Next Article in Journal
Influence of Electronic Non-Equilibrium on Energy Distribution and Dissipation in Aluminum Studied with an Extended Two-Temperature Model
Previous Article in Journal
Fabrication of Ultrafine, Highly Ordered Nanostructures Using Carbohydrate-Inorganic Hybrid Block Copolymers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantum Chemistry Study on the Structures and Electronic Properties of Bimetallic Ca2-Doped Magnesium Ca2Mgn (n = 1–15) Clusters

1
Quantum Materials Research Center, College of Physics and Electronic Engineering, Zhengzhou Normal University, Zhengzhou 450044, China
2
School of Chemical Engineering and Energy, Zhengzhou University, Zhengzhou 450001, China
3
School of Physics and Electronic Engineering, Sichuan University of Science & Engineering, Zigong 643000, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(10), 1654; https://doi.org/10.3390/nano12101654
Submission received: 18 April 2022 / Revised: 8 May 2022 / Accepted: 10 May 2022 / Published: 12 May 2022

Abstract

:
Here, by utilizing crystal structure analysis through the particle swarm optimization (CALYPSO) structural searching method with density functional theory (DFT), we investigate the systemic structures and electronic properties of Ca2Mgn (n = 1–15) clusters. Structural searches found that two Ca atoms prefer to occupy the external position of magnesium-doped systems at n = 2–14. Afterward, one Ca atom begins to move from the surface into the internal of the caged skeleton at n = 15. Calculations of the average binding energy, second-order difference of energies, and HOMO–LUMO gaps indicated that the pagoda construction Ca2Mg8 (as the magic cluster) has higher stability. In addition, the simulated IR and Raman spectra can provide theoretical guidance for future experimental and theoretical investigation. Last, further electronic properties were determined, including the charge transfer, density of states (DOS) and bonding characteristics. We hope that our work will provide theoretical and experimental guidance for developing magnesium-based nanomaterials in the future.

1. Introduction

Nanomaterials with small particle sizes, specific surface areas and high surface energies possess wide applications in chemistry, physics, biology, medicine, materials and nanodevices. Magnesium atoms contain s2 closed-shell electron configuration similarly to helium, which plays an essential role in aerospace, mobile electronics, automobile and biomedical applications [1,2,3,4]. Currently, metal-doped magnesium clusters with unique geometries and fascinating electronic properties have received considerable attention in magnesium-based multi-function materials.
In the past decades, many experimental techniques and theoretical studies have been reported for the related structures and properties of pure magnesium clusters [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19]. For instance, on the experimental side, the transition points of Mgn clusters were determined at n = 20 by Diederich’s group [5]. By measuring the photoelectron spectra and observing the sp band gaps of Mgn (n = 3–35) clusters, Thomas et al. found that the anion magnesium clusters exhibited a metallic character from n = 18 [6]. On the theoretical side, the ground state van der Waals potential of a magnesium dimer was described by five essential parameters of the Tang–Toennies potential model [7].
Janecek et al. studied the structures of neutral and cationic Mgn (n to 30) clusters using the local spin density functional [8]. Based on the MP4 (SDTQ) and CCSD(T) levels, the electron affinities were calculated for magnesium dimers and trimers [9]. By using the spin-unrestricted density functional theory with a local density approximation, Gong and co-workers studied the electronic structures of Mgn (n < 57) clusters [10]. Their results found that more 3p electrons will be involved in the sp hybridization with the cluster size increasing. More recently, Akola and co-workers focused on the structural and electronic properties of Mgn (n < 13) clusters [11]. Their investigations showed that the onset of metallization of Mgn clusters is difficult to assign due to the energy gap and sp hybridization.
Subsequently, electron binding energies, structural and electronic properties and the nonmetal-to-metal transition were studied systemically by Jellinek and Acioli [12,13]. In addition, by using ab initio theoretical methods (B3PW91, B3LYP and MP4), Lyalin et al. investigated the structure and electronic properties of neutral and singly charged Mgn (n = 2–21) clusters [14]. The above study by Lyalin et al. suggested that the hexagonal ring structure determines the cluster growth from Mg15. Moreover, the electronic shell effects and the Jellium-like behavior manifest themselves in the formation of geometrical properties; however, the shell effects do not determine the geometry of the Mg clusters completely.
In addition, the energetic structural properties of neutral magnesium clusters Mgn (n = 2–22) were investigated utilizing density functional theory [15]. More recently, Heidari and Xia’s groups presented further details about the structural transition and electronic properties of Mgn (n = 10–56) and Mgn0/− (n = 3–20) clusters, respectively [16,17]. Recently, Duanmu’s group investigated the geometries, binding energies, adiabatic ionization potentials and adiabatic electron affinities of Mgn+/0/− (n = 1–7) clusters [18] and cohesive energies by using the CCSD(T) scheme with MP2/CBS correction for Mgn (n = 2–19 and 28), respectively [19].
Most recently, research found that heteroatom doping is an effective strategy to stabilize geometrical structures or to tune electronic properties. Up to now, based on the different quantum chemistry calculations, studies on metal-doped (Be, Al, Ge, Sn, 3d and 4d TM atoms) and nonmetal-doped (B, C, N, O, F and Si) magnesium clusters have harvested many great achievements [20,21,22,23,24,25,26,27,28,29,30,31,32]. More importantly, based on the CALYPSO structural searching method and DFT, the structures and electronic properties of Be-, Be2-, Sr2- and Ba2-atom-doped differently sized magnesium clusters have been systemically discussed by Zeng’s, Zhao’s and our groups [33,34,35,36,37]. For example, Zeng and co-workers found that BeMg90, BeMg9+1 and BeMg8−1 clusters possess relative highly stability out of the studied systems [33].
Moreover, the stability mainly originating from the σ-type covalent bond, is formed by the interaction between Be-s and Mg-p orbitals. Zhao’s group performed the structural and electronic properties of BeMgn (n = 10–20) clusters and their anions [34]. The research concluded that the position of Be atom changes from completely encapsulated sites to surface sites after reverting to the caged magnesium motif. Subsequently, the structures and electronic properties have been investigated for two Be-, Sr- and Ba-atom-doped small-sized magnesium clusters by our group [35,36,37]. For Be2Mgn (n = 1–20) clusters, from n = 10, the structures transfer from 3D to filled cage-like frameworks [35].
Furthermore, in the small size, one Be atom prefers the surface sites, and the other Be atom tends to embed inside magnesium motif. However, for the large size clusters (n > 18), two Be atoms were completely encapsulated into magnesium cages. In addition, the Be2Mg8 cluster possesses robust stability, and Be-2p and Mg-3p orbitals revealed increasing metallic behavior. More interestingly, based on the same calculated method, studies on the structural evolution and electronic properties were performed for SrMgn0/− (n = 1–12) clusters [36]. As a result, the tower-like framework of the Sr2Mg8 cluster possesses higher stability out of the studied systems.
Moreover, the stronger sp hybridization leads to stronger Sr-Mg bonds, which is supported by the analysis of the multi-center bonds. Subsequently, our groups systemically reported the structures and electronic properties of two-barium-atom-doped magnesium in both neutral and anionic species [37]. A pagoda-like Ba2Mg8 was determined by analyzing the relative stability.
Analysis of the molecular orbitals indicated that the high stability comes mainly from the interaction between Ba-6s and Mg-3p orbitals. In conclusion, the stronger sp hybridization leads to stronger M-Mg bonds and metallic behavior. Last, the geometric structures and electronic properties have been investigated systemically for lithium-doped magnesium clusters [38]. The results indicated that lithium atoms prefer to occupy the convex sites of LiMgn structures. The LiMg9 cluster possesses relatively higher stability. In addition, the charges transfer from the Li to Mg atoms, and there exists strong hybridization among sp orbitals.
Thus, metal-atom-doped magnesium clusters provide an effective approach to creating novel structures and electronic properties. As the same group of alkaline earth metals, Ca and Be, Mg, Sr and Ba have the identical valence electronic configuration of ns2. Nevertheless, due to the different electronegativity and atomic radius of Ca atoms compared with Be, Mg, Sr and Ba atoms, are there similar frameworks for two-calcium-atom-doped magnesium clusters? If yes, do these clusters possess novel electronic and bonding properties? How does hybridization change?
To date, minor investigations on the structures and electronic properties of two-calcium-atom-doped magnesium clusters have been reported in theoretical calculations and experimental works. Thus, in the present work, motivated by Be-, Mg-, Sr- and Ba-doped magnesium, we performed a systematic investigation for two-calcium-atom-doped magnesium Ca2Mgn (n = 1–15) clusters.
First, we conducted wide structural searching and precise structural optimization to explore the structural evolution rule. Second, determining the stable configuration of Ca2Mgn (n = 1–15) clusters was conducted by analyzing the stability properties. Finally, some electronic properties, such as the charge transfer, IR and Raman spectra, DOS and bonding characteristics, are discussed for Mg-doped alkaline-earth clusters. We hope that our investigations will provide a theoretical and experimental basis for studying the microscopic mechanism of magnesium doped with alkaline-earth nanomaterials.

2. Computational Detail

In the present section, a stochastic global search algorithm based on the CALYPSO structural searching method was used to obtain the lowest and low-lying isomers of Ca2Mgn (n = 1–15) clusters and Mgn+2 (n = 1–15) clusters [39,40,41]. Based on this method, the corresponding stable or metastable structures can be successful with the chemical composition and given external conditions [42,43,44,45,46,47,48,49,50], and the detailed searching process can be found in our reported papers [42,43,44,45,46,47]. The structural optimization and energy calculations employed the B3PW91 functional and 6−311+G(d) basis set for Ca and Mg atoms, respectively [51,52].
In particular, the B3PW91 functional has been widely tested for magnesium and magnesium-based clusters [35,36]. In addition, the spin multiplicity (1, 3, 5 and 7) is included, no imaginary frequencies are validated. All calculations were performed using the Gaussian09 program package [53]. In the following works, the relative stabilities of the ground state Ca2Mgn (n = 1–15) and Mgn+2 (n = 1–15) clusters were studied by computing the average binding energy (Eb) and second-order difference energy ( Δ 2 E ). Subsequently, the IR and Raman spectra, DOS, molecular orbitals and AdNDP were systemically investigated using the Multiwfn software for the studied clusters [54]. To ensure the reliability of our computational method, the bond length (re), vibration frequencies (ωe) and dissociation energies (De) were calculated for Mg2, Ca2 and CaMg dimers, respectively.
For the Mg2 dimer, our calculated values were re = 3.651 Å, De = 0.0790 eV, which are in good agreement with the experimental results (3.891 Å, 0.0866 eV) [55]. For the Ca2 and CaMg dimers, there are no experimental values available. Our calculated results for the bond lengths, vibration frequencies and dissociation energies are 4.2667 Å, 72.32 cm1 and 0.1478 V for the Ca2 dimer and 3.909 Å, 82.63 cm1 and 0.1114 V for the CaMg dimer, respectively. Moreover, the bond length and frequency of Ca2 dimer are also in excellent agree with Soltani’s theoretical values (4.285 Å and 65.2 cm1), respectively [56]. This indicated the reliability of the proposed method in this work.

3. Results and Discussions

3.1. Geometric Structures

In Figure 1, the lowest and lower-lying energy structures are present for Ca2Mgn (n = 1–15) and Mgn+2 (n = 1–15) clusters. These isomers with energies from low to high are designated by na, nb, nc and nd. Moreover, the electronic states and point symmetry are also collected in Table 1. Simultaneously, the Cartesian coordinates of the lowest energy structures of Ca2Mgn (n = 1–15) clusters are given in Table S1 in the Supplementary Materials.
First, the lowest energy structures of the Mgn clusters agree with previous research by Zhang and Li et al. [35,37], which indicates that the present functional and basis sets are reliable. Moreover, the Ca2Mgn (n = 1–8) clusters possess similar geometric structures compared with the Mgn+2 clusters. Second, Ca2Mg possesses a triangular plane structure, and the lowest energy structures of Ca2Mgn (n = 2–15) clusters maintain a three-dimensional (3D) configuration.
Specifically, Ca2Mgn (n = 2–6) clusters can be obtained by adding one Mg atom to the Ca2Mgn−1 clusters, and Ca2Mgn (n = 7–9) clusters can be generated from the substitution of Mgn+2 clusters by two Ca atoms, respectively. Interestingly, the lowest energy structure of the Ca2Mg8 cluster possesses the same geometrical form as those of X2Mg8 (X = Be, Sr and Ba) clusters. Finally, from 2 to 14, the doped two Ca atoms prefer to locate outside the host Mgn+2 cluster. At n = 15, one Ca atom starts to move from the surface into the internal of the caged skeleton.

3.2. Relative Stability

To determine the relative stabilities of Ca2Mgn and Mgn+2 (n = 1–15) clusters, the average binding energy (Eb) and second-order difference of energy Δ 2 E are calculated as follows:
E b Ca 2 Mg n = n E ( Mg ) + 2 E Ca ) E ( Ca 2 Mg n n + 2 E b Mg n = n E Mg ) E ( Mg n n Δ 2 Ca 2 Mg n = E Ca 2 Mg n 1 + E Ca 2 Mg n + 1 2 E Ca 2 Mg n Δ 2 E Mg n = E Mg n 1 + E Mg n + 1 2 E Mg n
E denotes the total energy of the corresponding clusters or atoms. The calculated results are plotted in Table 1 and Figure 2A,B. The following information can be concluded: (1) The values of Eb(Mgn+2) clusters are lower than those of Eb(Ca2Mgn) clusters, indicating that the doped Ca atom can improve the stability of pure magnesium clusters. (2) With an increase of Mg atoms, the values of Eb (Mgn+2 and Ca2Mgn) monotonically increase, which indicates an enhanced effect on the stabilities for Ca2Mgn and Mgn+2 (n = 1–15) clusters. (3) Δ 2 E values show a certain degree of oscillation, and there exist four visible peaks in the curves at n = 2, 5, 8, 11 and 13, implying that the Ca2Mg2,5,8,11,13 and Mg4,7,10,13,15 clusters possess higher relative stability. Most interestingly, the Ca2Mg8 cluster with 20 valence electrons possesses an notably high Δ 2 E value (0.66 eV), which indicates superior relative stability.
The HOMO–LUMO energy gap (Eg) is the other powerful tool to study the relative stabilities. Generally speaking, large values indicate stronger stability. In the present section, the HOMO–LUMO energy gaps are presented in Table 2 and Figure 2C. First, values of Eg(Mgn+2) are larger than those of Eg(Ca2Mgn) clusters, indicating that Ca2Mgn clusters are more stable, which is in reasonable agreement with previous research on averaged binding energies.
Second, Ca2Mg2,6,8,9,15 clusters with local maxima of Egap suggests that those clusters are more stable than their neighbors. In summary, combining the conclusions of Eb, Δ2E and Eg values, the Ca2Mg8 cluster corresponds to the magic numbers and exhibits robust stability.

3.3. Charge Transfer

In this section, the charge-transfer information is analyzed by natural population analysis (NPA) in Table 1 and Figure 2D. Clearly, the doped Ca atoms possess positive charges in doped systems, meaning the charges transfer from calcium to magnesium atoms.
Thus, the role of magnesium atoms is as charge acceptors, and Ca atoms are the charge donors. This is expected as Ca (1.00) has a smaller electronegativity compared with that of Mg (1.31) [57]. Second, the transferred charges increase with increasing cluster size. However, the value falls sharply at n = 15, and this situation may be the result of the position of Ca atom in the caged skeleton. In addition, the transfer charges in the range of n = 5–14 are all greater than 1.0 eV except for Ca2Mg6.

3.4. Infrared (IR) and Raman Spectra

In order to facilitate the characterization of spectra, we computationally simulated the infrared (IR) and Raman spectra of the lowest energy structures of Ca2Mg8 cluster. The simulated spectra with atomic labels are shown in Figure 3. The results found the highest intense IR frequency was located at 197.74 cm−1 with 4Mg-7Mg-10Mg bond tensile vibration; however, two Ca atoms were almost silent. The second and third strongest peaks can be found at 129.62 and 218.78 cm−1. For Raman spectra, the strongest peak at 181.51 cm−1 corresponds to the breathing vibration of all atoms. The second- and third-strongest Raman frequencies at 138.28 and 129.62 cm−1 correspond to the swing vibration of all atoms. In addition, IR and Raman spectra revealed that the strongest spectral frequencies are displayed in the range of 100–200 cm−1.

3.5. The Density of States

To understand the nature of the chemical bonding, the total density of states (TDOS) and partial density of states (PDOS) of the lowest energy structures of Ca2Mgn (n = 1–15) are displayed in Figure 4. The TDOS is represented by the khaki shade; PDOS of Ca-s, Ca-p, Mg-s and Mg-p AOs (atomic orbitals) are represented by the red and blue solid lines as well as magenta and green dotted lines, respectively. We found that the contribution to TDOS mainly comes from the PDOS of Ca-s, Mg-s and Mg-p AOs in the region of occupied orbitals.
This indicates that sp hybridization has occurred in Ca-Mg atoms and Mg-Mg atoms. In fact, the sp hybridization of Mg-Mg promotes the formation of Mgn frames in Ca2Mgn clusters. The sp hybridization of Ca-Mg can promote the interaction between the two doped calcium atoms and magnesium frames of Ca2Mgn clusters, which is also the main reason why the stability of Ca2Mgn is higher than that of their corresponding pure magnesium clusters.

3.6. Bonding Characters

Based on the above analyses, a pagoda-like Ca2Mg8 structure possesses superior stability. To illustrate the source of higher stability, the bonding nature, such as the MOs (molecular orbitals) and multi-center bonds, are discussed for the lowest energy structure of Ca2Mg8 cluster. The molecular energy levels and corresponding orbitals are presented in Figure 5. First, calcium and magnesium atoms are composed of the same valence configuration of s2, and Ca2Mg8 with 20 valence electrons meets the requirement of the Jellium model in terms of the valence electron number.
Moreover, the shell structures consist of one 1S orbital, three 1P orbital, five 1D orbitals and one 2S orbital, all of which are occupied by the paired electrons. The energy of 1S, 1P, 1D and 2S states are arranged in order from low to high without energy levels overlapping, and the energy levels are also relatively concentrated. In addition, all the splitting energy levels of 1D orbital are lower in energy than the 2S orbital. As a result, Ca2Mg8 cluster is a closed shell 1S21P61D102S2 filled with 20 valence electrons.
In addition, the contributions of molecular orbital for Ca2Mg8 cluster were probed utilizing Multiwfn 3.8 program. The HOMO corresponding to 2S state involves Ca-s (30.98%), Mg-s (19.53%) and Mg-p (44.86%). The HOMO-m (m = 1–5) exhibit 1D state, in which HOMO-1 and HOMO-4 are mainly composed of Ca-s (10.62%, 26.60%), Mg-s (41.87%, 16.99% for) and Mg-p (43.70%, 53.33%), respectively, and the remaining 1D state orbital is mainly contributed to by Mg atoms. The compositions of HOMO-6, 7, 8 corresponding to three 1P orbitals comes mainly from the Mg-s and p AOs as well as to a small extent from the Ca-s and p AOs. In the case of HOMO-9 (1S), Mg atoms provide more than 90% of the orbital contribution. Hence, the sp hybridization between the Ca and Mg atoms could promote the interaction between the doped-Ca and host-Mg atoms and form stronger Ca-Mg bonds.
To further elucidate the bonding patterns of the Ca2Mg8 cluster, adaptive natural density partitioning (AdNDP) is included, representing the bonding of a molecule in terms of n-center two-electron (nc-2e) bonds. Notably, 1c-2e and 2c-2e bonds mean a localized character; the nc-2e bond (n ≥ 3) belongs to the delocalized character. For the Ca2Mg8 cluster, ten delocalized bonds with different occupation number (ON) are shown in Figure 6, which corresponds to three delocalized 3c-2e σ-bonds, two delocalized 4c-2e σ-bonds, two delocalized 5c-2e π-bonds, two delocalized 7c-2e π-bonds and one delocalized 9c-2e π-bond, respectively.
Three 3c-2e σ-bonds with an ON of 1.745–1.766 |e| are formed by two opposite trigonal 5Mg-8Mg-10Mg and 3Mg-6Mg-7Mg units and one 2Ca-5Mg-3Mg unit. Two 4c-2e σ-bonds are accountable for the bonding between the quadrangle 4Mg-6Mg-8Mg-9Mg units (ON = 1.89 |e|) and pyramid 1Ca-7Mg-9Mg-10Mg units (ON = 1.774 |e|), respectively. There are two 5c-2e π-type bonds with ON = 1.80 and 1.792 |e|, derived from two Ca-Mg units of 1Ca-2Ca-3Mg-5Mg-9Mg and 1Ca-3Mg-4Mg-5Mg-9Mg units, respectively. Two 7c-2e ON = 1.945, 1.903 |e| (two symmetric units) and one 9c-2e ON = 1.994 |e| exhibit π-type bonding character. In addition, to deeply understand the nature of bonding, the Wiberg bond orders are calculated and listed in Table 3. The results indicate that the strong Ca-Mg bonds are greater than Ca-Ca and most Mg-Mg bonds.

4. Conclusions

In summary, a detailed investigation of the structures and electronic properties of Ca2Mgn (n = 1–15) was performed using the CALYPSO searching method and DFT calculations. Structural searching found that two Ca atoms prefer to occupy the external position of magnesium-doped systems at n = 2–14 and that one Ca atom tends to move from the surface into the internal of the caged skeleton at n = 15. The size-dependent binding energies, second-order difference of energies, and HOMO–LUMO gaps found a pagoda-like Ca2Mg8 as a magic cluster that possessed higher stability.
Upon charge transfer analysis, charges transferred from calcium to magnesium atoms. The simulated IR and Raman spectra of the magic cluster revealed that the strongest spectral frequencies were displayed in the range of 100~200 cm−1. In addition, the high stability of Ca2Mg8 with a 20 valence electron cluster possessed a closed-shell electron configuration of 1S21P61D102S2 in terms of the Jellium model. Last, the sp hybridization of Ca-Mg and Mg-Mg bonds was confirmed by the molecular orbitals and AdNDP, which contribute to the high stability of the Ca2Mg8 cluster.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12101654/s1, Table S1: Cartesian coordinates for the lowest energy structures of Ca2Mgn (n = 1–15) clusters.

Author Contributions

C.L.: investigation, writing—original draft and funding acquisition. Y.C.: investigation and writing—review and editing. H.T. and B.R.: investigation and software. Q.L. and Y.L.: writing—review and editing and supervision. H.Y.: conceptualization and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Natural Science Foundation of China (No. U1404109, 11504334, 11904328 and 12104416), China Postdoctoral Science Foundation funded project (No. 2017M623310XB), Training Plan for Young backbone Teachers in Zhengzhou Normal University (No. QNGG-211361), Provincial student’s project for innovation and entrepreneurship training program (S202012949018 and S202112949002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are included in the main text.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nanan, M.E.; Bierwagen, G.P. Mg-Rich coatings: A new paradigm for Cr-free corrosion protection of Al arrospece alloy. JCT Res. 2014, 1, 69–80. [Google Scholar]
  2. Yoo, J.; Aksimentiev, A. Improved parametrization of Li+, Na+, K+ and Mg2+ ions for all-atom molecular dynamics simulations of nucleic acid systems. J. Phys. Chem. Lett. 2012, 3, 45–50. [Google Scholar] [CrossRef]
  3. Pollock, T.M. Weight loss with magnesium alloys. Science 2010, 328, 986–987. [Google Scholar] [CrossRef] [PubMed]
  4. Kim, S.H.; You, B.S.; Yim, C.D. Texture and microstructure changes in asymmetrically hot rolled Az31 magnesium alloy sheets. Mater. Lett. 2005, 59, 3876–3880. [Google Scholar] [CrossRef]
  5. Diederich, T.; Doppner, T.; Braune, J.; Tiggesbaumker, J.; Meiwes-Broer, K.H. Electron delocalization in magnesium clusters grown in superclod helium drolpets. Phys. Rev. Lett. 2001, 86, 4807–4810. [Google Scholar] [CrossRef]
  6. Thomas, O.C.; Zheng, W.J.; Xu, S.J.; Bowen, K.H. Onset of metallic behavior in magnesium clusters. Phys. Rev. Lett. 2002, 89, 213403. [Google Scholar] [CrossRef]
  7. Li, P.; Xie, W.; Tang, K.T. The van der Waals potential of the magnesium dimer. J. Chem. Phys. 2010, 133, 084308. [Google Scholar] [CrossRef]
  8. Janecek, S.; Krotscheck, E.; Liebrecht, M.; Wahl, R. Structure of Mgn and Mgn+ clusters up to n = 30. Eur. Phys. J. D-At. Mol. Opt. Plasma Phys. 2011, 63, 377–390. [Google Scholar] [CrossRef]
  9. Kaplan, I.G.; Diaz, C.C. Comparative study of the electron affinities of beryllium and magnesium dimmers and trimers. Int. J. Quantum Chem. 2005, 104, 468–474. [Google Scholar] [CrossRef]
  10. Gong, X.G.; Zheng, Q.Q.; He, Y.Z. Electronic structures of magnesium clusters. Phys. Lett. A 1993, 181, 459–464. [Google Scholar] [CrossRef]
  11. Akola, J.; Rytkonen, K.; Manninen, M. Metallic evolution of small magnesium clusters. Eur. Phys. J. D 2001, 16, 21–24. [Google Scholar] [CrossRef]
  12. Jellinek, J.; Acioli, P.H. Magnesium Clusters: Structural and Electronic Properties and the Size-Induced Nonmetal-to-Metal Transition. J. Phys. Chem. A 2002, 106, 10919–10925. [Google Scholar] [CrossRef]
  13. Acioli, P.H.; Jellinek, J. Electron binding energies of anionic magnesium clusters and the nonmetal-to-metal transition. Phys. Rev. Lett. 2002, 89, 213402. [Google Scholar] [CrossRef]
  14. Lyalin, A.; Solov’yov, I.A.; Solov’yov, A.V.; Greiner, W. Evolution of the electronic and ionic structure of Mg clusters with increase in cluster size. Phys. Rev. A 2003, 67, 063203. [Google Scholar] [CrossRef] [Green Version]
  15. Kohn, A.; Weigend, F.; Ahlrichs, R. Theoretical study on clusters of magnesium. Phys. Chem. Chem. Phys. 2001, 23, 711–719. [Google Scholar] [CrossRef]
  16. Herdari, I.; Ghazi, S.M.; Goedecker, S.; Kanhere, D.G. Growth and structural properties of Mgn (n = 10–56) clusters: Density functional theory study. J. Phys. Chem. A 2011, 115, 12307–12314. [Google Scholar] [CrossRef]
  17. Xia, X.X.; Kuang, X.Y.; Lu, C.; Jin, Y.Y.; Xing, X.D.; Merino, G. Deciphering the structural evolution and electronic properties of magnesium clusters: An aromatic homonuclear metal. J. Phys. Chem. A 2016, 120, 7947–7954. [Google Scholar] [CrossRef] [Green Version]
  18. Duanmu, K.; Roberto-Neto, O.; Machado, F.B.C.; Hansen, J.A.; Shen, J.; Piecuch, P.; Truhlar, D.G. Geometries, binding energies, ionization potentials, and electron affinities of metal clusters: Mgn 0,±1, n = 1–7. J. Phys. Chem. C 2016, 120, 13275. [Google Scholar] [CrossRef]
  19. Duanmu, K.; Friedrich, J.; Truhlar, D.G. Thermodynamics of metal nonparties: Energies and enthalpies of formation of magnesium clusters and nanoparticles as large as 1.3 nm. J. Phys. Chem. C 2016, 120, 26110. [Google Scholar] [CrossRef]
  20. Kong, F.J.; Hu, Y.F. Density functional theory study of small X-doped Mgn (X = Fe, Co, and Ni, n = 1–9) bimetallic clusters: Equilibrium structures, stabilities, electronic and magnetic properties. J. Mol. Model. 2014, 20, 2087. [Google Scholar] [CrossRef]
  21. Ge, G.X.; Han, Y.; Wan, J.G.; Zhao, J.J.; Wang, G.H. First-principles prediction of magnetic superatoms in 4d-transition-metal-doped magnesium clusters. J. Chem. Phys. 2013, 139, 174309. [Google Scholar] [CrossRef] [PubMed]
  22. Li, Z.; Zhao, Z.; Zhou, Z.H.; Wang, H.B.; Li, S.L. First-principles calculations on small MgnZn and Mgn−1Zn2 clusters: Structures, stabilities, and electronic properties. Mater. Chem. Phys. 2017, 199, 585–590. [Google Scholar] [CrossRef]
  23. Xing, X.D.; Wang, J.J.; Kuang, X.Y.; Xia, X.X.; Lu, C.; Maroulis, G. Probing the low-energy structures of aluminum-magnesium alloy clusters: A detailed study. Phys. Chem. Chem. Phys. 2016, 18, 26177–26183. [Google Scholar] [CrossRef] [PubMed]
  24. Trivedi, R.; Bandyopadhyay, D. Study of absorption and dissociation pathway of H2 molecule on MgnRh (n = 1–10) clusters: A first principle investigation. Int. J. Hydrogen Energy 2016, 41, 20113–20121. [Google Scholar] [CrossRef]
  25. Afshar, M.; Chegeni, M.H. Spin and orbital magnesium in XMg8 (X = Sc-Ni): A relativistic density functional theory study. Mol. Phys. 2016, 114, 3620–3628. [Google Scholar] [CrossRef]
  26. Medel, V.M.; Ulises, J.; Khanna, S.N.; Chauhan, V.; Sen, P.; Castleman, A.W. Hund’s rule in superatoms with transition metal impurities. Proc. Natl. Acad. Sci. USA 2011, 108, 10062–10066. [Google Scholar] [CrossRef] [Green Version]
  27. Zhu, B.C.; Zhang, S.; Zeng, L. The effect of silicon doping on the geometrical structures, stability, and electronic and spectral properties of magnesium clusters: DFT study of SiMgn (n = 1–12) clusters. Int. J. Quantum Chem. 2020, 120, e26143. [Google Scholar] [CrossRef]
  28. Medel, V.M.; Reber, R.C.; Reveles, J.U.; Khanna, S.N. Metallic and molecular orbital concepts in XMg8 clusters, X = Be-F. J. Chem. Phys. 2012, 136, 134311. [Google Scholar] [CrossRef] [Green Version]
  29. Zeng, L.; Liang, M.K.; Wei, X.F.; Guo, J.; Zhu, B.C. New potential stable structures of XMgn (X = Ge, C, Sn; n = 1–12) clusters: XMg8 with high stability. J. Phys. Condens. Matter 2021, 33, 065302. [Google Scholar] [CrossRef]
  30. Trivedi, R.; Bandyopadhyay, D. Hydrogen storage in small size MgnCo clusters: A density functional study. Int. J. Hydrogen Energy 2015, 40, 12727–12735. [Google Scholar] [CrossRef]
  31. Lu, C.; Ma, Y.L.; Li, F.; Zhu, H.; Zeng, X.Q.; Ding, W.J.; Deng, T.; Wu, J.B.; Zhou, J.X. Visualization of fast “hydrogen pump” in core–shell nanostructure Mg@Pt through hydrogen-stabilized Mg3Pt. J. Mater. Chem. A 2019, 7, 14629–14637. [Google Scholar] [CrossRef]
  32. Zeng, L.; Wei, X.F.; Liang, M.K.; Zhao, J.; Zhu, B.C. Probing on the Stable Structure of Silicon-Doped Charged Magnesium Nanomaterial Sensor: SiMgn±1 (n = 2−12) Clusters DFT Study. Front. Mater. 2020, 7, 221. [Google Scholar] [CrossRef]
  33. Zeng, L.; Liang, M.K.; Wei, X.F.; Guo, J.; Zhang, S.; Bi, J.; Dai, W.; Zhu, B.C. Probing the structural evolution, electronic and spectral properties of beryllium doped magnesium and their ions clusters. New J. Chem. 2020, 44, 16929–16940. [Google Scholar] [CrossRef]
  34. Zhao, Y.R.; Xu, Y.Q.; Chen, P.; Yuan, Y.Q.; Qian, Y.; Li, Q. Structural and electronic properties of medium-sized beryllium doped magnesium BeMgn clusters and their anions. Results Phys. 2021, 26, 104341. [Google Scholar] [CrossRef]
  35. Zhang, F.G.; Zhang, H.R.; Xin, W.; Chen, P.; Hu, Y.F.; Zhang, X.Y.; Zhao, Y.R. Probing the structural evolution and electronic properties of divalent metal Be2Mgn clusters from small to medium-size. Sci. Rep. 2020, 10, 6052. [Google Scholar] [CrossRef] [Green Version]
  36. Li, Q.Y.; Xi, S.G.; Hu, Y.F.; Yuan, Y.Q.; Zhao, Y.R.; Li, M.C.; Yuan, J.J.; Yang, Y.J. Probing the structural and electronic properties of neutral and anionic strontium-doped magnesium clusters. Comput. Mater. Sci. 2021, 197, 110605. [Google Scholar] [CrossRef]
  37. Li, Q.Y.; Hu, Y.F.; Xi, S.G.; Li, Y.Y.; Yang, H.; Yuan, Y.Q.; Yang, J.; Li, M.C. Structural and electronic properties of neutral and anionic magnesium clusters doped with two barium atoms. J. Mol. Liq. 2021, 343, 117622. [Google Scholar] [CrossRef]
  38. Li, C.G.; Cui, Y.Q.; Tian, H.; Shao, Q.Q.; Zhang, J.; Ren, B.Z.; Yuan, Y.Q. Systematic investigation of geometric structures and electronic properties of lithium doped magnesium clusters. Comput. Mater. Sci. 2021, 200, 110800. [Google Scholar] [CrossRef]
  39. Wang, Y.C.; Lv, J.; Zhu, L.; Ma, Y.M. Crystal structure prediction via particle swarm optimization. Phys. Rev. B Condens. Matter Mater. Phys. 2010, 82, 094116. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, Y.C.; Lv, J.; Zhu, L.; Ma, Y.M. CALYPSO: A method for crystal structure prediction. Comput. Phys. Commun. 2012, 183, 2063–2070. [Google Scholar] [CrossRef] [Green Version]
  41. Lv, J.; Wang, Y.C.; Zhu, L.; Ma, Y.M. Particle swarm structure prediction on clusters. J. Chem. Phys. 2012, 137, 084104. [Google Scholar] [CrossRef]
  42. 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]
  43. Li, C.G.; Li, H.J.; Cui, Y.Q.; Tian, H.; Shao, Q.Q.; Zhang, J.; Zhao, G.; Ren, B.Z.; Hu, Y.F. A density functional investigation on the structures, electronic, spectral and fluxional properties of VB20 cluster. J. Mol. Liq. 2021, 339, 116764. [Google Scholar] [CrossRef]
  44. Zhao, Y.R.; Bai, T.T.; Jia, L.N.; Xin, W.; Hu, Y.H.; Zheng, X.S. Probing the Structural and Electronic Properties of Neutral and Anionic Lanthanum-Doped Silicon Clusters. J. Phys. Chem. C 2019, 123, 28561–28568. [Google Scholar] [CrossRef]
  45. Lai, L.; Die, D.; Zheng, B.X.; Du, Q. Growth mechanism and electronic and magnetic properties of AgnTi alloy clusters. J. Phys. Chem. Solids 2021, 148, 109757. [Google Scholar] [CrossRef]
  46. Jin, S.Y.; Chen, B.L.; Kuang, X.Y.; Lu, C.; Sun, W.G.; Xia, X.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]
  47. Li, C.G.; Shen, Z.G.; Zhang, J.; Cui, Y.Q.; Li, J.J.; Xue, H.Y.; Li, H.F.; Ren, B.Z.; Hu, Y.F. Analysis of the structures, stabilities and electronic properties of MB16 (M = V, Cr, Mn, Fe, Co, Ni) clusters and assemblies. New J. Chem. 2020, 44, 5109–5119. [Google Scholar] [CrossRef]
  48. Lu, C.; Gong, W.G.; Li, Q.; Chen, C.F. Elucidating stress-strain of ZrB12 from first-priciples studies. J. Phys. Chem. Lett. 2020, 11, 9165–9170. [Google Scholar] [CrossRef] [PubMed]
  49. Die, D.; Zheng, B.X.; Yue, J.Y.; Guo, J.J.; Du, Q. The ground-state structure, optical-absorption and photoelectron spectrum of silver clusters. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 117, 113805. [Google Scholar] [CrossRef]
  50. Li, C.G.; Gao, J.H.; Zhang, J.; Song, W.T.; Liu, S.Q.; Gao, S.Z.; Ren, B.Z.; Hu, Y.F. Structures, stabilities and electronic properties of boron-doped silicon clusters B3Sin (n = 1–17) and their anions. Mol. Phys. 2019, 117, 382. [Google Scholar] [CrossRef]
  51. Krishnan, R. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  52. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401–146405. [Google Scholar] [CrossRef] [Green Version]
  53. 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. Fox, Gaussian 09 Revision D.01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  54. Lu, T.; Chen, F.W. Multiwfn: A multifunctional wavefunction analyzer. Comput. Phys. Commun. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  55. Ruette, F.; Sanchez, M.; Anez, R.; Bermudez, A.; Sierraalta, A. Diatomic molecule data for parametric methods. I. J. Mol. Struct. THEOCHEM 2005, 729, 19–37. [Google Scholar] [CrossRef]
  56. Soltani, A.; Boudjahem, A.G.; Bettahar, M. Electronic and magnetic properties of small RhnCa (n = 1–9) clusters: A DFT Study. Int. J. Quantum Chem. 2016, 116, 346–356. [Google Scholar] [CrossRef]
  57. Zhou, G.D.; Duan, L.Y. Free Movement in the EU: The Case of Great Britain Structural Chemistry Basis; Peking University Press: Beijing, China, 2002; pp. 1–10. [Google Scholar]
Figure 1. Optimized geometrical structures of lowest and low-lying isomers of Ca2Mgn (n = 1–15) clusters and Mgn+2 (n = 1–15) clusters along with the point group symmetry, electronic states and relative energy (eV). The pinkish and orange balls are magnesium and calcium atoms, respectively.
Figure 1. Optimized geometrical structures of lowest and low-lying isomers of Ca2Mgn (n = 1–15) clusters and Mgn+2 (n = 1–15) clusters along with the point group symmetry, electronic states and relative energy (eV). The pinkish and orange balls are magnesium and calcium atoms, respectively.
Nanomaterials 12 01654 g001
Figure 2. (A) Eb, (B) Δ2E, (C) Eg and (D) the total charges on Ca atoms in the ground state of Ca2Mgn (n = 1–15) clusters.
Figure 2. (A) Eb, (B) Δ2E, (C) Eg and (D) the total charges on Ca atoms in the ground state of Ca2Mgn (n = 1–15) clusters.
Nanomaterials 12 01654 g002
Figure 3. The Infrared and Raman spectra of the most stable cluster of the Ca2Mg8 cluster.
Figure 3. The Infrared and Raman spectra of the most stable cluster of the Ca2Mg8 cluster.
Nanomaterials 12 01654 g003
Figure 4. The calculated total densities of states (TDOS) and partial densities of states (PDOS) of Ca2Mgn (n = 1–15) clusters.
Figure 4. The calculated total densities of states (TDOS) and partial densities of states (PDOS) of Ca2Mgn (n = 1–15) clusters.
Nanomaterials 12 01654 g004
Figure 5. Molecular orbitals and the corresponding energy levels of the Ca2Mg8 cluster. The HOMO–LUMO gap is indicated (in azure).
Figure 5. Molecular orbitals and the corresponding energy levels of the Ca2Mg8 cluster. The HOMO–LUMO gap is indicated (in azure).
Nanomaterials 12 01654 g005
Figure 6. AdNDP chemical bonds and the corresponding structural units for the Ca2Mg8 cluster. (ON denotes the occupation number).
Figure 6. AdNDP chemical bonds and the corresponding structural units for the Ca2Mg8 cluster. (ON denotes the occupation number).
Nanomaterials 12 01654 g006
Table 1. Electronic states, symmetries, average binding energies Eb (eV), HOMO–LUMO energy gaps Eg (eV), the second-order difference energy (∆2E) and charges on the Ca atoms of the most stable Ca2Mgn (n = 1–15) clusters.
Table 1. Electronic states, symmetries, average binding energies Eb (eV), HOMO–LUMO energy gaps Eg (eV), the second-order difference energy (∆2E) and charges on the Ca atoms of the most stable Ca2Mgn (n = 1–15) clusters.
ClustersStateSym.Eb (eV)Δ2E (eV)Eg (eV)Charge (e)
Ca1Ca2
Ca2Mg1A1C2V0.20-1.910.110.11
Ca2Mg21A1C2V0.400.572.010.310.31
Ca2Mg31A1C2V0.40−0.211.850.370.37
Ca2Mg41A1C2V0.44−0.251.710.410.41
Ca2Mg51A1D5H0.500.331.390.530.53
Ca2Mg61AD5H0.51−0.371.510.370.37
Ca2Mg71A1C2V0.56−0.191.310.680.68
Ca2Mg81A’CS0.610.661.770.840.94
Ca2Mg91AC20.60−0.201.770.710.71
Ca2Mg101AC10.60−0.191.160.860.71
Ca2Mg111ACS0.620.141.130.870.71
Ca2Mg121A’CS0.63−0.421.120.850.85
Ca2Mg131A’CS0.660.061.190.860.86
Ca2Mg141A’CS0.69−0.060.931.061.06
Ca2Mg151A1C3V0.71-1.41−0.180.85
Table 2. The HOMO and LUMO energy of Ca2Mgn and Mgn+2 (n = 1–15).
Table 2. The HOMO and LUMO energy of Ca2Mgn and Mgn+2 (n = 1–15).
ClustersHOMO (eV)LUMO (eV)ClustersHOMO (eV)LUMO (eV)
Ca2Mg−3.834−1.920Mg3−4.799−1.937
Ca2Mg2−3.981−1.973Mg4−4.961−2.076
Ca2Mg3−3.831−1.979Mg5−4.258−2.099
Ca2Mg4−3.860−2.148Mg6−4.294−2.311
Ca2Mg5−3.531−2.140Mg7−4.436−2.347
Ca2Mg6−3.897−2.384Mg8−4.284−2.507
Ca2Mg7−3.917−2.611Mg9−4.393−2.868
Ca2Mg8−3.954−2.188Mg10−4.306−2.350
Ca2Mg9−4.048−2.275Mg11−4.385−2.338
Ca2Mg10−3.750−2.588Mg12−4.012−2.583
Ca2Mg11−3.671−2.543Mg13−4.105−2.652
Ca2Mg12−3.746−2.623Mg14−4.092−2.881
Ca2Mg13−3.730−2.539Mg15−3.973−2.750
Ca2Mg14−3.518−2.592Mg16−3.985−2.521
Ca2Mg15−3.965−2.558Mg17−3.917−2.823
Table 3. The Wiberg bond orders of the Ca2Mg8 cluster.
Table 3. The Wiberg bond orders of the Ca2Mg8 cluster.
AtomsCa-1Ca-2Mg-3Mg-4Mg-5Mg-6Mg-7Mg-8Mg-9
Ca-20.130
Mg-30.3830.462
Mg-40.1510.1430.233
Mg-50.3830.4620.5240.233
Mg-60.1390.1310.5170.5290.228
Mg-70.3980.1490.5280.4390.2370.523
Mg-80.1390.1310.2280.5290.5170.1820.199
Mg-90.1790.4100.4650.4580.4650.4870.2310.487
Mg-100.3980.1490.2370.4390.5280.1990.5230.5230.231
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, C.; Cui, Y.; Tian, H.; Ren, B.; Li, Q.; Li, Y.; Yang, H. Quantum Chemistry Study on the Structures and Electronic Properties of Bimetallic Ca2-Doped Magnesium Ca2Mgn (n = 1–15) Clusters. Nanomaterials 2022, 12, 1654. https://doi.org/10.3390/nano12101654

AMA Style

Li C, Cui Y, Tian H, Ren B, Li Q, Li Y, Yang H. Quantum Chemistry Study on the Structures and Electronic Properties of Bimetallic Ca2-Doped Magnesium Ca2Mgn (n = 1–15) Clusters. Nanomaterials. 2022; 12(10):1654. https://doi.org/10.3390/nano12101654

Chicago/Turabian Style

Li, Chenggang, Yingqi Cui, Hao Tian, Baozeng Ren, Qingyang Li, Yuanyuan Li, and Hang Yang. 2022. "Quantum Chemistry Study on the Structures and Electronic Properties of Bimetallic Ca2-Doped Magnesium Ca2Mgn (n = 1–15) Clusters" Nanomaterials 12, no. 10: 1654. https://doi.org/10.3390/nano12101654

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop