Next Article in Journal
Evaluation of Ruthenium(II) N-Heterocyclic Carbene Complexes as Enzymatic Inhibitory Agents with Antioxidant, Antimicrobial, Antiparasitical and Antiproliferative Activity
Next Article in Special Issue
Adsorption Features of Tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 Borophene and Pristine Graphene Nanosheets: A Comparative DFT Study
Previous Article in Journal
Dual-Enzyme Cascade Composed of Chitosan Coated FeS2 Nanozyme and Glucose Oxidase for Sensitive Glucose Detection
Previous Article in Special Issue
A Novel Two-Dimensional ZnSiP2 Monolayer as an Anode Material for K-Ion Batteries and NO2 Gas Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structures and Stabilities of Carbon Chain Clusters Influenced by Atomic Antimony

1
School of Pharmaceutical and Chemical Engineering, Taizhou University, Taizhou 318000, China
2
Department of Chemistry, Tongji University, Shanghai 200092, China
3
School of Intelligent Engineering, Shaoguan University, Shaoguan 512005, China
4
Department of Physics and Astronomy, London Centre for Nanotechnology, University College London, London WC1E 6BT, UK
5
Research Institute of Zhejiang University-Taizhou, Zhejiang University, Taizhou 318000, China
6
Key Laboratory of Yunnan Provincial Higher Education Institutions for Organic Optoelectronic Materials and Devices, Kunming University, Kunming 650214, China
7
Yunnan Key Laboratory of Metal-Organic Molecular Materials and Devices, Kunming University, Kunming 650091, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(3), 1358; https://doi.org/10.3390/molecules28031358
Submission received: 7 December 2022 / Revised: 18 January 2023 / Accepted: 21 January 2023 / Published: 31 January 2023
(This article belongs to the Special Issue Computational Chemistry for Material Research)

Abstract

:
The C-C bond lengths of the linear magnetic neutral CnSb, CnSb+ cations and CnSb anions are within 1.255–1.336 Å, which is typical for cumulene structures with moderately strong double-bonds. In this report, we found that the adiabatic ionization energy (IE) of CnSb decreased with n. When comparing the IE~n relationship of CnSb with that of pure Cn, we found that the latter exhibited a stair-step pattern (n ≥ 6), but the IE~n relationship of CnSb chains took the shape of a flat curve. The IEs of CnSb were lower than those of corresponding pure carbon chains. Different from pure carbon chains, the adiabatic electron affinity of CnSb does not exhibit a parity effect. There is an even-odd alternation for the incremental binding energies of the open chain CnSb (for n = 1–16) and CnSb+ (n = 1–10, when n > 10, the incremental binding energies of odd (n) chain of CnSb+ are larger than adjacent clusters). The difference in the incremental binding energies between the even and odd chains of both CnSb and pure Cn diminishes with the increase in n. The incremental binding energies for CnSb anions do not exhibit a parity effect. For carbon chain clusters, the most favorable binding site of atomic antimony is the terminal carbon of the carbon cluster because the terminal carbon with a large spin density bonds in an unsaturated way. The C-Sb bond is a double bond with Wiberg bond index (WBI) between 1.41 and 2.13, which is obviously stronger for a carbon chain cluster with odd-number carbon atoms. The WBI of all C-C bonds was determined to be between 1.63 and 2.01, indicating the cumulene character of the carbon chain. Generally, the alteration of WBI and, in particular, the carbon chain cluster is consistent with the bond length alteration. However, the shorter C-C distance did not indicate a larger WBI. Rather than relying on the empirical comparison of bond distance, the WBI is a meaningful quantitative indicator for predicting the bonding strength in the carbon chain.

1. Introduction

Small magnetic carbon clusters and related carbon-based materials have attracted much attention on the account of their important role in astrophysics, terrestrial processes, electronic spintronics, catalysis, and chemical engineering [1,2,3,4,5]. A deep understanding of the carbon nanostructure with adjustable bonding should facilitate the designing and synthesizing of size- and morphology-controlled carbon-based functional materials [6,7,8,9]. Previous theoretical and experimental studies have revealed that small carbon clusters are mainly of a linear structure [10,11,12]. The pure carbon clusters are ascertained to be intermediate in the production of diamonds, silicon carbine films, and a variety of chemical systems involving hydrocarbons [13,14,15]. Transition-metal doping for hybridization is an effective way to adjust the performance and electronic properties of nanomaterials [16,17,18,19,20]. In the interstellar medium, the reactivity of small carbon clusters is forfeited by quasi-collisionless conditions, and carbon atoms take the highly thermally stable, albeit highly reactive, form linear chains. When these metastable carbon chains encounter heteroatoms such as sulfur, oxygen, hydrogen, and nitrogen, they form more stable complexes.
Mass spectrometry is useful for detecting pure and doped carbon clusters with high stabilities. Theoretical studies of the linear carbon clusters show that doped by different heteroatom X, CnX/CnX+/CnX clusters have different parity effects in their stabilities [21,22,23]. Zhan and Iwata have used more sophisticated post-HF methods, including Møller-Plesset2 (MP2), MP4SDTQ, and the QCISD(T) method with different basis set to study the properties of CnN- and CnP [24,25]. Zheng et al. produced cluster anions CnN and CnP from the laser ablation of appropriate samples and studied them by TOF mass spectrometry [26,27]. Small, doped carbon clusters have been extensively investigated theoretically and experimentally. Theoretical methods have been useful for studying heavier-atom-doped carbon clusters. Recently, theoretical work for lead [28], gallium [29], indium [30], gold [31], and tantalum [32] doped carbon clusters has been conducted. With the electronegativity modulation of heteroatom X, the parity effect in CnX may undergo dramatic changes. Their geometries, electronic structures, and bonding of small metal carbide clusters, such as antimony-doped carbon clusters, are yet to be determined systematically. In this paper, we report a density-functional-theory (DFT) study of the linear CnSb, CnSb+, and CnSb clusters.

2. Results and Discussion

2.1. Structural Optimization

Table 1, Table 2 and Table 3 present the optimized bond lengths obtained at the B3LYP/6-31G(d) level, while Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6 analyzed the C-C and C-Sb bond lengths. Table 4, Table 5 and Table 6 summarize the calculated total energies En of optimized structures, differential energies ΔEn, binding energies BE, incremental binding energies ΔEI, dipole moments, and rotational constants. As shown in Figure 1, we can find that the C-C bond lengths of neutral CnSb are within the range of 1.270–1.304 Å, exhibiting the character of both double-bonds (C-C bond length of ethylene is computed to be 1.331 nm by using B3LYP/6-31G(d)) and triple-bonds (C-C bond length of acetylene 1.205 Å by B3LYP/6-31G(d)); The C-Sb bond length of neutral CnSb is within the range of 1.941 Å to 1.968 Å. The C-C bond length alteration (BLA) of neutral CnSb is obvious and tends to be irregular in the vicinity of Sb. For CnSb+ clusters, the C-C bond lengths are within a range from 1.259 Å to 1.334 Å, C-Sb bond lengths are in a range from 1.889 Å to 2.009 Å. The optimized BLA of corresponding C-C and C-Sb showed diverse tendencies compared with that of neutral CnSb. The odd CnSb+ chains also have a significant BLA effect; instead of being weakened and irregular C-C bond length, the C-C BLA near the antimony atom is slightly enhanced, which is markedly different from the corresponding neutral ones. The C-C bond lengths of CnSb anions are within a range from 1.255 Å to 1.336 Å, and C-Sb bond lengths are within a range from 1.889 Å to 2.014 Å; the fluctuation trend of the BLA effect in even CnSb anions have the same tendencies with that of neutral CnSb; different from neutral CnSb, odd CnSb anions also have a distinct BLA effect, and this effect is more obvious than CnSb+ cations. Unlike the CnSb+ and CnSb, the C-Sb bond lengths of even CnSb anions are larger than the adjacent odd cluster anions, as depicted in Figure 5. We can preliminarily infer from C-C BLA and C-Sb BLA that the antimony atom of even CnSb+ and odd CnSb might combine with the carbon clusters more firmly than adjacent clusters (this speculation can be proved roughly right for small clusters by our calculation on their dissociation channels). When comparing the BLA of neutral and charged antimony-doped carbon chains with that of pure carbon chains, it is easy to deduce that through doping the antimony element, the properties of carbon clusters have been changed to a certain extent (Figure 1 and Figure 2). The C-C bond lengths are within the range of 1.255 Å–1.336 Å, typical for cumulene structures with moderately strong double-bonds. On the other hand, the clear alternation in C-C distances suggests that there is a substantial contribution of polyacetylenic valence-bond structures with an alternating triple and single C-C bond.
We mainly focused on the linear carbon chain clusters initially because many reported carbon clusters mainly adopt a linear shape as the ground state structure. For the antimony-doped larger carbon clusters (n > 10), we conducted density functional calculations to obtain the energies of ring structures. The total energies of C11Sb and C14Sb with a ring structure are 0.35 and 0.11 eV higher in energy than linear C11Sb and C14Sb clusters, while C12Sb, C13Sb, C15Sb, and C16Sb clusters with a ring structure are −0.14, −0.57, −0.53, and −0.83 eV lower in energy. Although the linear carbon cluster is not always the global minima in the potential surface, the linear carbon clusters possess a practical significance that the reactivity of carbon clusters increases dramatically with the increasing number of consecutive cumulene-like double bonds, and the linear carbon clusters can be formed in space-confined nanotube materials [33,34,35]. The linear carbon chain cluster with different lengths and doping atoms is a special allotrope type of carbon material different from graphite, diamond, graphene, carbon nanotubes, and fullerenes.
We made comparisons between 6-31G(d) and def2-TZVP basis sets for selected structural parameters, C-Sb stretching frequencies, and HOMO-LUMO gaps in Table 7. The C-C bonds adjacent antimony atoms of C4Sb, C5Sb, C6Sb, C10Sb, and C11Sb show only slight differences less than 0.011 Å, and the C-Sb bond distances are almost the same through using def2-TZVP and 6-31G(d) basis sets. C-Sb stretching frequencies tend to show negligible differences for the larger C11Sb and C12Sb. Although the C-Sb stretching frequency differences for C4Sb and C5Sb, it is slightly larger (6.2 cm−1 and 3.9 cm−1) under different basis sets, and we can clearly recognize that the theoretically predicted vibration modes are the same by comparing normal coordinates. The HOMO-LUMO gaps for C4Sb, C5Sb, C6Sb, C10Sb, and C11Sb are 1.27, 1.25, 1.17, 1.03, and 1.01 eV with 6-31G(d), while the HOMO-LUMO gaps are nearly identical with def2-TZVP basis set (1.26, 1.27, 1.16, 1.04, and 0.99 eV, respectively).

2.2. Electronic Configuration

The total symmetry of the pure carbon chain is D h The core electron configuration for carbon is 1s2, and for antimony, 1s22s22p63s23p63d104s24p64d10. CnSb has, therefore, 4n + 5 valence electrons. Based on molecular orbital theory, quantum chemical computations predict the electronic configurations of the linear ground-state CnSb clusters as:
C n Sb { ( core ) 1 σ 2 2 σ 2 1 π 4 3 σ 1 , n = 1 ( core ) 1 σ 2 2 σ 2 3 σ 2 ( n + 2 ) σ 2 ( core ) 1 σ 2 2 σ 2 3 σ 2 ( n + 2 ) σ 2 n π 2 π 1 ,   n   is   even 2 n + 1 2 π 2 ,   n   is   odd
Table 8 lists the valence orbital configuration of CnSb (n = 1–16) clusters. CnSb possesses (2n + 1) valence π-electrons and (2n + 4) valence σ-electrons. The (2n + 4) valence σ-electrons fully occupy the (n + 2) σ-orbitals. CnSb has a π1 electronic configuration, 2Π state for n-even members and a π3 electronic configuration, 2Π state for n-odd members. CnSb+ anions contain 2n valence π-electrons; even for n, these 2n π-electrons should fully populate n and doubly degenerate π-orbitals, resulting in a …π4 electronic configuration and a 1Σ+ state while for odd n, the highest occupied molecular orbital (HOMO) with π-symmetry is half-filled with two electrons, resulting in a π2 electronic configuration and a 3Σ state. For CnSb anions, the situation is just the opposite. Two more π-electrons result in fully filled π-orbitals in n-odd CnSb clusters and a half-filled π-orbital in n-odd CnSb clusters.
The frontier molecular orbitals are depicted in Figure 7 with an isovalue of 0.02. Except for the CSb molecule, the LUMO of the other carbon chain clusters exhibits the same orbital shape as HOMO. The smallest CSb shows the σ-character of HOMO, while all the other CnSb chain clusters show π-character HOMO with overlapping p orbitals shoulder to shoulder. For HOMO, the p orbital of the terminal carbon of the even-numbered carbon cluster always presents a different sign from the p orbital of antimony. While the p orbital of the terminal carbon of odd-numbered carbon cluster exhibits the same sign with a p orbital of antimony, it can form π bonding orbital with antimony. Therefore, the C-Sb bond of CnSb with even-numbered carbon should be weaker than the C-Sb bond of Cn ± 1Sb with odd-numbered carbon. This result agrees well with the Wiberg bond index analysis that the C-Sb WBI of CnSb with an even n tends to be smaller than the C-Sb WBI of CnSb with odd n.

2.3. Electronic Properties

As is known, the ion signal intensity in a mass spectrum is related to the electron affinity (EA) or ionization energies (IE, also called electron detachment energy) of the molecule. Thus, adiabatic ionization energies, defined as the energy required to remove an electron from the neutral clusters with a geometric change, are important parameters to understand the relative stability of the antimony-doped clusters with different sizes. Usually, there are three types of IE: Koopmans IE, vertical IE, and adiabatic IE. Koopmans IE is the HOMO energy, vertical IE is the energy difference between the neutral and ionic clusters at the neutral equilibrium geometry, and adiabatic IE is the energy difference between the neutral and ionic clusters at their respective equilibrium geometry (i.e., IE = E(optimized cation) − E(optimized neutral)). In this work, the adiabatic IE of CnSb and Cn clusters for their optimized structures were calculated and shown in Table 4 and Figure 8. As a whole, the ionization energies decrease with the size of the clusters, suggesting that larger CnSb chains become less stable, e.g., when exposed to a strong electrical field or high temperature. As shown in Figure 8, the adiabatic IE of pure carbon chains Cn has a stair-step shape (n ≥ 6), whereas CnSb chains take the shape of a flat curve with a smaller gradient than pure carbon chains. The adiabatic IEs of the even and odd pure carbon chains (n ≥ 6) follow some nonlinear relationship; the IEs of carbon chains are larger than that of corresponding CnSb, but the energy difference between them (IECn-IECnSb) decreases when n increases. The electron affinity (EA) of CnSb is defined as the energy released when an electron is attached to neutral CnSb:
EA = E(optimized neutral) – E(optimized anion).
This property is also related to the stability of the molecule in the TOF experiment and molecule. A higher EA means that more energy is released when an electron is added to a neutral molecule, and the generation of the corresponding anion is more readily performed. The calculated adiabatic EA data are presented in Table 4. In Figure 9, we compared the adiabatic EAs of CnSb with that of pure carbon chains: the EAs of carbon chains have a clear alternation parity effect, the EAs of carbon chains with an even n are larger than the EAs of adjacent n-odd members but carbon chains doped with the antimony atom do not exhibit a parity effect. The computed EAs of the even and odd carbon chains followed a non-linear relationship with the number of carbons, respectively. The energy difference between EACn and EACnSb was much smaller than (IECn - IECnSb), except when n = 1, and EACn was slightly larger than EACnSb when n ≥ 10.
All the antimony atoms of CnSb show positive charges, indicating that antimony denotes electrons to the carbon cluster. The antimony of CSb carries the least positive charge of 0.33 |e|, while the antimony of C2Sb carries the most positive charge of 0.52 |e|. Antimony atoms of larger CnSb (n > 3) chain clusters show positive charges 0.43, 0.44, 0.49, 0.44, 0.48, 0.44, 0.46, 0.44, 0.46, 0.44, 0.45, 0.44, and 0.44 |e|, respectively. Thus, the carbon chain with an even-number of carbon atoms tends to exhibit higher antimony charges. The carbon atom nearest to antimony showed a negative charge due to the electron donation of antimony (Table 9). Without doping antimony, the two anomeric carbons at the right and the left end of the pure carbon chain show an identical charge value. With the antimony doping, the anomeric carbon shows significant charge differences (Table 9 and Figure 10). The charge differences between anomeric carbon atoms are calculated to be 0.43, 0.31, 0.51, 0.53, 0.62, 0.6, 0.65, 0.63, 0.66, 0.64, 0.67, 0.65, 0.67, 0.65, and 0.67 |e| for CnSb (n = 2–16). This result suggests that the doping of antimony definitely changed the charge population of the carbon cluster.
The Wiberg bond indices of CnSb are listed in Figure 11. The C-Sb bond is a double bond with WBI between 1.41 and 2.13, which is obviously stronger for a carbon chain cluster with odd-number carbon atoms (compared with neighboring carbon chain clusters with even-number carbon atoms). The WBI of all C-C bonds was determined to be between 1.63 and 2.01, indicating the cumulene character of the carbon chain. Generally, the alteration of WBI and, in particular, the carbon chain cluster is consistent with the bond length alteration. However, the shorter C-C distance did not indicate a larger WBI. For example, the largest WBI for CnSb (n > 1) was calculated to be the terminal C-C bond index, while this terminal C-C bond was not the shortest C-C bond. Therefore, rather than relying on the empirical comparison of bond distance, the WBI is a meaningful quantitative indicator for predicting the bonding strength in the carbon chain.

2.4. Incremental Energies, Fragmental Energies, and Binding Sites

The relative stability of clusters can be also analyzed in terms of the energy differences between the neighboring size of the clusters, which is correlated with the “magic number” in cluster science [36]. Energy differences between CnSb and Cn−1Sb, CnSb+ and Cn−1Sb+, CnSb and Cn−1Sb (differential energies ΔEn, defined as ΔEn = EnEn−1) are listed in Table 5, Table 6 and Table 7, respectively. For the clusters with different sizes, the concept of the incremental binding energy, labeled as ΔEI, was introduced to compare their relative stabilities. As suggested by Pascoli and Lavendy [37,38], ΔEI is just the reaction energy of the following processes:
CnSb → Cn−1Sb + C (DN1)
CnSb+ → Cn−1Sb+ + C (DC1)
CnSb → Cn−1Sb + C (DA1)
They can be computed as the consecutive binding energy (BE, atomization energy, listed in Table 5, Table 6 and Table 7) differences between the adjacent clusters,
ΔEI(CnSb/CnSb+/ CnSb) = BE(CnSb/CnSb+/CnSb) − BE(Cn−1Sb/Cn−1Sb+/Cn−1Sb)
where BE can be defined as the energy difference between a molecule and its component atoms:
BE(CnSb/ CnSb+/ CnSb) = nE(C) + E(Sb) − E(CnSb/ CnSb+/ CnSb)
The results for the incremental binding energy as a function of the number of carbon atoms for the different open-chain CnSb/CnSb+/CnSb clusters and pure carbon chains are shown in Figure 12, Figure 13 and Figure 14. It can be observed that there is an even-odd alternation for the open chain CnSb and CnSb+ (n = 1–10, when n > 10, the parity effect is less obvious and n-odd members of CnSb+ are slightly larger), with even species being comparatively more stable than odd ones; the parity variation tendency of neutral CnSb is opposite to that of pure Cn (i.e., n-even carbon chains are less stable than adjacent n-odd ones) and the variation amplitude is much smaller than pure Cn. The difference in ΔEI between the even and odd species of both CnSb and pure Cn chains diminishes with the increase in carbon atoms; different from Cn- anions, the ΔEI for CnSb anions did not exhibit a parity effect. However, these patterns of ΔEI for CnSb+ and CnSb cannot be simply explained by the “valence π-electron number” rule.
The fragmentation energies accompanying channels DN1, DC1, and DA1, i.e., △EI, have been discussed. In addition, the fragmentation energies for many other dissociation reactions are calculated and exhibited in Figure 15, Figure 16 and Figure 17, including the following seven channels for neutral CnSb clusters:
CnSb → Cn−2Sb + C2(DN2)
CnSb → Cn−3Sb + C3 (DN3)
CnSb → Cn + Sb (DN4)
CnSb → Cn−1 + CSb (DN5)
CnSb → Cn−2 + C2Sb (DN6)
CnSb → Cn−3 + C3Sb (DN7)
CnSb → Cn- + Sb+ (DN8)
The following ten channels for cationic CnSb+ cations:
CnSb+ → Cn−2Sb+ + C2 (DC2)
CnSb+ → Cn−3Sb+ + C3 (DC3)
CnSb+ → Cn+ + Sb (DC4)
CnSb+ → Cn−1+ + CSb (DC5)
CnSb+ → Cn−2+ + C2Sb (DC6)
CnSb+ → Cn−3+ + C3Sb (DC7)
CnSb+ → Cn + Sb+ (DC8)
CnSb+ → Cn−1 + CSb+ (DC9)
CnSb+ → Cn−2 + C2Sb+ (DC10)
CnSb+ → Cn−3 + C3Sb+ (DC11)
and the following thirteen channels for anionic CnSb anions:
CnSb → Cn−2Sb + C2 (DA2)
CnSb → Cn−3Sb + C3 (DA3)
CnSb → Cn−1Sb + C (DA4)
CnSb → Cn−2Sb + C2 (DA5)
CnSb → Cn−3Sb + C3 (DA6)
CnSb → Cn + Sb (DA7)
CnSb → Cn−1+ CSb (DA8)
CnSb → Cn−2+ C2Sb (DA9)
CnSb → Cn−3+ C3Sb (DA10)
CnSb → Cn + Sb (DA11)
CnSb → Cn−1 + CSb (DA12)
CnSb → Cn−2 + C2Sb (DA13)
CnSb → Cn−3 + C3Sb (DA14)
These channels can be divided into four types: (1) losing neutral small carbon particles such as C, C2, or C3; (2) losing neutral antimony-contained small fragments such as Sb, CSb, C2Sb, or C3Sb; (3) losing ionic defects, such as Sb+/Sb, CSb+/CSb, C2Sb+/C2Sb, or C3Sb+/C3Sb fragments (for CnSb+/CnSb); and (4) losing anionic carbons, such as C, C2, and C3 fragments (only for CnSb). Comparing the fragmentation energies can help us to find some dominant channels for each kind of cluster in the discussion.
The fragmentation energies of channels DN1, DC1, and DA1 are also included for comparison. It is clear that losing an antimony atom is the dominant channel for neutral CnSb (channel DN4). For small CnSb+ cations (n = 1–9), the most favorable fragmentation channel is the loss of the Sb+ ion (channel DC8). However, when n ≥ 10, losing an antimony atom (DC4) becomes the dominant channel. The most favorable dissociation pathway for CSb and CnSb (n = 2–16) is the loss of the Sb ion and the loss of the antimony atom, respectively. The most favorable dissociation channels for CnSb/CnSb+/CnSb are illustrated in Figure 18, from which we can draw some conclusions: the fragmentation energies for DN4 exhibit a parity effect, i.e., the even CnSb clusters are more stable while the odd CnSb clusters are easy to dissociate; the fragmentation energies for DC8 (n = 1–9) also showed an alternation effect with the same alternation trend as DN4. the fragmentation energies for DC4 did not show fluctuation or a decrease monotonically as the n number rose; when n > 4 CnSb+, is more stable than CnSb anions and neutral CnSb. The subtle alternation for DA7 exists when n ≤ 10; when n > 10, the CnSb anion is in its least stable form compared with CnSb and CnSb+.
For carbon chain clusters, the most favorable binding site of atomic antimony is the terminal carbon of the carbon cluster because the terminal carbon with a large spin density bonds in an unsaturated way. We have constructed initial structures with antimony binding on nonterminal carbon and found that the optimization of these structures often encounters a convergence problem or loop back to the linear structure with antimony binding to the terminal carbon. For the structures which encounter convergence problems or converge to geometry with antimony binding on nonterminal carbon, we further optimized them and obtained the local minima (Figure 19). For antimony binding on C2, the 2a nonlinear structure showed a C-C distance of 1.315 Å, longer than the linear structure. The results indicate that the construction of the nonlinear shape CnSb can further activate the C-C bond. Nonlinear C3Sb and C4Sb show C2V symmetric structures. The C-C bonds of C3Sb are all 1.335 Å, while the C-C bond adjacent to antimony shows a substantially longer distance of 1.380 Å. Nonlinear C5Sb with antimony binding on the second or third carbon does not exhibit bond length alteration. Nonlinear C6Sb and C7Sb with antimony binding on the third carbon or binding on the bridge site between the third and fourth carbon do not exhibit bond length alteration effects. The C-C bond distances adjacent to antimony are calculated to be 1.346 and 1.360 Å for 5a and 5b, respectively, which are much longer than 1.288 Å for linear C5Sb. The C-C bond distances adjacent to the antimony of 6a and 6b are calculated to be 1.355 and 1.440 Å, which is much longer than 1.277 Å for the linear carbon chain C6Sb. The bond length alteration does not exist for nonlinear C7Sb. The antimony binding on the second or third carbon of C7 leads to the reconstruction and formation of the linear carbon chain. The bond length alteration effect exists in unstable nonlinear C8Sb (8c) with s high RBE of 3.93 eV. Structures 8a and 8b do not show the bond length alteration effect.
For the low energy minia of nonlinear C2Sb to C8Sb, the adiabatic ionization energies were calculated to be 8.54, 7.19, 9.09, 6.14, 4.82, 8.06, and 8.26 eV, respectively, indicating that the nonlinear C5Sb and C6Sb could be more easily ionized. Comparatively, the linear carbon chains CnSb show more constant adiabatic ionization energy with the increase in the carbon number. The relative binding energies of nonlinear C2Sb to C8Sb with antimony binding on the sides of carbon clusters (relative to the binding energy of the linear chain cluster) are calculated to be 0.4, 1.23, 1.53, 1.8, 2.86, 2.54, and 2.82 eV, respectively. The results indicate that the relative binding energy is significantly large except for the much smaller carbon cluster C2Sb. Quantitatively, the binding energy of nonlinear C2Sb to C8Sb with antimony binding on the sides of carbon clusters were calculated to be −3.33, −3.24, −1.85, −2.64, −0.34, −1.52, and −0.28 eV, respectively. Antimony is a metallic element exhibiting low electron affinity. If the antimony atom is changed to nitrogen with strong electronegativity, we can test the structural geometry, adiabatic ionization energy, and natural charge population of C6N and C7N. The bond length alteration effect of C6N and C7N is much different from that in C6Sb and C7Sb. The bond length difference between the adjacent C-C bonds is larger than that in the antimony-doped carbon chain cluster. The C-C bond adjacent to nitrogen is substantially longer than that adjacent to antimony. The adiabatic ionization energies of C6N and C7N were calculated to be 8.82 and 9.17 eV, which are larger than the C6Sb and C7Sb, indicating the higher stability of the nitrogen-doped carbon chain. Different from the charging state of antimony, the natural charge population analysis of C6N and C7N indicates that nitrogen atoms are both negatively charged with −0.42 |e|.

3. Computational Methods

To explore the structure and energetics in the linear antimony-doped carbon clusters, full geometry optimizations were performed using density-functional theory methods implemented in the Gaussian 03 program [39]. The Molecular mechanic’s algorithm is frequently employed to investigate very large carbon-based materials because of the efficiency in predicting the binding and delivery mechanism. Density-functional calculations are verified to be effective and have an accurate strategy to reveal the geometric structures [40,41] and electronic properties of various nanomaterials [42,43,44,45]. The B3LYP exchange-correlation function consists of Fock’s exact exchange and Beck’s three-parameter nonlocal exchange function, along with the nonlocal correlation function developed by Lee et al. [46]. B3LYP was chosen here because the previous research suggests that hybrid-functional is reliable and highly efficient for molecules and clusters [47,48], while the post-HF Ab initio method is accurate but time-consuming [49,50]. A medium-size basis set 6-31G(d) was used for a carbon atom, and the Stuttgart/Bonn relativistic effective core potential (SDD) basis set was used for antimony. The geometries and relative energies of heteroatom-doped carbon clusters obtained with the B3LYP method were very close to those with the coupled cluster single and double (triple) (CCSD(T)) and QCISD(T) method [51,52]. Vibrational frequencies were computed at the same level using a harmonic approximation to assess the nature of the optimized structures. Zero-point energies (ZPE) were evaluated as well using the same methodology. The optimized structures were then used for single-point calculations at the B3LYP/6-311++G(3df,3pd) level. In the computation, bond lengths of the magnetic neutral CnSb, CnSb+ cations, and CnSb anions (n = 1–16) clusters have been optimized through the use of B3LYP methods with a 6-31G(d) basis set. Subsequently, the corresponding harmonic vibrational frequencies are evaluated at the same level. To further verify the reliability of the optimized geometries, we have carefully checked every computation step that might cause possible numerical calculation errors. We have adopted the default convergence criteria for self-consistent-field (SCF) calculation, i.e., 10−8 for the root mean square density and 10−6 for the maximum density, and for geometry optimization, 0.00030 This includes the Hartree/Bohr radius for the root mean square force and 0.00045 Hartree/Bohr radius for the maximum force. In addition, the convergence criteria for the energy change in the final step of geometry optimization was set to be 10−7 Hartree.

4. Conclusions

The linear carbon chain cluster with different lengths and doping atoms is a special allotrope type of carbon material different from graphite, diamond, graphene, carbon nanotubes, and fullerenes. We have conducted a systematic DFT study on linear CnSb/CnSb+/CnSb clusters with sizes of n = 1–16 and compared these with pure Cn clusters. C-C bond lengths of the linear neutral CnSb, CnSb+ cations and CnSb anions are within 0.1255−0.1336 nm, which is typical of cumulene structures with moderately strong double bonds. However, the alternation in C-C distances suggests that there is a substantial contribution of polyacetylenic valence-bond structures with an alternating triple and single C-C bond. The C-C BLA of neutral CnSb is obvious for n-even clusters, and for n-odd clusters, the BLA tends to be irregular in the vicinity of Sb. We can deduct from C-C BLA and C-Sb BLA that the antimony atom of n-odd CnSb anions and n-even CnSb+ (n = 1–10) cations are combined more firmly than adjacent clusters. This is roughly proved to be right by our calculation of their dissociation channels. When comparing the BLA of neutral and charged antimony-doped carbon chains to that of pure carbon chains, it is easy to deduce that through doping the antimony element, the properties of carbon clusters are changed. The adiabatic IE of CnSb decreased with the rise in the n number, suggesting that larger CnSb chains become less stable. When comparing the IE of CnSb with that of pure Cn, we found that the latter had a stair-step pattern (n ≥ 6), but CnSb chains took the shape of a flat curve. The IEs of carbon chains are larger than that of corresponding CnSb, but the energy difference (IECn-IECnSb) decreases with increasing n. Different from pure carbon chains, the adiabatic electron affinity of CnSb do not exhibit a parity effect. There is an even-odd alternation for △EI of the open chain CnSb (n = 1–16) and CnSb+ (n = 1–10, when n > 10, △EI of n-odd members of CnSb+ are larger), with the n-even species being comparatively more stable than n-odd ones. The △EI for CnSb anions does not exhibit a parity effect. For carbon chain clusters, the most favorable binding site of atomic antimony is the terminal carbon of the carbon cluster because the terminal carbon with a large spin density bonds in an unsaturated way. The C-Sb bond is a double bond with WBI between 1.41 and 2.13, which is obviously stronger for a carbon chain cluster with odd-number carbon atoms (compared with neighboring carbon chain clusters with even-number carbon atoms). The WBI of all C-C bonds was determined to be between 1.63 and 2.01, indicating the cumulene character of the carbon chain. Generally, the alteration of WBI and, in particular, the carbon chain cluster is consistent with the bond length alteration. However, the shorter C-C distance did not indicate a larger WBI. For example, the largest WBI for CnSb (n > 1) was calculated to be the terminal C-C bond index, while this terminal C-C bond was not the shortest C-C bond. Therefore, rather than relying on the empirical comparison of bond distance, the WBI is a meaningful quantitative indicator for predicting the bonding strength in the carbon chain. For HOMO, the p orbital of the terminal carbon of the even-numbered carbon cluster always presents a different sign from the p orbital of antimony. While the p orbital of the terminal carbon of odd-numbered carbon cluster exhibits the same sign with a p orbital of antimony, it can form π bonding orbital with antimony. Therefore, the C-Sb bond of CnSb with even-numbered carbon should be weaker than the C-Sb bond of Cn±1Sb with odd-numbered carbon. This result agrees well with the Wiberg bond index analysis.

Author Contributions

Conceptualization, Z.S. and M.Y.; methodology, Z.S., X.S., W.W. and H.W.; software, Z.S. and M.Y.; validation, Z.S., M.L. and M.Y.; formal analysis, Z.S.; investigation, Z.S., X.S. and Z.L.; resources, Z.S., Z.L. and H.W.; data curation, Z.S.; writing—original draft preparation, Z.S., M.Y., M.L., X.S. and Z.L.; writing—review and editing, Z.S., H.W. and M.Y.; visualization, Z.S. and M.Y.; supervision, Z.S., H.W., M.Y. and M.L.; project administration, Z.S., M.Y. and H.W.; funding acquisition, Z.S., M.Y., M.L. and H.W. All authors have read and agreed to the published version of the manuscript.

Funding

Foundation of China (NSFC grant 22101198), and the Science & Technology Planning Project (grant 1901GY21, Taizhou Science & Technology Bureau).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was financially supported by the National Natural Science Foundation (NSFC grant 22101198) and the Science & Technology Planning Project (grant 1901GY21, Taizhou Science & Technology Bureau). We thank X.F. Wang (Department of Chemistry, Tongji University) for helpful discussions and valuable experimental assistance.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Sample Availability

Not applicable.

References

  1. Leyva-Perez, A. Sub-nanometre metal clusters for catalytic carbon-carbon and carbon-heteroatom cross-coupling reactions. Dalton T. 2017, 46, 15987–15990. [Google Scholar] [CrossRef] [PubMed]
  2. de Lara-Castells, M.P.; Mitrushchenkov, A.O. Mini Review: Quantum Confinement of Atomic and Molecular Clusters in Carbon Nanotubes. Front. Chem. 2021, 9, 796890. [Google Scholar] [CrossRef] [PubMed]
  3. Shi, W.; Song, Z.; Wang, J.; Li, Q.; An, Q. Phytic acid conversion film interfacial engineering for stabilizing zinc metal anode. Chem. Eng. J. 2022, 446, 137295. [Google Scholar] [CrossRef]
  4. Shao, X.; Li, L.; Huang, S.; Song, Z.; Wang, K. First-Principles Calculations of Magnetic Moment Modulation of 3D Transition Metal Atoms Encapsulated in C-60/C-70 Cages on Si(100) Surfaces: Implications for Spintronic Devices. ACS Appl. Nano Mater. 2021, 4, 12356–12364. [Google Scholar] [CrossRef]
  5. Chen, W.; Huang, J.; He, Z.-C.; Ji, X.; Zhang, Y.-F.; Sun, H.-L.; Wang, K.; Su, Z.-W. Accelerated photocatalytic degradation of tetracycline hydrochloride over CuAl2O4/g-C3N4 p-n heterojunctions under visible light irradiation. Sep. Purif. Technol. 2021, 277, 119461. [Google Scholar] [CrossRef]
  6. Liu, M.; Yu, T.; Huang, R.; Qi, W.; He, Z.; Su, R. Fabrication of nanohybrids assisted by protein-based materials for catalytic applications. Catal. Sci. Technol. 2020, 10, 3515–3531. [Google Scholar] [CrossRef]
  7. Pu, D.; Pan, Y. First-principles investigation of solution mechanism of C in TM-Si-C matrix as the potential high-temperature ceramics. J. Am. Ceram. Soc. 2022, 105, 2858–2868. [Google Scholar] [CrossRef]
  8. Liu, M.; Shan, C.; Chang, H.; Zhang, Z.; Huang, R.; Lee, D.W.; Qi, W.; He, Z.; Su, R. Nano-engineered natural sponge as a recyclable and deformable reactor for ultrafast conversion of pollutants from water. Chem. Eng. Sci. 2022, 247, 117049. [Google Scholar] [CrossRef]
  9. Tuktarov, A.R.; Khuzin, A.A.; Dzhemilev, U.M. Light-controlled molecular switches based on carbon clusters. Synthesis, properties and application prospects. Russ. Chem. Rev. 2017, 86, 474–509. [Google Scholar] [CrossRef]
  10. Hadad, C.Z.; Florez, E.; Merino, G.; Cabellos, J.L.; Ferraro, F.; Restrepo, A. Potential Energy Surfaces of WC6 Clusters in Different Spin States. J. Phys. Chem. A 2014, 118, 5762–5768. [Google Scholar] [CrossRef]
  11. Stein, T.; Bera, P.P.; Lee, T.J.; Head-Gordon, M. Molecular growth upon ionization of van der Waals clusters containing HCCH and HCN is a pathway to prebiotic molecules. Phys. Chem. Chem. Phys. 2020, 22, 20337–20348. [Google Scholar] [CrossRef] [PubMed]
  12. Yao, Y.; Xie, S. Structures and Progress of Carbon Clusters. Prog. Chem. 2019, 31, 50–62. [Google Scholar]
  13. Ticknor, B.W.; Bandyopadhyay, B.; Duncan, M.A. Photodissociation of Noble Metal-Doped Carbon Clusters. J. Phys. Chem. A 2008, 112, 12355–12366. [Google Scholar] [CrossRef] [PubMed]
  14. Lacovig, P.; Pozzo, M.; Alfe, D.; Vilmercati, P.; Baraldi, A.; Lizzit, S. Growth of dome-shaped carbon nanoislands on Ir(111): The intermediate between carbidic clusters and quasi-free-standing graphene. Phys. Rev. Lett. 2009, 103, 166101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Niu, T.C.; Zhou, M.; Zhang, J.L.; Feng, Y.P.; Chen, W. Growth Intermediates for CVD Graphene on Cu(111): Carbon Clusters and Defective Graphene. J. Am. Chem. Soc. 2013, 135, 8409–8414. [Google Scholar] [CrossRef] [PubMed]
  16. Song, Z.J. First-principle investigation on growth patterns and properties of cobalt-doped lithium nanoclusters. J. Mol. Model. 2016, 22, 133. [Google Scholar] [CrossRef]
  17. Song, Z.; Zhao, B.; Wang, Q.; Cheng, P. Homolytic cleavage of water on magnesia film promoted by interfacial oxide-metal nanocomposite. Appl. Surf. Sci. 2019, 487, 1222–1232. [Google Scholar] [CrossRef] [Green Version]
  18. Song, Z.; Shao, X.; Wang, Q.; Ma, C.; Wang, K.; Han, D. Generation of molybdenum hydride species via addition of molecular hydrogen across metal-oxygen bond at monolayer oxide/metal composite interface. Int. J. Hydrog. Energ. 2020, 45, 2975–2988. [Google Scholar] [CrossRef]
  19. Song, Z.; Fang, S.; Xie, P.; Zhong, A.; Ma, C.; Han, D. Unveiling the interaction profile of cisplatin with gold-supported magnesia film. Appl. Surf. Sci. 2021, 540, 148365. [Google Scholar] [CrossRef]
  20. Pei, C.-Y.; Li, T.; Zhang, M.; Wang, J.-W.; Chang, L.; Xiong, X.; Chen, W.; Huang, G.-B.; Han, D.-M. Synergistic effects of interface coupling and defect sites in WO3/InVO4 architectures for highly efficient nitrogen photofixation. Sep. Purif. Technol. 2022, 290, 120875. [Google Scholar] [CrossRef]
  21. Guo, X.-G.; Zhang, J.-L.; Zhao, Y. Ab initio Characterization of Size Dependence of Electronic Spectra for Linear Anionic Carbon Clusters Cn(n = 4–17). J. Comput. Chem. 2012, 33, 93–102. [Google Scholar] [CrossRef] [PubMed]
  22. Zhao, Y.; He, H.; Zhang, J.; Wang, L. Direct dynamics study of the hydrogen abstraction reaction of CF3CH2Cl + Cl -> CF3CHCl + HCl. Int. J. Chem. Kinet. 2012, 44, 661–667. [Google Scholar] [CrossRef]
  23. Fan, Y.; Wen, J.; Zhao, Y.; Wang, L. Theoretical study on the hydrogen abstraction reactions of CF3CHFCF3 and CF3CF2CHF2 with X atoms (X = F, Cl, and Br). J. Fluorine Chem. 2013, 150, 39–45. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Guo, J.; Zhang, J. A theoretical study of electronic spectra in the linear cationic chains NC2n+1N+ (n = 1–6). Chem. Phys. 2011, 386, 23–28. [Google Scholar] [CrossRef]
  25. Zhan, C.-G.; Iwata, S. Ab initio studies on the structures, vertical electron detachment energies, and stabilities of CnP clusters. J. Chem. Phys. 1997, 107, 7323–7330. [Google Scholar] [CrossRef]
  26. Sun, S.; Cao, Y.; Sun, Z.; Tang, Z.; Zheng, L. Experimental and Theoretical Studies on Carbon−Nitrogen Clusters C2nN7. J. Phys. Chem. A 2006, 110, 8064–8072. [Google Scholar] [CrossRef]
  27. Tang, Z.; BelBruno, J.J.; Huang, R.; Zheng, L. Collision-induced dissociation and density functional study of the structures and energies of cyclic C2nN5 clusters. J. Chem. Phys. 2000, 112, 9276–9281. [Google Scholar] [CrossRef]
  28. Li, G.; Xing, X.; Tang, Z. Structures and properties the lead-doped carbon clusters PbCn/PbCn+/PbCn (n = 1–10). J. Chem. Phys. 2003, 118, 6884–6897. [Google Scholar] [CrossRef]
  29. Zhang, C.-J.; Jiang, Z.-Y.; Wang, Y.-L.; Zhu, H.-Y. Structures and stabilities of GaCn (n = 1–10) clusters. Comput. Theor. Chem. 2013, 1004, 12–17. [Google Scholar] [CrossRef]
  30. Chen-Jun, Z.; Yang-Li, W.; Chao-Kang, C. Density functional theory of InCn+ (n = 1–10) clusters. Acta Phys. Sin. 2018, 67, 113101. [Google Scholar] [CrossRef]
  31. Wang, P.; Yan, S.-T.; Xu, H.-G.; Xu, X.-L.; Zheng, W.-J. Anion photoelectron spectroscopy and density functional theory studies of AuCn/0 (n = 3−8): Odd-even alternation in electron binding energies and structures. Chin. J. Chem. Phys. 2022, 35, 177–184. [Google Scholar] [CrossRef]
  32. Zhang, C.; Xu, H.; Xu, X.-L.; Zheng, W.-J. Electronic structures, chemical bonds, and stabilities of Ta4C/0n (n = 0–4) clusters: Anion photoelectron spectroscopy and theoretical calculations. Acta Phys. Sin. 2021, 70, 023601. [Google Scholar] [CrossRef]
  33. Freitas, A.; Azevedo, S.; Kaschny, J.R. Structural and electronic properties of linear carbon chains encapsulated by flattened nanotubes. Phys. E-Low-Dimens. Syst. Nanostruct. 2016, 84, 444–453. [Google Scholar] [CrossRef]
  34. Rocha, R.A.; dos Santos, R.B.; Ribeiro Junior, L.A.; Aguiar, A.L. On the Stabilization of Carbynes Encapsulated in Penta-Graphene Nanotubes: A DFT Study. J. Mol. Model. 2021, 27, 318. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Su, Y.; Wang, L.; Kong, E.S.-W.; Chen, X.; Zhang, Y. A one-dimensional extremely covalent material: Monatomic carbon linear chain. Nanoscale Res. Lett. 2011, 6, 577. [Google Scholar] [CrossRef] [Green Version]
  36. Yang, S.; Li, W.; Li, Y.F.; Chen, X.M.; Zhang, H.; Xu, B.Q.; Yang, B. Structural, electronic and catalytic properties of AgnSnn (n = 2–14) clusters by density functional theory. Phys. Chem. Chem. Phys. 2022, 24, 26631–26641. [Google Scholar] [CrossRef] [PubMed]
  37. Pascoli, G.; Lavendy, H. Theoretical Study of CnP, CnP+, CnP(n = 1–7) Clusters. J. Phys. Chem. A 1999, 103, 3518–3524. [Google Scholar] [CrossRef]
  38. Pascoli, G.; Lavendy, H. Are CnNclusters really bent? Chem. Phys. Lett. 1999, 312, 333–340. [Google Scholar] [CrossRef]
  39. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.K.; Kudin, K.N.; Burant, J.C.; et al. Gaussian 03 Rev. A01; Gaussian Inc.: Wallingford, CT, USA, 2004. [Google Scholar]
  40. Pan, Y.; Yu, E. New insight into the structural and physical properties of AlH3. Int. J. Energ. Res. 2022, 46, 19678–19685. [Google Scholar] [CrossRef]
  41. Song, Z.; Zhao, B.; Wang, Q.; Cheng, P. Steering reduction and decomposition of peroxide compounds by interface interactions between MgO thin film and transition-metal support. Appl. Surf. Sci. 2018, 459, 812–821. [Google Scholar] [CrossRef]
  42. Chen, S.; Pan, Y. Enhancing catalytic properties of noble metal@MoS2/WS2 heterojunction for the hydrogen evolution reaction. Appl. Surf. Sci. 2022, 591, 153168. [Google Scholar] [CrossRef]
  43. Pan, Y. First-principles investigation of the effect of noble metals on the electronic and optical properties of GaN nitride. Mat. Sci. Semicon. Proc. 2022, 151, 107051. [Google Scholar] [CrossRef]
  44. Zhou, Y.; Shao, X.J.; Lam, K.H.; Zheng, Y.; Zhao, L.Z.; Wang, K.D.; Zhao, J.Z.; Chen, F.M.; Hou, X.H. Symmetric Sodium-Ion Battery Based on Dual-Electron Reactions of NASICON-Structured Na3MnTi(PO4)3 Material. ACS Appl. Mater. Inter. 2020, 12, 30328–30335. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, W.P.; Zhao, J.Z.; Xu, H. Adsorption Induced Indirect-to-Direct Band Gap Transition in Monolayer Blue Phosphorus. J. Phys. Chem. C 2018, 122, 15792–15798. [Google Scholar] [CrossRef]
  46. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wu, J.Y.; Yan, H.; Zhong, A.G.; Chen, H.; Jin, Y.X.; Dai, G.L. Theoretical and conceptual DFT study of pnicogen-and halogen-bonded complexes of PH2X—BrCl. J. Mol. Model. 2019, 25, 28. [Google Scholar] [CrossRef] [PubMed]
  48. Zhu, H.Y.; Wu, J.Y.; Dai, G.L. Study on the halogen bond and pi-pi stacking interaction between fluoro substituted iodobenzene and pyrazine. J. Mol. Model. 2020, 26, 333. [Google Scholar] [CrossRef]
  49. Wu, J.Y.; Yan, H.; Chen, H.; Jin, Y.X.; Zhong, A.G.; Wang, Z.X.; Dai, G.L. Three types of noncovalent interactions studied between pyrazine and XF. J. Mol. Model. 2022, 28, 15. [Google Scholar] [CrossRef]
  50. Wu, J. Investigations into the nature of halogen-and hydrogen-bonding interactions of some heteroaromatic rings with dichlorine monoxide. J. Mol. Model. 2014, 20, 2424. [Google Scholar] [CrossRef]
  51. Pascoli, G.; Lavendy, H. Geometrical structures of the phosphorus-doped carbon cluster cations CnP+ (n = 1–20). Int. J. Mass Spectrom. 1999, 189, 125–132. [Google Scholar] [CrossRef]
  52. Pascoli, G.; Lavendy, H. Structures and energies of CnSi+ (4 ≤ n ≤ 15) silicon carbide clusters. Int. J. Mass Spectrom. 1998, 173, 41–54. [Google Scholar] [CrossRef]
Figure 1. C-C bond length alteration of neutral C15Sb, C15Sb+ cation, C15Sb anion, and pure carbon chain C15 as a function of number of carbon atoms.
Figure 1. C-C bond length alteration of neutral C15Sb, C15Sb+ cation, C15Sb anion, and pure carbon chain C15 as a function of number of carbon atoms.
Molecules 28 01358 g001
Figure 2. C-C bond length alteration of neutral C16Sb,C16Sb+ cation, C16Sb anion, and pure carbon chain C16 as a function of number of carbon atoms.
Figure 2. C-C bond length alteration of neutral C16Sb,C16Sb+ cation, C16Sb anion, and pure carbon chain C16 as a function of number of carbon atoms.
Molecules 28 01358 g002
Figure 3. C-C bond length alteration of neutral CnSb (n = 11–16) as a function of number of carbon atoms.
Figure 3. C-C bond length alteration of neutral CnSb (n = 11–16) as a function of number of carbon atoms.
Molecules 28 01358 g003
Figure 4. C-C bond length alteration of CnSb+ cations (n = 11–16) as a function of number of carbon atoms.
Figure 4. C-C bond length alteration of CnSb+ cations (n = 11–16) as a function of number of carbon atoms.
Molecules 28 01358 g004
Figure 5. C-C bond length alteration of CnSb anions (n = 11–16) as a function of number of carbon atoms.
Figure 5. C-C bond length alteration of CnSb anions (n = 11–16) as a function of number of carbon atoms.
Molecules 28 01358 g005
Figure 6. C-Sb bond length of neutral CnSb, CnSb+ cation, and CnSb anion as a function of number of carbon atoms.
Figure 6. C-Sb bond length of neutral CnSb, CnSb+ cation, and CnSb anion as a function of number of carbon atoms.
Molecules 28 01358 g006
Figure 7. Highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO) for carbon chain clusters. The surface isovalue for molecular orbitals is 0.02.
Figure 7. Highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO) for carbon chain clusters. The surface isovalue for molecular orbitals is 0.02.
Molecules 28 01358 g007
Figure 8. Adiabatic ionization energies of neutral CnSb and pure carbon clusters Cn as a function of the number of carbons.
Figure 8. Adiabatic ionization energies of neutral CnSb and pure carbon clusters Cn as a function of the number of carbons.
Molecules 28 01358 g008
Figure 9. Adiabatic electron affinities of neutral CnSb and pure carbon clusters Cn as a function of the number of carbons.
Figure 9. Adiabatic electron affinities of neutral CnSb and pure carbon clusters Cn as a function of the number of carbons.
Molecules 28 01358 g009
Figure 10. The NPA (natural population analysis) charge distribution for CnSb chain clusters. The color range is depicted from −1 to 1. The small balls and the large balls in the right end indicate carbon and antimony atoms, respectively.
Figure 10. The NPA (natural population analysis) charge distribution for CnSb chain clusters. The color range is depicted from −1 to 1. The small balls and the large balls in the right end indicate carbon and antimony atoms, respectively.
Molecules 28 01358 g010
Figure 11. Computed Wiberg bond indices (WBI) for CnSb.
Figure 11. Computed Wiberg bond indices (WBI) for CnSb.
Molecules 28 01358 g011
Figure 12. Increment binding energies of CnSb and pure Cn chains as a function of the number of carbon atoms.
Figure 12. Increment binding energies of CnSb and pure Cn chains as a function of the number of carbon atoms.
Molecules 28 01358 g012
Figure 13. Incremental binding energies of CnSb+ cations and pure Cn cations as a function of the number of carbon atoms.
Figure 13. Incremental binding energies of CnSb+ cations and pure Cn cations as a function of the number of carbon atoms.
Molecules 28 01358 g013
Figure 14. Incremental binding energies of CnSb anions and pure Cn anions.
Figure 14. Incremental binding energies of CnSb anions and pure Cn anions.
Molecules 28 01358 g014
Figure 15. Fragmentation energies of neutral CnSb clusters. DNx corresponds to different dissociation channels, as shown in the text.
Figure 15. Fragmentation energies of neutral CnSb clusters. DNx corresponds to different dissociation channels, as shown in the text.
Molecules 28 01358 g015
Figure 16. Fragmentation energies of CnSb+ cations. DCx corresponds to different dissociation channels, as shown in the text.
Figure 16. Fragmentation energies of CnSb+ cations. DCx corresponds to different dissociation channels, as shown in the text.
Molecules 28 01358 g016
Figure 17. Fragmentation energies of CnSb anions. DAx corresponds to different dissociation channels, as shown in the text.
Figure 17. Fragmentation energies of CnSb anions. DAx corresponds to different dissociation channels, as shown in the text.
Molecules 28 01358 g017
Figure 18. The most favorable dissociation channels for CnSb/ CnSb+ / CnSb.
Figure 18. The most favorable dissociation channels for CnSb/ CnSb+ / CnSb.
Molecules 28 01358 g018
Figure 19. Structural geometry, adiabatic ionization energy (AIE), relative binding energy (RBE), and charge population influenced by binding site of antimony atom. Relative binding energy is calculated by equation: RBE = E b ( nt ) E b ( t ) , where E b ( nt ) and E b ( t ) stand for structure with antimony binding on nonterminal carbon and linear structure with antimony binding on terminal carbon, respectively. Notations na, nb, and nc represent the structures with n carbon atoms.
Figure 19. Structural geometry, adiabatic ionization energy (AIE), relative binding energy (RBE), and charge population influenced by binding site of antimony atom. Relative binding energy is calculated by equation: RBE = E b ( nt ) E b ( t ) , where E b ( nt ) and E b ( t ) stand for structure with antimony binding on nonterminal carbon and linear structure with antimony binding on terminal carbon, respectively. Notations na, nb, and nc represent the structures with n carbon atoms.
Molecules 28 01358 g019
Table 1. Optimized bond lengths (Å) for neutral CnSb at b3lyp/6-31G(d) theoretical level.
Table 1. Optimized bond lengths (Å) for neutral CnSb at b3lyp/6-31G(d) theoretical level.
CSb1.888
C2Sb1.3001.941
C3Sb1.3011.2961.949
C4Sb1.2911.3041.2771.941
C5Sb1.2941.2971.2801.2881.946
C6Sb1.2911.3011.2711.2951.2771.949
C7Sb1.2921.2991.2741.2881.2821.2861.945
C8Sb1.2911.3011.2711.2931.2721.2961.2761.956
C9Sb1.2911.3001.2721.2891.2761.2851.2841.2851.947
C10Sb1.2901.3011.2701.2931.2721.2921.2721.2971.2751.960
C11Sb1.2911.3001.2721.2901.2751.2871.2781.2841.2861.2841.949
C12Sb1.2901.3011.2701.2931.2721.2921.2721.2931.2711.2981.2741.964
C13Sb1.2911.3001.2711.2911.2741.2881.2771.2861.2801.2831.2871.2831.951
C14Sb1.2901.3021.2701.2931.2711.2931.2721.2931.2721.2941.2711.3001.2731.966
C15Sb1.2901.3011.2711.2911.2741.2891.2761.2871.2781.2851.2811.2821.2881.2821.952
C16Sb1.2901.3021.2701.2941.2711.2931.2711.2931.2721.2941.2711.2951.2701.3001.2731.968
Table 2. Optimized bond length (Å) for CnSb+ cations at b3lyp/6-31G(d) theoretical level.
Table 2. Optimized bond length (Å) for CnSb+ cations at b3lyp/6-31G(d) theoretical level.
CSb1.963
C2Sb1.3261.889
C3Sb1.3341.2702.009
C4Sb1.3151.2771.3001.891
C5Sb1.3211.2721.3021.2641.995
C6Sb1.3111.2761.2911.2681.3011.894
C7Sb1.3151.2761.2931.2641.3031.2651.985
C8Sb1.3081.2781.2901.2661.2951.2661.3031.896
C9Sb1.3111.2781.2901.2681.2951.2641.3021.2671.979
C10Sb1.3061.2791.2891.2671.2951.2631.2981.2651.3041.897
C11Sb1.3091.2801.2881.2701.2921.2671.2961.2651.3021.2681.976
C12Sb1.3041.2811.2871.2681.2941.2641.2981.2621.3001.2641.3041.898
C13Sb1.3071.2821.2861.2711.2901.2691.2931.2671.2961.2661.3011.2691.973
C14Sb1.3031.2831.2861.2701.2921.2651.2971.2621.3001.2601.3011.2641.3051.897
C15Sb1.3051.2831.2851.2721.2891.2701.2921.2691.2931.2681.2951.2661.3001.2701.971
C16Sb1.3021.2841.2851.2711.2911.2661.2961.2631.3001.2601.3021.2591.3031.2631.3061.898
Table 3. Optimized bond length (Å) for CnSb anions at b3lyp/6-31G(d) theoretical level.
Table 3. Optimized bond length (Å) for CnSb anions at b3lyp/6-31G(d) theoretical level.
CSb1.952
C2Sb1.2802.014
C3Sb1.2781.3261.924
C4Sb1.2731.3361.2602.000
C5Sb1.2731.3261.2621.3141.921
C6Sb1.2731.3291.2561.3231.2611.997
C7Sb1.2721.3261.2561.3151.2631.3111.919
C8Sb1.2731.3261.2571.3171.2581.3191.2621.994
C9Sb1.2731.3251.2561.3161.2571.3131.2641.3101.917
C10Sb1.2741.3241.2571.3141.2591.3141.2591.3171.2631.992
C11Sb1.2731.3241.2561.3151.2561.3141.2571.3121.2631.3101.914
C12Sb1.2751.3221.2581.3121.2591.3121.2601.3111.2601.3151.2641.990
C13Sb1.2741.3221.2561.3141.2561.3141.2561.3141.2571.3121.2631.3101.913
C14Sb1.2761.3211.2581.3121.2591.3111.2601.3101.2611.3101.2611.3141.2651.987
C15Sb1.2751.3211.2571.3131.2571.3141.2561.3151.2551.3151.2561.3121.2621.3101.912
C16Sb1.2761.3201.2581.3111.2601.3101.2601.3101.2611.3091.2611.3091.2621.3121.2661.985
Table 4. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies BE, adiabatic ionization energies IE, adiabatic electron affinities EA, dipole moments, and rotational constants (RC) of neutral CnSb.
Table 4. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies BE, adiabatic ionization energies IE, adiabatic electron affinities EA, dipole moments, and rotational constants (RC) of neutral CnSb.
CnSbEn
(a. u.)
ΔEn
(a. u.)
BE
(a. u.)
ΔEI
(kcal/mol)
IE
(kcal/mol)
EA
(kcal/mol)
μ
(debye)
RC
(MHz)
1−43.35455 0.10241 212.469.23.1312990
2−81.48162−38.127070.37201169.2196.362.65.573497
3−119.56561−38.083990.59853142.1196.172.15.361519
4−157.65877−38.093160.83422147.9187.772.97.46821
5−195.74409−38.085321.06207143.0186.378.77.10496
6−233.83823−38.094141.29874148.5183.976.78.98326
7−271.91968−38.081451.52272140.5179.183.08.75227
8−310.00802−38.088341.75359144.9176.184.110.39165
9−348.09409−38.086071.98219143.4173.886.210.31125
10−386.18172−38.087632.21235144.4172.187.511.7497
11−424.26787−38.086152.44103143.5169.588.611.7877
12−462.35529−38.087422.67098144.3169.289.513.0462
13−500.44149−38.086202.89971143.5166.290.513.1951
14−538.52879−38.08733.12954144.2166.892.414.3042
15−576.61521−38.086423.35849143.7163.592.014.5336
16−614.70241−38.087203.58822144.2164.894.215.5130
Table 5. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies (BE), dipole moments (μ), and rotational constants (RC) of CnSb+ cations.
Table 5. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies (BE), dipole moments (μ), and rotational constants (RC) of CnSb+ cations.
CnSb+En
(a. u.)
ΔEn
(a. u.)
BE
(a. u.)
ΔEI
(kcal/mol)
μ(Debye)RC
(MHz)
1−43.01612 −0.23602 1.0812013
2−81.16886−38.152740.059250.295272.683584
3−119.25339−38.084530.286310.227050.851474
4−157.35996−38.106570.535410.249112.26835
5−195.44758−38.087620.765560.230140.52489
6−233.54567−38.098091.006180.240631.97331
7−271.63483−38.089161.237870.231690.43226
8−309.72820−38.093371.473770.23591.89167
9−347.81792−38.089721.706020.232250.48124
10−385.90836−38.090441.938990.232970.8398
11−423.99872−38.090362.171880.23290.6577
12−462.08689−38.088172.402580.230691.9962
13−500.17791−38.091022.636130.233550.8751
14−538.26446−38.086552.865210.229082.0643
15−576.35619−38.091733.099470.234261.1136
16−614.44148−38.085293.327290.227822.1031
Table 6. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies BE, dipole moments (μ), and rotational constants (RC) of CnSb- anions.
Table 6. The total energies En, differential energies ΔEn, incremental energies ΔEI, binding energies BE, dipole moments (μ), and rotational constants (RC) of CnSb- anions.
CnSbEn
(a. u.)
ΔEn
(a. u.)
BE
(a. u.)
ΔEI
(kcal/mol)
μ
(Debye)
RC
(MHz)
1−43.46493 0.21279 5.5012153
2−81.58147−38.116540.471860.259088.893348
3−119.68068−38.099210.71360.241749.781528
4−157.77531−38.094630.950760.2371513.03800
5−195.86996−38.094651.187940.2371813.52496
6−233.96096−38.091001.421470.2335316.53321
7−272.05265−38.091691.655690.2342216.93227
8−310.14274−38.090091.888310.2326219.74163
9−348.23223−38.089492.120330.2320220.16125
10−386.32214−38.089912.352770.2324322.7696
11−424.41010−38.087962.583260.2304923.2677
12−462.49915−38.089052.814840.2315825.0462
13−500.58704−38.087893.045260.2304226.3051
14−538.67748−38.090443.278230.2329628.5242
15−576.76336−38.085883.506640.2284129.3336
16−614.85418−38.090823.739990.2333531.3330
Table 7. The selected structural parameters (the C-C bond adjacent to Sb and the C-Sb bond in Å), C-Sb stretching frequencies (in cm−1), and HOMO-LUMO gaps (in eV) for results obtained at the B3LYP/6-31G(d)/SDD and B3LYP/def2-TZVP/SDD level (denoted as normal and italic fonts, respectively) for C4Sb, C5Sb, C6Sb, C10Sb, and C11Sb.
Table 7. The selected structural parameters (the C-C bond adjacent to Sb and the C-Sb bond in Å), C-Sb stretching frequencies (in cm−1), and HOMO-LUMO gaps (in eV) for results obtained at the B3LYP/6-31G(d)/SDD and B3LYP/def2-TZVP/SDD level (denoted as normal and italic fonts, respectively) for C4Sb, C5Sb, C6Sb, C10Sb, and C11Sb.
C4SbC5SbC6SbC11SbC12Sb
C-C1.277
1.267
1.288
1.281
1.277
1.267
1.284
1.275
1.274
1.263
C-Sb1.941
1.941
1.946
1.941
1.949
1.949
1.949
1.949
1.964
1.966
C-Sb Stretching frequency378.3
384.5
331.3
335.2
291.7
294.9
204.8
204.9
190.4
189.6
HOMO-LUMO gap1.27
1.26
1.25
1.27
1.17
1.16
1.03
1.04
1.01
0.99
Table 8. Valence orbital configuration of CnSb.
Table 8. Valence orbital configuration of CnSb.
n   orbital configuration
1   (core) 1σ2241
2   (core)1σ222241
3   (core)1σ2222423
4   (core)1σ222224421
5   (core)1σ2222242423
6   (core)1σ222222424421
7   (core)1σ2222222424423
8   (core)1σ222222224244410σ21
9   (core)1σ2222222224410σ24411σ23
10   (core)1σ22222222210σ24411σ244412σ21
11   (core)1σ22222222210σ211σ24412σ244413σ23
12   (core)1σ22222222210σ211σ212σ24413σ2444414σ21
13   (core)1σ22222222210σ211σ212σ213σ24414σ2444415σ23
14   (core)1σ22222222210σ211σ212σ213σ214σ244415σ2444416σ21
15   (core)1σ22222222210σ211σ212σ213σ214σ215σ244416σ2444417σ23
16   (core)1σ22222222210σ211σ212σ213σ214σ215σ216σ244417σ24444418σ21
Table 9. The charge population and charge difference ( | Δ C | ) of terminal carbon atoms which are adjacent antimony (C1) and far away from antimony (C2).
Table 9. The charge population and charge difference ( | Δ C | ) of terminal carbon atoms which are adjacent antimony (C1) and far away from antimony (C2).
SpeciesC1C2 | Δ C | SpeciesC1C2 | Δ C |
C1Sb−0.33 C9Sb−0.76−0.140.63
C2Sb−0.47−0.050.42C10Sb−0.80−0.140.66
C3Sb−0.45−0.140.31C11Sb−0.77−0.130.64
C4Sb−0.73−0.220.51C12Sb−0.80−0.130.67
C5Sb−0.72−0.190.53C13Sb−0.77−0.130.65
C6Sb−0.80−0.180.62C14Sb−0.80−0.130.67
C7Sb−0.75−0.150.60C15Sb−0.78−0.130.65
C8Sb−0.80−0.150.65C16Sb−0.80−0.130.67
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

Song, Z.; Shao, X.; Wu, W.; Liu, Z.; Yang, M.; Liu, M.; Wang, H. Structures and Stabilities of Carbon Chain Clusters Influenced by Atomic Antimony. Molecules 2023, 28, 1358. https://doi.org/10.3390/molecules28031358

AMA Style

Song Z, Shao X, Wu W, Liu Z, Yang M, Liu M, Wang H. Structures and Stabilities of Carbon Chain Clusters Influenced by Atomic Antimony. Molecules. 2023; 28(3):1358. https://doi.org/10.3390/molecules28031358

Chicago/Turabian Style

Song, Zhenjun, Xiji Shao, Wei Wu, Zhenzhong Liu, Meiding Yang, Mingyue Liu, and Hai Wang. 2023. "Structures and Stabilities of Carbon Chain Clusters Influenced by Atomic Antimony" Molecules 28, no. 3: 1358. https://doi.org/10.3390/molecules28031358

Article Metrics

Back to TopTop