Next Article in Journal
A Diverse and Versatile Regiospecific Synthesis of Tetrasubstituted Alkylsulfanylimidazoles as p38α Mitogen-Activated Protein Kinase Inhibitors
Previous Article in Journal
Synthesis and Antimicrobial Activity of Sulfur Derivatives of Quinolinium Salts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

icMRCI+Q Study of the Spectroscopic Properties of the 14 Λ-S and 49 Ω States of the SiN Anion in the Gas Phase

1
School of Materials Science and Engineering, Henan University of Science and Technology, Luoyang 471023, China
2
College of Physics and Electronic Engineering, Xinyang Normal University, Xinyang 464000, China
3
College of Physics and Material Science, Henan Normal University, Xinxiang 453007, China
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(1), 210; https://doi.org/10.3390/molecules23010210
Submission received: 26 November 2017 / Revised: 23 December 2017 / Accepted: 16 January 2018 / Published: 20 January 2018
(This article belongs to the Section Physical Chemistry)

Abstract

:
This paper calculates the potential energy curves of the 14 Λ-S and 49 Ω states, which come from the first three dissociation channels of the SiN anion. These calculations are conducted using the valence internally contracted multireference configuration interaction and the Davidson correction approach. Core-valence correlation and scalar relativistic corrections are taken into account. The potential energies are extrapolated to the complete basis set limit. The spin-orbit coupling is computed using the state interaction approach with the Breit–Pauli Hamiltonian. We found that the X1Σ+ (υ′′ = 0–23) and a3Σ+ (υ′ = 0–2) states of SiN are stable at the computed adiabatic electron affinity value of 23,262.27 cm−1 for SiN. Based on the calculated potential energy curves, the spectroscopic parameters and vibrational levels were determined for all stable and metastable Λ-S and Ω states. The computed adiabatic electron affinity of SiN and the spectroscopic constants of SiN (X1Σ+) are all in agreement with the available experimental data. The d3Σ+, 25Σ+, 15Δ, and 15Σ quasi-bound states caused by avoided crossings were found. Calculations of the transition dipole moment of a3Σ+1 to X1Σ+0+ are shown. Franck-Condon factors, Einstein coefficients, and radiative lifetimes of the transition from the a3Σ+1 (υ′ = 0–2) to the X1Σ+0+ state are evaluated.

1. Introduction

SiN is analogous with the known interstellar CN anion [1]. Several silicon-containing nitrogen chains (SiN, SiCN, and SiNC) have been detected in the interstellar medium [2,3,4]. Consequently, SiN is a tantalizing potential interstellar anion candidate [5]. Furthermore, the SiN anion is a compound of considerable importance in gas-phase ion chemistry [6], laser-induced plasmas [7], semiconductor chemistry [8], and microelectronics [9]. However, very few experiments [10] and computations [11,12,13,14,15,16] have been conducted for its spectroscopic properties. Based on the aforementioned facts, we systematically investigated the spectroscopic properties of the SiN anion using a highly accurate ab initio approach.
Experimentally, Meloni et al. [10] used anion photoelectron spectroscopy to obtain the SiN (X2Σ+, A2Π) + e ← SiN (X1Σ+) transitions for the first time in 2004. They determined the equilibrium bond length (Re), harmonic frequency (ωe), and dissociation energy (D0) for the X1Σ+ state of SiN. The adiabatic electron affinity (AEA) of SiN was also determined.
Theoretically, only six groups of computations [11,12,13,14,15,16] were performed. In 1989, Peterson and Woods [11] used the Møller–Plesset perturbation theory with single, double, and quadruple substitutions (MP4SDQ) and two large Gaussian basis sets to calculate the AEA (26,826.21 cm−1) of SiN and the spectroscopic constants of the SiN (X1Σ+). In 1992, McLean et al. [12] studied the potential energy curve (PEC) of SiN (X1Σ+) at the singles and double excitation from a single reference configuration interaction plus Davidson correction (SDCI + Q) levels of theory. Kalcher [13] in 2002 studied the AEA (23,083.58 cm−1) of SiN and the PECs of X1Σ+ and a3Σ+ states using the complete active space in conjunction with the averaged coupled pair functional (CAS-ACPF) approach using the cc-pVQZ basis set. In 2003, Midda and Das [14] calculated the spectroscopic constants and molecular properties of X1Σ+ states using a hybrid HF/DF B3LYP method with four different basis sets. Kerkines and Mavridis [15] in 2005 calculated the spectroscopic constants and energetics for the SiN (X2Σ+, A2Π) and SiN (X1Σ+, a3Σ+) at the level of the theory of the restricted coupled-cluster method with all singles and doubles and noniterative inclusion of triples [RCCSD(T)]. Their best estimated AEA is 24,212.75 cm−1. Recently, Mogren et al. [16] used Gaussian-3 theory to study the Re and ωe of the X1Σ+ state of SiN. Furthermore, they derived the AEA (25,648.42 cm−1) and vertical electron affinity (VEA = 24,115.97 cm−1) at the MP2/6-31G* level. After summarizing these theoretical spectroscopic results [11,12,13,14,15,16], we firstly found that all of the calculations were mainly focused on the X1Σ+ and a3Σ+ states, and few results achieved high quality. Secondly, it was observed that no spin-orbit coupling (SOC) interactions were involved, though the SOC effect could influence the accurate prediction of spectroscopic properties [17,18]. Finally, it was found that no transition probabilities, such as Franck-Condon (FC) factors and the radiative lifetimes of the a3Σ+1 (υ′ = 0–2) to X1Σ+0+ transition, were calculated despite the fact that the transition probabilities were very useful in observing the corresponding states. Therefore, to improve the quality of the spectroscopic parameters of the anion, more accurate calculations should be done.
This paper is organized as follows. The methodology is briefly introduced in the next section. The PECs are reported in Section 3. The spectroscopic parameters and vibrational properties are predicted. The SOC effect on the spectroscopic parameters and vibrational levels is evaluated. The transition dipole moments (TDMs) between the a3Σ+1 and X1Σ+0+ states are determined. Calculations of FC factors and the radiative lifetimes of the a3Σ+1 to X1Σ+0+ transition are shown. Some conclusions are drawn in Section 4. The spectroscopic parameters, vibrational levels, and transition probabilities obtained here can be considered very reliable.

2. Methodology

The electron affinities of ground-state Si and N atoms are 11,207.24 and −560 cm−1 [19], respectively. Thus, the first dissociation channel of the SiN anion should be Si(4Su) + N(4Su). The first and second excited states of Si anion are 2Du and 2Pu; their energy levels relative to the ground state (4Su) are approximately 6961.85 and (10,977.24 + xc)/2 cm−1 [19], respectively. The first excited state of the N atom is 2Du; its energy level relative to its ground state (4Su) is approximately 19,228.82 cm−1 [20]. Using the electron affinities and energy levels, we can determine that the second and third dissociation channels of the SiN anion are Si(2Du) + N(4Su) and Si(2Pu) + N(4Su), respectively. With the help of group theory, 4 states (X1Σ+, a3Σ+, 15Σ+, and 17Σ+) were correlated to the first dissociation channel, 6 states (b3Δ, c3Π, d3Σ+, 25Σ+, 15Π, and 15Δ) were correlated to the second dissociation limit, and 4 states (e3Σ, 15Σ, f3Π, and 25Π) were correlated to the third dissociation limit. These states and asymptotes are collected in Table 1. Using the approaches outlined above, we have determined the energy separation between each higher dissociation channel and the lowest one, i.e., Si(4Su) + N(4Su). For reasons of comparison, we also collected the energy separation obtained using the icMRCI + Q/56 + CV + DK calculations and experiments [19] in Table 1. As seen in Table 1, the present results agree favorably with the measurements [19]. The results indicate that our calculations can properly describe the dissociation properties of SiN.
The PECs were calculated using the complete active space self-consistent field (CASSCF) method followed by the internally contracted multireference configuration interaction (icMRCI) plus the Davidson modification (icMRCI + Q) approach [21,22] for internuclear separations from approximately 0.10 to 1.08 nm. As a result, the CASSCF was used as the reference wavefunction for the icMRCI calculations. In the calculations for the Si and N atoms, the basis sets aug-cc-pV5Z (AV5Z) and aug-cc-pV6Z (AV6Z) [23,24] were employed. The calculations were done with the MOLPRO 2010.1 program package [25] in the C2v point group. The point spacing interval used for calculating the PECs is 0.02 nm for each state. To accurately determine each PEC, the point spacing interval used was further reduced to 0.002 nm for internuclear separations from approximately 0.11 to 0.50 nm; this is because the equilibrium separations fall into this range for all the bound and quasi-bound states involved. It should be noted that these point spacing intervals were suitable for all the calculations including the core–valence correlation, scalar relativistic corrections, and the SOC effect.
The molecular orbitals (MOs) used for the icMRCI calculations come from the CASSCF results. The state-averaged technique was used in the CASSCF calculations. To accurately determine the interaction between different PECs (such as avoided crossings), we put the 18 electronic states together into the calculations. Each state has the same weight factor of 0.0588235. In this paper, we only reported the PECs of 14 states arising from the first three dissociation limits. In the icMRCI calculations, we put the 8 outermost MOs (4a1, 2b1, and 2b2) into the active space. This corresponds to the 5-8σ, 2π, and 3π MOs in the anion. No additional MOs were included in the active space. That is, the 10 valence electrons were distributed into the 8 valence MOs of the SiN anion. Consequently, this active space was referred to as CAS (10, 8). The rest of the 12 inner electrons were put into the 6 lowest MOs (4a1, 1b1, and 1b2). For the AV6Z basis set, the total number of external orbitals is 368, corresponding to 126a1, 90b1, 90b2, and 62a2. In summary, we used 14 MOs (8a1, 3b1, and 3b2) to calculate the PECs of all the 14 Λ-S states.
Scalar relativistic correction was computed using the cc-pV5Z-DK basis set [26]. Its contribution to the total energy is denoted as DK. The core-valence correlation correction was calculated with the cc-pCVTZ basis set [27]. Its contribution to the total energy is denoted as CV. The SOC effect was determined by the state interaction method with the Breit-Pauli operator [28] at the level of the icMRCI theory with the all-electron cc-pCVTZ basis set. The all-electron cc-pCVTZ basis set with and without the Breit–Pauli operator was used to calculate the contribution to potential energy via the SOC effect. The difference between the two energies is the SOC splitting energy, which is denoted as SOC. The extrapolation of potential energy to the complete basis set (CBS) limit was performed with the AV5Z and AV6Z basis sets. The energy obtained from the extrapolation is denoted as 56. The extrapolation scheme [29] is as follows:
Δ E X r e f = E r e f + A r e f X α
Δ E X c o r r = E c o r r + A c o r r X β .
Here, Δ E X r e f and Δ E X c o r r are the CASSCF and correlation energies obtained by the aug-cc-pVXZ basis set, respectively. E r e f and E c o r r are the CASSCF and correlation energies extrapolated to the CBS limit, respectively. The extrapolation parameters α and β are taken as 3.4 and 2.4 for the Hartree–Fock and correlation energies [29], respectively.
With the PECs, the spectroscopic parameters Te, Re, ωe, ωexe, ωeуe, αe, Be, and De were evaluated. The meanings of these spectroscopic notations are explained in our earlier paper [30,31]. All the PECs were fitted to an analytical form by cubic splines. The rovibrational constants were first obtained from the analytic potential by solving the rovibrational Schrödinger equation, and the spectroscopic parameters were then evaluated by fitting the vibrational levels.

3. Results and Discussion

Figure 1 and Figure 2 show the PECs of 14 electronic states obtained by the icMRCI + Q/56 + CV + DK calculations. To clearly display the details of each PEC, we plotted them only over a small internuclear separation range from approximately 0.11 to 0.50 nm. The dissociation channel of each state is also indicated in the figure.
From the PECs shown in Figure 1 and Figure 2, we can summarize the following features. (1) All the bound states (i.e., X1Σ+, a3Σ+, b3Δ, c3Π, e3Σ, f3Π, 15Σ+, 15Π, and 25Π) have a single well. (2) Two states, namely 15Δ and 15Σ, each have a single well and one barrier. The potential energies of the two barriers are obviously larger than those at their respective dissociation limits. (3) Besides the 15Δ and 15Σ states, the d3Σ+ and 25Σ+ states also have barriers on their PECs. In detail, the d3Σ+ state has a double well and two barriers; the 25Σ+ state has a double well and one barrier, but the second well of the 25Σ+ state is so shallow that it cannot be clearly distinguished. The barriers and the double wells of the d3Σ+ and 25Σ+ states will be discussed in detail later in this paper. (4) The 17Σ+ state is repulsive. (5) The avoided crossings are found in four paired states: the 23Σ+ and 33Σ+ states, the 25Σ+ and 35Σ+ states, the 15Δ and 25Δ states, and the 15Σ and 25Σ states.
To better study the transition probabilities between two electronic states, we present the leading valence electronic configurations of all the bound and quasi-bound states near their respective internuclear equilibrium positions in Table S1 of the Supporting Information section. These were determined by the icMRCI/AV6Z calculations. Due to length limitations, we only tabulated these valence configurations in Table S1 if the coefficients-squared of the configuration-state function (CSF) were larger than 0.1.

3.1. Electron Affinity of SiN and the Stable States of SiN

3.1.1. Electron Affinity of SiN

At the icMRCI + Q/56 + CV + DK level of theory, the AEA and VEA of SiN are computed to be 23,262.27cm−1 and 22,766.75 cm−1, respectively. These correspond to the SiN (X2Σ+, υ = 0) + e → SiN (X1Σ+, υ′′ = 0) and SiN (X2Σ+, υ = 0) + e → SiN (X1Σ+, 0 < υ′′ < 1) transitions, respectively. Our AEA agrees well with the experimental result of 23,785.28 cm1 [10]. Only the theoretical AEA obtained by Kerkines and Mavridis [15] is slightly closer to the experimental result than this one. In addition, the vertical detachment energy (VDE) of SiN was calculated at the icMRCI + Q/56 + CV + DK level of theory as 23,639.13cm1. This corresponds to the SiN (X2Σ+, 0 < υ+ < 1) + e ← SiN (X1Σ+, υ′′ = 0) transition. The computed electron affinities adhere to the expected trend: VEA < AEA < VDE.

3.1.2. Spectroscopic Parameters of the 13 Stable and Metastable Λ-S States of SiN

Employing the PECs determined by the icMRCI + Q/56 + CV + DK calculations, we evaluated the spectroscopic parameters of 13 Λ-S states. For the purposes of this discussion and due to length limitations, the spectroscopic parameters of the 13 Λ-S states are given in Table S2, along with the experimental only [10] and other theoretical [11,12,13,14,15,16] spectroscopic results.
The ground state X1Σ+ of the SiN anion is mainly characterized by the closed-shell electronic configuration 5σ222400 (0.827). A group of experimental work [10] and six groups of calculations [11,12,13,14,15,16] reported the spectroscopic parameters of this state. Values of Re and ωe for this state obtained in this work deviate from the experimental values [10] by only 0.00117 nm and 8.87 cm−1, respectively. Using the values of ωe, ωexe and ωeye obtained here in combination with the equation De = D0 + ωe/2 − ωexe/4 + ωeye/8, we determined the ground state D0 to be approximately 6.2466 eV. Obviously, the present D0 is close to the experimental values [10] as the deviation is only 0.0066 eV. We found, as shown in Table 3, that only the theoretical Re values obtained by Peterson and Woods [11] and Kalcher [13] are slightly closer to the measurements [10] than what we have obtained. This state has a well depth of approximately 50,592.79 cm−1. It has 81 vibrational states, as tabulated in Table S3. Nevertheless, X1Σ+ (υ′′ = 24) lies 23,613.44cm−1 above the X1Σ+ (υ′′ = 0), which is larger than the calculated AEA. This means the X1Σ+ υ′′ ≥ 24 vibrational levels are difficult to observe in a spectroscopy experiment.
The a3Σ+, b3Δ, c3Π, e3Σ, f3Π, 15Σ+, 15Π, and 25Π states also possess the single reference character near the equilibrium position. The dominant electronic configurations of the a3Σ+, b3Δ, e3Σ, and 15Σ+ states arise from the 2π → 3π singlet electronic excitation of the ground state. The dominant electronic configurations of the c3Π and f3Π states are generated predominantly from the 7σ → 3π and 2π → 8σ singlet electronic excitations of the ground state, respectively, whereas the dominant electronic configurations of the 15Π and 25Π states come from the 2π → 3π, 7σ → 3π, and 2π → 3π, 2π → 8σ double electronic excitations of the ground state, respectively.
Two groups of calculations have been performed [13,15], but no measurements on the spectroscopic parameters of the a3Σ+ state had been reported. Compared with the results obtained by Kalcher [13], the present Re and De results are slightly smaller, but the present ωe and Te values are obviously larger. In addition, the present Te value is obviously smaller, but the present Re and De values are larger than the previous results [15]. It should be noted that the previous results [13,15] were calculated using a small basis set [13] and the single reference method [15], whereas the present ones are derived by the icMRCI + Q method with extrapolation to the CBS limit and including various corrections. For this reason, we believe the present results should be more accurate and reliable [13,15]. To date, no experimental or other theoretical spectroscopic parameters have been reported in the literature for the b3Δ, c3Π, e3Σ, f3Π, 15Σ+, 15Π, or 25Π states.
For the a3Σ+ state, the depth of the well is approximately 29,492.12 cm−1. It has 68 vibrational states, as tabulated in Table S3. However, the a3Σ+ (υ′ = 3) lies 23,604.85 cm−1 above the X1Σ+ (υ′′ = 0), and is also larger than the calculated AEA. Hence, the a3Σ+ (υ′ ≥ 3) would be metastable towards autodetachment.
As with X1Σ+ (υ′′ ≥ 24) and a3Σ+ (υ′ ≥ 3), the b3Δ, c3Π, e3Σ, f3Π, 15Σ+, 15Π, and 25Π states are metastable states.
Two-pair avoided crossings are located at approximately R = 0.2284 nm (between the 15Δ and 25Δ states) as well as at approximately R = 0.2604 nm (between the 15Σ and 25Σ states). Therefore, the 15Δ and 15Σ states have a potential well induced by the avoided crossings in the range of R < 0.2284 and R < 0.2604 nm, respectively. They are repulsive at larger internuclear distances. The potential energy at the top of the barrier of each state is higher than that at their respective dissociation limits by approximately 4250.02 and 3750.36 cm−1, respectively. Thus, both the dissociation energy and well depth of each state are relative to their respective barriers and are equal to 6696.83 and 7213.21 cm−1, respectively. The 15Δ and 15Σ states possess 11 and 12 vibrational states, respectively. These vibrational levels are summarized in Table S4. The dominant electronic transitions between the 15Δ and 15Π states and between the 15Σ and 15Π states can be regarded as 3π to 8σ promotions.
The d3Σ+ state has a double well and two barriers. The two barriers lie at approximately 0.1824 and 0.3004 nm, respectively. They are formed by the avoided crossings of the d3Σ+ and 33Σ+ states. The potential energy at the top of the first barrier is lower, whereas that of the second barrier is higher than that at the dissociation limit for the d3Σ+ state. Therefore, the depths of the two wells should be relative to the first barrier, but their dissociation energies must be relative to the second barrier. The depth of the first well is approximately 19,808.90 cm−1. It has 12 vibrational states, as listed in Table S4. The depth of the second well is approximately 1355.04 cm−1. It has three vibrational states, as presented in Table S4.
The 25Σ+ state has a double well and one barrier. The barrier is generated by the avoided crossing of this state with the 35Σ+ state at approximately 0.2604 nm. For the 25Σ+ state, the potential energy at the top of the barrier is lower than that at the dissociation limit. Thus, the depths of the double well should be relative to the barrier, whereas the dissociation energies of the double well must be relative to the dissociation limit. For the double well of the 25Σ+ state, the well depths are 6970.08 and 450.14 cm−1, with 11 and 5 vibrational levels, respectively. These vibrational levels are presented in Table S4.
As with the b3Δ, c3Π, e3Σ, f3Π, 15Σ+, 15Π, and 25Π states, neither theoretical nor experimental studies have been reported in the literature regarding the 15Δ, 15Σ, d3Σ+, and 25Σ+ metastable states.

3.2. Spectroscopic Parameters and Vibrational Levels of the 45 Ω Bound States

When the SOC is taken into account, the Si(4Su) + N(4Su) dissociation channel splits into one dissociation asymptote. Each of the Si(2Du) + N(4Su) and Si(2Pu) + N(4Su) dissociation limits splits into two asymptotes. These Ω states and their dissociation asymptotes are listed in Table 2. In employing the icMRCI + Q/56 + CV + DK + SOC calculations, we determined the energy separations between each higher asymptote and the lowest one, i.e., Si(4Su) + N(4Su). The dissociation relationships for the possible Ω states and corresponding energy separations are also listed in Table 2. Simultaneously, we also collected the corresponding experimental energy separations [19] for comparison. As seen in Table 2, the energy separations obtained in this paper agree favorably with the experimental values [19].
The 14 Λ-S states are split into 49 Ω states. In detail, 10 Ω states (X1Σ+0+, a3Σ+1, a3Σ+0, 15Σ+2, 15Σ+1, 15Σ+0+, 17Σ+0, 17Σ+1, 17Σ+2, and 17Σ+3) arise from the Si(4S3/2) + N(4S3/2) channel, 10 Ω states (c3Π0, c3Π0+, c3Π1, c3Π2, 15Π3, 15Π2, 15Π1, 15Π0+, 15Π0, and 15Π−1) are produced from the Si(2D3/2) + N(4S3/2) channel, 14 Ω states (b3Δ1, b3Δ2, b3Δ3, d3Σ+0, d3Σ+1, 15Δ0, 15Δ0+, 15Δ1, 15Δ2, 15Δ3, 15Δ4, 25Σ+2, 25Σ+1, and 25Σ+0+) are associated with the Si(2D5/2) + N(4S3/2) asymptote, 5 Ω states (25Π−1, 25Π0, 25Π0+, 25Π1, and 25Π2) are yielded from the Si(2P1/2) + N(4S3/2) asymptote, and 10 Ω states (e3Σ0+, e3Σ1, f3Π2, f2Π1, f3Π0+, f3Π0, 25Π3, 15Σ0, 15Σ1, and 15Σ2) belong to the Si(2P3/2) + N(4S3/2) channel. Of these 49 Ω states, only the 17Σ+0, 17Σ+1, 17Σ+2, and 17Σ+3 states are repulsive.
To conveniently discuss the spectroscopic parameters and vibrational levels of these Ω states, we divide them into three types according to their symmetries. The first group comprises 16 Ω states, which arise from the X1Σ+, a3Σ+, d3Σ+, e3Σ, 15Σ+, 25Σ+, and 15Σ states. The second group comprises 20 Ω states, which come from the c3Π, f3Π, 15Π, and 25Π states. The third group comprises nine Ω states, which are generated from the b3Δ and 15Δ states.

3.2.1. Spectroscopic and Vibrational Properties of 16 Ω States with the Σ Symmetry

Using the PECs obtained by the icMRCI + Q/56 + CV + DK + SOC calculations, we evaluated the Te, De, Re, and ωe values of these 16 Ω states with Σ symmetry. The spectroscopic parameters are presented in Table S5. For reasons of discussion, the leading Λ-S state compositions of each Ω state near their respective equilibrium positions are also presented in Table S5. For clarity, we neglected the Λ-S states that contribute less than 0.08% to the total Λ-S state composition.
The X1Σ+ state does not split when the SOC effect is included. The X1Σ+0+ state completely consists of the X1Σ+ state in the FC region. Comparing Re, ωe, and De values collected in Tables S2 and S5, we confirmed that the SOC effect on these spectroscopic parameters can be negligible. Additionally, the vibrational levels of X1Σ+0+ states are almost equal to the corresponding values of the X1Σ+ state.
The SOC interaction causes the a3Σ+ and e3Σ states to split into four Ω states: a3Σ+1, a3Σ+0, e3Σ0+, and e3Σ1, in the order of increasing energy. The a3Σ+ state is inverted, whereas the e3Σ1 state is regular when the SOC effect is taken into account. The SOC effect on the Re, ωe, and De values of these two states is small. This is consistent with the fact that the Λ-S state compositions of each Ω state are almost pure near their respective equilibrium positions. The SOC splitting energies of a3Σ+ and e3Σ states are only 0.87 and 0.88 cm−1, respectively. This is not large. In comparing the Te, Re, ωe, and De values of each Ω state with those seen in Table S2, we affirm that the difference between them is small. This shows that the SOC effect is very small in the spectroscopic parameters of these Ω states. In addition, the SOC effect on their vibrational levels is also small.
Under the SOC effect, the d3Σ+ state is regular and still has a double well. As seen in Table S5, for the double well, the Λ-S state compositions of each Ω state are pure around their respective equilibrium positions. Accordingly, the spectroscopic parameters and vibrational levels of the d3Σ+0 and d3Σ+1 states are almost equal to those of the d3Σ+ state for both the first and second well.
Each of the 15Σ+ and 15Σ states splits into three Ω states with the SOC effect taken into account. As seen in Table S5, the Λ-S state compositions of each Ω state are almost pure around their respective equilibrium positions. Accordingly, the SOC effect on their spectroscopic parameters is small. Through comparison, we confirm that the SOC effect on their vibrational levels is also very small. In addition, the 15Σ+ state is inverted when the SOC effect is included.
The 25Σ+2, 25Σ+1, and 25Σ+0+ states have a double well. The Te of the 25Σ+2 state is smaller than that of the 25Σ+0+ state for both the first and second well. Therefore, the 25Σ+ state is inverted when the SOC effect is taken into account. For the first well, as seen in Table S5, the Λ-S state compositions of each Ω state are pure near the equilibrium positions. This is also true with the 15Σ+ and 15Σ states. The SOC effect on the spectroscopic parameters and vibrational levels of all the Ω states is tiny. For the second well, the Λ-S state compositions of the three Ω states slightly mix with the c3Π and 15Π states near their respective equilibrium positions. The largest deviations of Re, ωe, and De values of each Ω state from those originating from the 25Σ+ state are 0.00019 nm, 1.275 cm−1, and 0.0003 eV, respectively. The SOC splitting energies between the two neighboring Ω states from 25Σ+2 to 25Σ+0+ are only 0.65 and 0.22 cm−1, respectively. Each Ω state has five vibrational levels, which are almost equal to those of the second well of the 25Σ+ state.
In conclusion, (1) the a3Σ+, 15Σ+ and 25Σ+ states are inverted with the SOC effect taken into account. (2) The SOC effect to the spectroscopic parameters and the vibrational levels of all the Σ states is small.

3.2.2. Spectroscopic and Vibrational Properties of 20 Ω States with the Π Symmetry

Using the PECs obtained by the icMRCI + Q/56 + CV + DK + SOC calculations, we evaluated Te, Re, ωe, and De values of the 20 Ω states with the Π symmetry. The spectroscopic parameters are tabulated in Table S6. As with Table S5, also presented the leading Λ-S state compositions of each Ω state around their respective equilibrium positions.
The c3Π state splits into four Ω states (i.e., c3Π0−, c3Π0+, c3Π1, and c3Π2) when the SOC effect is accounted for. The SOC splitting energies between two consecutive Ω states from the c3Π0− to c3Π2 states are 0.44, 47.19, and 47.62 cm−1, respectively. These are relatively large when compared with those of other states. The SOC effect on Re, ωe and De values of each Ω state is not large. The largest deviations of Re, ωe, and De values of all four Ω states of the split c3Π state are 0.00002 nm, 0.274 cm−1, and 0.0059eV, respectively. In comparison, we can confirm that the SOC effect on the vibrational levels of these Ω states is also small.
Under the SOC effect, the f3Π also splits into four Ω states (f3Π0−, f3Π0+, f3Π1, and f3Π2). Their energies increase in the order of f3Π0−, f3Π0+, f3Π1, and f3Π2 in the range of R < 0.1644 nm. However, the order of f3Π0−, f3Π0+, f3Π1, and f3Π2 changes in the internuclear distance region from 0.1644 to 1.08 nm. This phenomenon leads to the obvious change of the PECs for the f3Π0, f3Π0+, and f3Π2 states. As a result, the spectroscopic parameters and the vibrational levels of f3Π0, f3Π0+, and f3Π2 states are obviously different from those of the corresponding Λ-S state.
Each of the 15Π and 25Π states splits into six Ω states with the SOC effect taken into account, and the 15Π state is inverted. For the 15Π state, the largest deviations of Re, ωe, and De of each Ω state from those of the 15Π state are 0.00001 nm, 0.298 cm−1, and 0.0047eV, respectively; the SOC splitting energies between the two neighboring Ω states from the 15Π3 to the 15Π−1 are only 17.99, 18.22, 18.31, 0.13, and 18.65 cm−1, respectively; on the whole, the SOC effect on the splitting energies is not large; each Ω state has 38 vibrational levels, which are almost equal to those of the first well of 15Π state. For the 25Π state, the Λ-S state compositions of each Ω state are pure near the equilibrium positions; as with the c3Π and 15Π states, the SOC effect on Re, ωe and De of all the Ω states is tiny; each Ω state has the 36 vibrational states, in which vibrational levels are almost equal to those of the 25Π state; the energy splitting of 25Π−1 to 25Π0, 25Π0 to 25Π0+, 25Π0+ to 25Π1, 25Π1 to 25Π2, and 25Π2 to 25Π3 are 22.17, 0.22, 21.95, 22.16, and 22.17cm−1, respectively, which are not large; thus, we affirm that the SOC effect on the spectroscopic and vibrational properties of all the 6 Ω states is very small.
Our conclusions are as follows: (1) the SOC effect on the spectroscopic parameters is tiny except for the f3Π state and the spitting energies of c3Π state; (2) the 15Π state is inverted, and the f3Π state is also inverted in the internuclear distance range from 0.1644 to 1.08 nm with the SOC effect taken into account; (3) the SOC effect on the vibrational levels of all the Π states is small expect for the f3Π0, f3Π0+, and f3Π2 states.

3.2.3. Spectroscopic and Vibrational Properties of Nine Ω States with the Δ Symmetry

Using the PECs obtained by the icMRCI + Q/56 + CV + DK + SOC calculations, we evaluated the Te, De, Re, and ωe values of nine Ω states with the Δ symmetry. The spectroscopic parameters are presented in Table S7. As with Tables S5 and S6, the leading Λ-S state compositions of each Ω state around their respective internuclear equilibrium positions are also tabulated in Table S7.
The b3Δ state splits into three Ω states when the SOC effect is taken into account. For the three Ω states, the energy arrangement from low to high is b3Δ3, b3Δ2, and b3Δ1. The Te, Re, ωe, and De values of each Ω state are very close to those of the b3Δ state. Each Ω state has 70 vibrational states, the vibrational levels of which are almost equal to those of the b3Δ state. As a result, we conclude that the SOC effect on the spectroscopic parameters and vibrational states is tiny.
The 15Δ state splits into the 15Δ0, 15Δ0+, 15Δ1, 15Δ2, 15Δ3, and 15Δ4 states when the SOC effect is included, as tabulated in Table S7. The Λ-S state compositions of each Ω state are almost pure around the internuclear equilibrium positions. Consequently, the SOC effect on the spectroscopic parameters and vibrational levels of each Ω state are inconspicuous.
In conclusion, (1) the SOC effect on the spectroscopic parameters and vibrational properties of the b3Δ and 15Δ states is very small, and (2) the b3Δ state is inverted.

3.3. Transition Properties

The TDMs between the a3Σ+1 and X1Σ+0+ states were obtained with the Breit-Pauli Hamiltonian in combination with the all-electron CVTZ basis set at the level of the icMRCI theory. The curves of TDM versus internuclear separation are depicted in Figure 3. As with the PECs, to clearly show the main features of each TDM curve, we display them over a small range of internuclear separations. The figure clearly shows that the TDMs of the transitions from the a3Σ+1 state to the X1Σ+0+ state very small in the FC region. This is consistent with the fact that the triplet-singlet transitions are forbidden. In addition, the TDMs go to the zero asymptote when the internuclear distance is larger than 0.4 nm.
Using the PECs and TDMs obtained here, we calculated the FC factors and Einstein coefficients of the a3Σ+1 (υ′ = 0–2) to X1Σ+0+ with the LEVEL program [32]. We only list the lowest 10 vibrational levels of the X1Σ+0+ state due to the length limitation. For the purposes of this discussion, we present these results in Table 3. Table 4 lists the radiative lifetimes for the a3Σ+1 (υ′ = 0–2) to X1Σ+0+ transition. Overall, the a3Σ+1 states are not easy to detect spectroscopically by observing these transitions. This can be explained based on the data in Table 3 and Table 4. As presented in Table 3 and Table 4, almost all of the Einstein coefficients of the a3Σ+1 to X1Σ+0+ transition are very small and their radiative lifetimes are very long. That is, these transitions are weak.

4. Conclusions

In this work, we calculated the PECs of 14 Λ-S states of the SiN anion using the icMRCI + Q/56 + CV + DK approach, computed the TDMs of the a3Σ+1 to X1Σ+0+ transition using the icMRCI approach with the all-electron CVTZ basis set, and determined the spectroscopic parameters of 49 Ω states employing the icMRCI + Q/56 + CV + DK + SOC method. The spectroscopic parameters and vibrational levels were evaluated. In addition, the transition probabilities of a3Σ+1 to X1Σ+0+ were studied. These results were compared in detail with those reported in the literature. Excellent agreement is found between these results and those of other measurements. Of these 14 states, only the 17Σ+ state is repulsive. The avoided crossings exist between the d3Σ+ and 33Σ+ states, between the 25Σ+ and 35Σ+ states, between the 15Δ and 25Δ states, and between the 15Σ and 25Σ states. The X1Σ+ (υ′′ ≥ 24), a3Σ+ (υ′ ≥ 3), b3Δ, c3Π, d3Σ+, e3Σ, f3Π, 15Σ+, 15Π, 25Σ+, 25Π, 15Δ, and 15Σ states are metastable states. The d3Σ+, b3Δ, f3Π, 15Σ+, 15Π, and 25Σ states are inverted when the SOC effect is included. The SOC effect on the spectroscopic parameters and vibrational levels is small except for in the f3Π state and the spitting energies of c3Π state. The spectroscopic parameters, vibrational levels, and transition probabilities obtained in this paper can be considered very reliable and can be employed as helpful guidelines for detecting these states in an appropriate spectroscopy experiment in the near future. In addition, the data provided in this work should assist in spectral searches for this anion in the interstellar medium.

Supplementary Materials

The supplementary materials are available online.

Acknowledgments

This work was sponsored by the National Natural Science Foundation of China under Grant No. 11274097 and the Program for Science and Technology of Henan Province in China under Grant No. 142300410201.

Author Contributions

Wei Xing computed the data and wrote the manuscript. Deheng Shi and Zunlue Zhu analyzed the data. Jinfeng Sun revised the manuscript and conceived and designed the project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Agúndez, M.; Cernicharo, J.; Guélin, M.; Kahane, C.; Roueff, E.; Kłos, J.; Aoiz, F.J.; Lique, F.; Marcelino, N.; Goicoechea, J.R.; et al. Astronomical identification of CN, the smallest observed molecular anion. Astron. Astrophys. 2010, 517, L2. [Google Scholar] [CrossRef]
  2. Turner, B.E. Detection of SiN in IRC + 10216. Astrophys. J. 1992, 388, L35–L38. [Google Scholar] [CrossRef]
  3. Guèlin, M.; Muller, S.; Cernicharo, J.; Apponi, A.J.; McCarthy, M.C.; Gottlieb, C.A.; Thaddeus, P. Astronomical detection of the free radical SiCN. Astrophys. J. 2000, 363, L9–L12. [Google Scholar]
  4. Guèlin, M.; Muller, S.; Cernicharo, J.; McCarthy, M.C.; Thaddeus, P. Detection of the SiNC Radical in IRC + 10216. Astron. Astrophys. 2004, 426, L49–L52. [Google Scholar] [CrossRef]
  5. Fortenberry, R.C.; Daniel Crawford, T. Theoretical prediction of new dipole-bound singlet states for anions of interstellar interest. J. Chem. Phys. 2011, 134, 154304. [Google Scholar] [CrossRef] [PubMed]
  6. Drzaic, P.S.; Marks, J.; Brauman, J.I. Gas-Phase Ion Chemistry; Academic Press: New York, NY, USA, 1984; pp. 1–63. [Google Scholar]
  7. Damrauer, R.; Hankin, J.A. Chemistry and thermochemistry of silicon-containing anions in the gas phase. Chem. Rev. 1995, 95, 1137–1160. [Google Scholar] [CrossRef]
  8. Zhu, W.; Kochanski, G.P.; Jin, S. Low-field electron emission from undoped nanostructured diamond. Science 1998, 282, 1471–1473. [Google Scholar] [CrossRef] [PubMed]
  9. Shabanov, S.V.; Gornushkin, I.B. Anions in laser-induced plasmas. Appl. Phys. A 2016, 122, 676. [Google Scholar] [CrossRef]
  10. Meloni, G.; Sheehan, S.M.; Ferguson, M.J.; Neumark, D.M. Negative ion photoelectron spectroscopy of SiN. J. Phys. Chem. A 2004, 108, 9750–9754. [Google Scholar] [CrossRef]
  11. Peterson, K.A.; Woods, R.C. Ground state spectroscopic and thermodynamic properties of AlO, SiN, CP, BS, BO, and CN from Møller-Plesset perturbation theory. J. Chem. Phys. 1989, 90, 7239–7250. [Google Scholar] [CrossRef]
  12. McLean, A.D.; Liu, B.; Chandler, G.S. Computed self-consistent field and singles and doubles configuration interaction spectroscopic data and dissociation energies for the diatomics B2, C2, N2, O2, F2, CN, CP, CS, PN, SiC, SiN, SiO, SiP, and their ions. J. Chem. Phys. 1992, 97, 8459–8464. [Google Scholar] [CrossRef]
  13. Kalcher, J. Trends in ground and excited state electron affinities of group 14, 15, and 16 mixed diatomic anions: A computational study. Phys. Chem. Chem. Phys. 2002, 4, 3311–3317. [Google Scholar] [CrossRef]
  14. Midda, S.; Das, A.K. Theoretical study of spectroscopic constants and molecular properties of diatomic anions using B3LYP method. J. Mol. Struct. 2003, 640, 183–189. [Google Scholar] [CrossRef]
  15. Kerkines, I.S.K.; Mavridis, A. On the electron affinity of SiN and spectroscopic constants of SiN. J. Chem. Phys. 2005, 123, 124301. [Google Scholar] [CrossRef] [PubMed]
  16. Mogren, M.M.A.; El-Azhary, A.A.; Alkiali, W.Z.; Hochlaf, M. Electronic structure and properties of neutral, anionic and cationic silicon-nitrogen nanoclusters. J. Mol. Model. 2013, 19, 2657–2668. [Google Scholar] [CrossRef] [PubMed]
  17. Luo, W.; Li, R.; Gai, Z.Q.; Ai, R.B.; Zhang, H.M.; Zhang, X.M.; Yan, B. Configuration interaction studies on the spectroscopic properties of PbO including spin-Orbit coupling. Chin. Phys. B 2016, 25, 073101. [Google Scholar] [CrossRef]
  18. Zhao, S.T.; Yan, B.; Li, R.; Wu, S.; Wang, Q.L. MRCI + Q study of the low-lying electronic states of CdF including spin-orbit coupling. Chin. Phys. B 2017, 26, 023105. [Google Scholar] [CrossRef]
  19. Andersen, T.; Haugen, H.K.; Hotop, H. Binding energies in atomic negative ions: III. J. Phys. Chem. Ref. Data 1999, 28, 1511–1533. [Google Scholar] [CrossRef]
  20. Moore, C.E. CRC Series in Evaluated Data in Atomic Physics; CRC Press: Boca Raton, FL, USA, 1993; p. 339. [Google Scholar]
  21. Langhoff, S.R.; Davidson, E.R. Configuration interaction calculations on the nitrogen molecule. Int. J. Quantum. Chem. 1974, 8, 61–72. [Google Scholar] [CrossRef]
  22. Richartz, A.; Buenker, R.J.; Peyerimhoff, S.D. Ab initio MRD-CI study of ethane: The 14–25 eV PES region and Rydberg states of positive ions. Chem. Phys. 1978, 28, 305–312. [Google Scholar] [CrossRef]
  23. Van Mourik, T.; Wilson, A.K.; Dunning, T.H. Benchmark calculations with correlated molecular wavefunctions. XIII. Potential energy curves for He2, Ne2 and Ar2 using correlation consistent basis sets through augmented sextuple zeta. Mol. Phys. 1999, 99, 529–547. [Google Scholar] [CrossRef]
  24. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  25. Werner, H.-J.; Knowles, P.J.; Lindh, R.; Knizia, G.; Manby, F.R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; et al. MOLPRO 2010.1 is a Package of Ab Initio Programs. Available online: http://www.molpro.net (accessed on 17 September 2010).
  26. De Jong, W.A.; Harrison, R.J.; Dixon, D.A. Parallel Douglas-Kroll energy and gradients in NWChem: Estimating scalar relativistic effects using Douglas-Kroll contracted basis sets. J. Chem. Phys. 2001, 114, 48–53. [Google Scholar] [CrossRef]
  27. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. V. Core-valence basis sets for boron through neon. J. Chem. Phys. 1995, 103, 4572–4585. [Google Scholar] [CrossRef]
  28. Berning, A.; Schweizer, M.; Werner, H.-J.; Knowles, P.J.; Palmieri, P. Spin-orbit matrix elements for internally contracted multireference configuration interaction wavefunctions. Mol. Phys. 2000, 98, 1823–1833. [Google Scholar]
  29. Oyeyemi, V.B.; Krisiloff, D.B.; Keith, J.A.; Libisch, F.; Pavone, M.; Carter, E.A. Size-extensivity-corrected multireference configuration interaction schemes to accurately predict bond dissociation energies of oxygenated hydrocarbons. J. Chem. Phys. 2014, 140, 044317. [Google Scholar] [CrossRef] [PubMed]
  30. Xing, W.; Shi, D.H.; Sun, J.F.; Zhu, Z.L. Calculations of 21 Λ-S and 42 Ω states of BC molecule: Potential energy curves, spectroscopic parameters and spin-orbit coupling effect. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 153, 722–734. [Google Scholar] [CrossRef] [PubMed]
  31. Xing, W.; Shi, D.H.; Sun, J.F.; Zhu, Z.L. Accurate spectroscopic calculations of the 19 Λ-S states and 36 Ω states of the BC+ cation. Mol. Phys. 2017, 115, 387–402. [Google Scholar] [CrossRef]
  32. Le Roy, R.J. LEVEL 8.0: A Computer Program for Solving the Radial Schrödinger Equation for Bound and Quasi-Bound Levels; University of Waterloo Chemical Physics Research Report CP-663; University of Waterloo: Waterloo, ON, Canada, 2007. [Google Scholar]
Sample Availability: Not Available.
Figure 1. PECs of the 8 Λ-S states with the Σ symmetry of the SiN anion.
Figure 1. PECs of the 8 Λ-S states with the Σ symmetry of the SiN anion.
Molecules 23 00210 g001
Figure 2. PECs of the 6 Λ-S states with the Π and Δ symmetries of the SiN anion.
Figure 2. PECs of the 6 Λ-S states with the Π and Δ symmetries of the SiN anion.
Molecules 23 00210 g002
Figure 3. TDMs versus internuclear separations of transitions from the a3Σ+1 state to the X1Σ+0+ state.
Figure 3. TDMs versus internuclear separations of transitions from the a3Σ+1 state to the X1Σ+0+ state.
Molecules 23 00210 g003
Table 1. Dissociation relationships of the 14 states generated from the first three dissociation channels of the SiN anion.
Table 1. Dissociation relationships of the 14 states generated from the first three dissociation channels of the SiN anion.
Dissociation ChannelElectronic StateRelative Energy/cm−1
This Work aExp. [19]
Si (4Su) + N(4Su)X1Σ+, a3Σ+, 15Σ+, 17Σ+0.00.0
Si (2Du) + N(4Su)b3Δ, c3Π, d3Σ+, 25Σ+, 15Π, 15Δ6973.486961.85 b
Si (2Pu) + N(4Su)e3Σ, 15Σ, f3Π, 25Π10,974.90(10,977.24 + x c)/2
a Obtained using the icMRCI + Q/56 + CV + DK calculations; b obtained by averaging the atomic energy levels of the 2D3/2 and 2D5/2 states; c no experimental energy level of 2P3/2 state obtained by Andersen et al. [19].
Table 2. Dissociation relationships of possible Ω states yielded from the first three dissociation channels of the SiN anion.
Table 2. Dissociation relationships of possible Ω states yielded from the first three dissociation channels of the SiN anion.
Atomic StatePossible Ω StatesRelative Energy/cm−1
This Work aExp. [19]
Si (4S3/2) + N(4S3/2)0 (2), 0+(2), 1(3), 2(2), 30.000.00
Si (2D3/2) + N(4S3/2)−1, 0(2), 0+(2), 1(2), 2(2), 36968.326954.81
Si (2D5/2) + N(4S3/2)0(2), 0+(2), 1(4), 2(3), 3(2), 46978.636968.89
Si (2P1/2) + N(4S3/2)−1, 0, 0+, 1, 210,970.0810,977.24
Si (2P3/2) + N(4S3/2)0(2), 0+(2), 1(3), 2(2), 310,979.76x b
a Obtained by the icMRCI + Q/56 + CV + DK + SOC calculations; b no experimental energy level of 2P3/2 state obtained by Andersen et al. [19].
Table 3. FC factors (1st line) and Einstein coefficients (s−1, 2nd line) for the a3Σ+1 to X1Σ+0+ transition.
Table 3. FC factors (1st line) and Einstein coefficients (s−1, 2nd line) for the a3Σ+1 to X1Σ+0+ transition.
State to Stateυ′′ = 0υ′′ = 1υ′′ = 2υ′′ = 3υ′′ = 4υ′′ = 5υ′′ = 6υ′′ = 7υ′′ = 8υ′′ = 9
a3Σ+1 to X1Σ+0+
υ′ = 00.09700.23930.27960.20680.10990.04570.01590.00470.00110.0002
0.33700.71360.71290.45060.20540.07360.02200.00560.00120.0002
υ′ = 10.18860.16660.01310.04550.16630.19290.13280.06440.02300.0057
0.72450.54930.03590.11400.35190.34840.20640.08700.02740.0062
υ′ = 20.23730.02620.07090.13620.02070.02890.13780.16610.11010.0479
1.00720.09540.22580.36920.04620.06130.24480.25440.14680.0556
Table 4. Radiative lifetime values of the transition from the a3Σ+1 (υ′= 0–2) excited Ω states to the X1Σ+0+ state for the SiN anion.
Table 4. Radiative lifetime values of the transition from the a3Σ+1 (υ′= 0–2) excited Ω states to the X1Σ+0+ state for the SiN anion.
Radiative Lifetimes
Transitionsυ′ = 0υ′ = 1υ′ = 2
a3Σ+1 to X1Σ+0+ (ms)396.5407.9396.4

Share and Cite

MDPI and ACS Style

Xing, W.; Sun, J.; Shi, D.; Zhu, Z. icMRCI+Q Study of the Spectroscopic Properties of the 14 Λ-S and 49 Ω States of the SiN Anion in the Gas Phase. Molecules 2018, 23, 210. https://doi.org/10.3390/molecules23010210

AMA Style

Xing W, Sun J, Shi D, Zhu Z. icMRCI+Q Study of the Spectroscopic Properties of the 14 Λ-S and 49 Ω States of the SiN Anion in the Gas Phase. Molecules. 2018; 23(1):210. https://doi.org/10.3390/molecules23010210

Chicago/Turabian Style

Xing, Wei, Jinfeng Sun, Deheng Shi, and Zunlue Zhu. 2018. "icMRCI+Q Study of the Spectroscopic Properties of the 14 Λ-S and 49 Ω States of the SiN Anion in the Gas Phase" Molecules 23, no. 1: 210. https://doi.org/10.3390/molecules23010210

Article Metrics

Back to TopTop