization and delocalization of different molecular regions. The ELF results for PB7 and PB15 are shown in Figure 6, as well as in Figures S31 and S32, respectively. As can be observed from Figure 6 and Figure S31, the isosurface map of PB7 covers six peripheral B–B bonds that correspond to six peripheral 2c–2e σ bonds, as well as the entire region of B and P atoms corresponding to the 8c–2e bonds. Combining Figure S32 and Figure 6 shows that the isosurface map of PB15 covers nine peripheral B–B bonds and three B–P bonds, which correspond to twelve 2c–2e σ bonds, as well as seven B3 triangles that correspond to seven 3c–2e σ bonds. Additionally, the three regions around PB15 are fatter, indicating the presence of 3c–2e and 4c–2e bonds.

2.4. Photoelectron Spectra

Photoelectron spectra (PES) is an effective approach for exploring the energy levels of valence electrons in nanoclusters. In photoelectron spectra, the positions of the peaks represent the energy differences between the initial and final electronic state after photons absorption. In order to identify the structures of the PBn (n = 3−17) clusters, we calculated the vertical detachment energies (VDEs) of the anionic clusters and simulated the photoelectron spectra of the PBn (n = 3−17) clusters using the time-dependent density functional theory (TD-DFT). The first few peaks of the photoelectron spectrum are commonly used to identify boron clusters [10,15]; thus, studying the peaks on the low-binding-energy side is of significant importance. Figure 7 shows the photoelectron spectra of the PBn (n = 3–17) clusters. According to the photoelectron spectra, the anion PB16 exhibits the largest first VDE value (3.88 eV, list in Table S1), while PB4 shows the lowest first VDE (1.78 eV). In addition, the energy gap (about 1.95 eV) between the first and second PB6 energy bands is the largest.
The first peaks of the photoelectron spectrum (except for those of PB4, PB5 and PB11) are derived from the calculated ground-state VDEs of PB3, PB6, PB7, PB8, PB9, PB10, PB12, PB13, PB14, PB15, PB16 and PB17 at 2.19, 2.47, 3.43, 3.24, 3.55, 3.35, 3.59, 3.44, 3.51, 3.53, 3.88, and 3.17 eV, respectively. The first peak of PB4 comes from the second VDE (2.2 eV). For open-shell PB5 and PB11, their first peaks come from the first and second VDEs (3.33 and 3.41 eV for PB5, 3.47 and 3.54 eV for PB11). Furthermore, the second peaks of PB7, PB8, PB9, PB10, PB12, PB14, PB15, PB16, and PB17 come from the second VDEs at 3.72, 4.06, 4.47, 3.88, 4.44, 4.3, 4.64, 3.9, 4.4, and 3.66 eV, respectively. The second peak of the photoelectron spectrum of PB4 comes from the ground-state VDE at 1.78 eV. However, the second peaks of PB3, PB6, and PB13 are derived from the second and third VDEs (2.736 and 2.744 eV for PB3; 4.4 and 4.43 eV for PB6; 4.27 and 4.33 eV for PB13). The second peak of PB5 comes from the third VDE (4.01 eV) and the fourth VDE (4.05 eV). For PB11, the second peak comes from the third VDE (3.74 eV). Furthermore, peaks with higher binding energies are derived from the separation of electrons from lower molecular orbitals.
Partially anionic boron clusters doped with P atoms have similar structures to those of the corresponding bare boron clusters; however, a comparison of their photoelectron spectra revealed that the addition of a phosphorus atom results in large changes in the photoelectron spectra of most clusters (except for PB11) [23,50]. For instance, the doped P atom causes the first peaks of PB3, PB6, and PB13 to move 0.63, 0.54, and 0.34 eV, respectively, towards the low-binding-energy side. At the same time, the first peaks of PB5, PB7, PB8, PB10, PB12, PB14, PB15, and PB16 moved 0.93, 0.58, 0.22, 0.29, 1.33, 0.41, 0.1, and 0.49 eV, respectively, to the high-binding-energy side. As can be seen from Figure 7, all photoelectron spectra are different, which indicates that the addition of P atoms not only alters the geometric structure of the clusters, but also leads to changes in their electronic structures. In summary, it is of great significance to study the peaks at the low-binding-energy side, as these simulated spectra may be used as fingerprints with which to identify PBn (n = 3–17) structures in the future.

3. Computation Details

In this paper, the geometrical structure searches for the neutral and anionic clusters of PBn0/− (n = 3–17) were performed using CALYPSO, an efficient and reliable method for searching for geometrical cluster configurations which has been successfully applied to the study of both boron clusters and doped boron clusters [25,27,33,36,40,54,56,57,58,59]. CALYPSO 5.0 software generates 70% of the structure in each generation, and the remaining 30% is formed randomly. When the size of a boron atom is in the range of n = 3–10, nearly 100–1200 isomers are initially obtained for each cluster. The number of isomers increases with the number of boron atoms; when n = 11–14, there are approximately 2000 isomers for each cluster, and when n = 15–17, the number of isomers increases to 2500.
After the initial structural search, the lower energy structures of PBn0/− (n = 3–17) were optimized at the PBE0/6-311+G(d) level, which is a reliable level for analyzing boron clusters, since the theoretical simulation values (photoelectron spectra) are consistent with the experimental values [15,25,38,59,60,61]. In order to obtain more accurate relative energy values, we performed CCSD(T) [62] calculations [CCSD(T)/6-311+G(d)//PBE0/6-311+G(d)]+ZPE(zero-point energy corrections) using the optimized PBE0 geometries for the collected isomers. The harmonic frequency and electronic structures were analyzed at the PBE0/6-311+G(d) level; therefore, the calculations below were obtained using the PBE0/6-311+G(d) and CCSD(T)/6-311+G(d)//PBE0/6-311+G(d)+ZPE methods, employing Gaussian16 software [63]. In addition, AdNDP and ELF are implemented in Multiwfn 3.7 software [64], and the AdNDP results were visualized using Visual Molecular Dynamics (VMD) 1.9.3 software [65].

4. Conclusions

In this work, the ground state structures of P-doped boron clusters PBn0/− (n = 3–17) were identified using the CALYPSO method. In addition, the bonding properties of PB7 and PB15 were discussed, based on AdNDP and ELF analyses, and the photoelectron spectra of the anionic clusters were also calculated. The conclusions can be summarized as follows: (1) The lowest-lying energy structures of PBn (n = 3–17) clusters, except for PB17, exhibit planar or quasi-planar structures. The lowest energy structures of PBn (n = 3–17) are planar or quasi-planar, with the exceptions of PB7, PB9, and PB15. The lowest energy structure of PB7 has an umbrella-like structure, with high symmetry (C6V), and the ground-state configuration of PB10 exhibits a chiral symmetric structure. (2) The lowest energy structures of the neutral ion PB15 have a half-sandwich structure, and they possess relatively high ∆2E and Egap values, indicating that they possess superior relative and chemical stability. (3) AdNDP bonding analysis and ELF analysis further verified the stability and validity of the lowest energy structures of PB7 and PB15. (4) The PBn anionic clusters (n = 3–17) exhibit different photoelectron spectra on their low-binding-energy sides, which can provide a theoretical basis for the identification of doped boron clusters.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29143384/s1, Figures S1–S30: Low-lying isomers of doped boron clusters PBn0/− (n = 3–17); Figure S31: Electron localization function (ELF) of PB7 with the isovalue set to 0.59; Figure S32: Electron localization function (ELF) of PB15 with the isovalue set to 0.7; Figure S33: Charge population of PB7 and PB15; Table S1: The first VDE values of bare B clusters and P-doped boron clusters.

Author Contributions

Conceptualization, S.L. and Y.Y.; methodology, D.W. and S.L.; software, S.L.; investigation, S.L.; data processing, D.W.; writing—original draft preparation, D.W.; writing—review and editing, S.L. and D.W.; funding acquisition, D.C. and Y.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), as well as the Growth Foundation for Young Scientists of the Education Department of Guizhou Province (Grant No: QJH KY[2022]310, QJJ [2022]260), 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 the Supplementary Materials.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Harmgarth, N.; Liebing, P.; Lorenz, V.; Engelhardt, F.; Hilfert, L.; Busse, S.; Goldhahn, R.; Edelmann, F.T. Synthesis and Structural Characterization of p-Carboranylamidine Derivatives. Molecules 2023, 28, 3837. [Google Scholar] [CrossRef] [PubMed]
  2. Diaw-Ndiaye, F.; Sanz Miguel, P.J.; Rodríguez, R.; Macías, R. The Synthesis, Characterization, and Fluxional Behavior of a Hydridorhodatetraborane. Molecules 2023, 28, 6462. [Google Scholar] [CrossRef]
  3. Das, S.; Shareef, M.A.; Das, B.C. Design and Synthesis of New Boron-Based Benzo[c][1,2,5] oxadiazoles and Benzo[c][1,2,5] thiadiazoles as Potential Hypoxia Inhibitors. Inorganics 2023, 11, 34. [Google Scholar] [CrossRef]
  4. Teixidor, F.; Núñez, R.; Viñas, C. Towards the Application of Purely Inorganic Icosahedral Boron Clusters in Emerging Nanomedicine. Molecules 2023, 28, 4449. [Google Scholar] [CrossRef]
  5. Sivaev, I.B.; Bregadze, V.I. Borane, Carborane and Metallacarborane Anions for Stabilization of Transient and Highly Reactive Intermediates; World Scientific: Singapore, 2018; Volume 1, pp. 147–203. [Google Scholar]
  6. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  7. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  8. Kroto, H.W.; Heath, J.R.; O’Brien, S.C.; Curl, R.F.; Smalley, R.E. C60: Buckminsterfullerene. Nature 1985, 318, 162–163. [Google Scholar] [CrossRef]
  9. Zhai, H.; Wang, L.; Alexandrova, A.N.; Boldyrev, A.I. Electronic structure and chemical bonding of B5 and B5 by photoelectron spectroscopy and ab initio calculations. J. Chem. Phys. 2002, 117, 7917–7924. [Google Scholar] [CrossRef]
  10. Zhai, H.; 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]
  11. 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]
  12. 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]
  13. 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] [PubMed]
  14. Bean, D.E.; Fowler, P.W. Double Aromaticity in “Boron Toroids”. J. Phys. Chem. C 2009, 113, 15569–15575. [Google Scholar] [CrossRef]
  15. 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] [PubMed]
  16. 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. 2014, 54, 941–945. [Google Scholar]
  17. 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]
  18. 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] [PubMed]
  19. Tang, C.; Zhang, X. The hydrogen storage capacity of Sc atoms decorated porous boron fullerene B40: A DFT study. Int. J. Hydrogen Energy 2016, 41, 16992–16999. [Google Scholar] [CrossRef]
  20. 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]
  21. 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]
  22. 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] [PubMed]
  23. 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]
  24. Piazza, Z.A.; Hu, H.S.; Li, W.L.; Zhao, Y.F.; Li, J.; Wang, L.S. Planar hexagonal B36 as a potential basis for extended single-atom layer boron sheets. Nat. Commun. 2014, 5, 3113. [Google Scholar] [CrossRef]
  25. 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] [PubMed]
  26. Jin, S.; Chen, B.; Kuang, X.; Lu, C.; Gutsev, G.L. Structural evolution and electronic properties of medium-sized boron clusters doped with scandium. J. Phys. Condens. Matter 2019, 31, 485302. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. 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]
  29. 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] [PubMed]
  30. 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]
  31. 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]
  32. Cheung, L.F.; Kocheril, G.S.; Czekner, J.; Wang, L.-S. Observation of Möbius Aromatic Planar Metallaborocycles. J. Am. Chem. Soc. 2020, 142, 3356–3360. [Google Scholar] [CrossRef]
  33. 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]
  34. 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]
  35. 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]
  36. 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]
  37. 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]
  38. Chen, B.L.; 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]
  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]
  40. 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]
  41. 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]
  42. 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] [PubMed]
  43. Tuyet Mai, D.T.; Van Duong, L.; Pham-Ho, M.P.; Nguyen, M.T. Electronic Structure and Properties of Silicon-Doped Boron Clusters BnSi with n = 15–24 and Their Anions. J. Phys. Chem. C 2020, 124, 6770–6783. [Google Scholar] [CrossRef]
  44. Van Duong, L.; Nguyen, M.T. Electronic structure of the boron fullerene B14 and its silicon derivatives B13Si+, B13Si and B12Si2: A rationalization using a cylinder model. Phys. Chem. Chem. Phys. 2016, 18, 17619–17626. [Google Scholar] [CrossRef]
  45. Li, P.F.; Zhai, H.J. Structures and chemical bonding of boron-based B12O and B11Au clusters. A counterexample in boronyl chemistry. Phys. Chem. Chem. Phys. 2022, 24, 10952–10961. [Google Scholar] [CrossRef]
  46. Li, S.X.; Yang, Y.J.; Chen, D.L. PB12+ and P2B12+/0/–: The Novel B12 Cage Doped by Nonmetallic P Atoms. ACS Omega 2023, 8, 44831–44838. [Google Scholar] [CrossRef] [PubMed]
  47. 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. [Google Scholar] [CrossRef] [PubMed]
  48. Lv, J.; Wang, Y.; Zhu, L.; Ma, Y. Particle-swarm structure prediction on clusters. J. Chem. Phys. 2012, 137, 084104. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
  51. 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]
  52. Gong, L.F.; Guo, W.L.; Wu, X.M.; Li, Q.S. B7 as a novel ligand: Theoretical investigations on structures and chemical bonding of LiB7 and BeB7+. Chem. Phys. Lett. 2006, 429, 326–334. [Google Scholar] [CrossRef]
  53. Kheshti, T.; Mahdavifar, Z.; Noorizadeh, S. Umbrella-shaped vs planar; evolutionary search for BnQ, Be©BnQ (n = 6–12, Q = 0, −1) clusters. J. Mol. Liq. 2021, 328, 115389. [Google Scholar] [CrossRef]
  54. Đorđević, S.; Radenković, S. Spatial and Electronic Structures of BeB8 and MgB8: How far Does the Analogy Go? Chem. Phys. Chem. 2022, 23, e202200070. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, N.; Wang, C.; Wu, Q.; Lan, J.; Chai, Z.; Shi, W. Highly stable actinide(iii) complexes supported by doubly aromatic ligands. Phys. Chem. Chem. Phys. 2022, 24, 5921–5928. [Google Scholar] [CrossRef] [PubMed]
  56. Lv, J.; Wang, Y.; Zhu, L.; Ma, Y. B38: An all-boron fullerene analogue. Nanoscale 2014, 6, 11692–11696. [Google Scholar] [CrossRef] [PubMed]
  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] [PubMed]
  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. 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]
  60. 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]
  61. 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]
  62. Bartlett, R.J.; Musiał, M. Coupled-cluster theory in quantum chemistry. Rev. Mod. Phys. 2007, 79, 291–352. [Google Scholar] [CrossRef]
  63. 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]
  64. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput.Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  65. 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 PBn, where pink balls represent boron atoms and orange ball represents phosphorus atom. (a) PB3 C2V; (b) PB4 CS; (c) PB5 CS; (d) PB6 CS; (e) PB7 C6V; (f) PB8 C2V; (g) PB9 CS; (h) PB10 C1; (i) PB11 C1; (j) PB12 CS; (k) PB13 C1; (l) PB14 CS; (m) PB15 CS; (n) PB16 C1; (o) PB17 CS.
Figure 1. Structures of PBn, where pink balls represent boron atoms and orange ball represents phosphorus atom. (a) PB3 C2V; (b) PB4 CS; (c) PB5 CS; (d) PB6 CS; (e) PB7 C6V; (f) PB8 C2V; (g) PB9 CS; (h) PB10 C1; (i) PB11 C1; (j) PB12 CS; (k) PB13 C1; (l) PB14 CS; (m) PB15 CS; (n) PB16 C1; (o) PB17 CS.
Molecules 29 03384 g001
Figure 2. Structures of PBn, where pink balls represent boron atoms and orange ball represents phosphorus atom. (a) PB3 C2V; (b) PB4 C2V; (c) PB5 CS; (d) PB6 CS; (e) PB7 C2V; (f) PB8 C2V; (g) PB9 CS; (h) PB10 C1; (i) PB11 CS; (j) PB12 CS; (k) PB13 C1; (l) PB14 CS; (m) PB15 C1; (n) PB16 C1; (o) PB17 C1.
Figure 2. Structures of PBn, where pink balls represent boron atoms and orange ball represents phosphorus atom. (a) PB3 C2V; (b) PB4 C2V; (c) PB5 CS; (d) PB6 CS; (e) PB7 C2V; (f) PB8 C2V; (g) PB9 CS; (h) PB10 C1; (i) PB11 CS; (j) PB12 CS; (k) PB13 C1; (l) PB14 CS; (m) PB15 C1; (n) PB16 C1; (o) PB17 C1.
Molecules 29 03384 g002
Figure 3. Average binding energy (Eb), second-order energy differences (Δ2E), and HOMO–LUMO gaps (Egap) of doped boron clusters PBn0/− (n = 3−17). Red represents the neutral cluster, blue represents the anion cluster, and n represents the number of boron atoms in the cluster.
Figure 3. Average binding energy (Eb), second-order energy differences (Δ2E), and HOMO–LUMO gaps (Egap) of doped boron clusters PBn0/− (n = 3−17). Red represents the neutral cluster, blue represents the anion cluster, and n represents the number of boron atoms in the cluster.
Molecules 29 03384 g003
Figure 4. AdNDP analysis of PB7. ON is the occupation number, and the yellow ball is the P atom.
Figure 4. AdNDP analysis of PB7. ON is the occupation number, and the yellow ball is the P atom.
Molecules 29 03384 g004
Figure 5. AdNDP analysis of PB15. ON is the occupation number, and the yellow ball is the P atom.
Figure 5. AdNDP analysis of PB15. ON is the occupation number, and the yellow ball is the P atom.
Molecules 29 03384 g005aMolecules 29 03384 g005b
Figure 6. ELF at the PBE0 level. (a) C6v PB7; the isovalue is 0.8. (b) CS PB15; the isovalue is 0.8.
Figure 6. ELF at the PBE0 level. (a) C6v PB7; the isovalue is 0.8. (b) CS PB15; the isovalue is 0.8.
Molecules 29 03384 g006
Figure 7. Photoelectron spectra at the PBE0/6-311+G * level: (a) PB3; (b) PB4; (c) PB5; (d) PB6; (e) PB7; (f) PB8; (g) PB9; (h) PB10; (i) PB11; (j) PB12; (k) PB13; (l) PB14; (m) PB15; (n) PB16; (o) PB17.
Figure 7. Photoelectron spectra at the PBE0/6-311+G * level: (a) PB3; (b) PB4; (c) PB5; (d) PB6; (e) PB7; (f) PB8; (g) PB9; (h) PB10; (i) PB11; (j) PB12; (k) PB13; (l) PB14; (m) PB15; (n) PB16; (o) PB17.
Molecules 29 03384 g007
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

Wang, D.; Yang, Y.; Li, S.; Chen, D. Structural Evolution of Small-Sized Phosphorus-Doped Boron Clusters: A Half-Sandwich-Structured PB15 Cluster. Molecules 2024, 29, 3384. https://doi.org/10.3390/molecules29143384

AMA Style

Wang D, Yang Y, Li S, Chen D. Structural Evolution of Small-Sized Phosphorus-Doped Boron Clusters: A Half-Sandwich-Structured PB15 Cluster. Molecules. 2024; 29(14):3384. https://doi.org/10.3390/molecules29143384

Chicago/Turabian Style

Wang, Danyu, Yueju Yang, Shixiong Li, and Deliang Chen. 2024. "Structural Evolution of Small-Sized Phosphorus-Doped Boron Clusters: A Half-Sandwich-Structured PB15 Cluster" Molecules 29, no. 14: 3384. https://doi.org/10.3390/molecules29143384

Article Metrics

Back to TopTop