Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Ab initio electronic structures and total internal partition sums of FeH+/2+

Isuru R. Ariyarathna *, Jeffery A. Leiding , Amanda J. Neukirch and Mark C. Zammit
Physics and Chemistry of Materials (T-1), Los Alamos National Laboratory, Los Alamos, NM 87545, USA. E-mail: isuru@lanl.gov

Received 22nd August 2024 , Accepted 3rd December 2024

First published on 4th December 2024


Abstract

In the present work, we studied 27 FeH+ and 6 FeH2+ electronic states using multireference configuration interaction (MRCI), Davidson-corrected MRCI (MRCI+Q), and coupled cluster singles doubles and perturbative triples [CCSD(T)] wavefunction theory (WFT) calculations conjoined with large quadruple-ζ and quintuple-ζ quality correlation consistent basis sets. We report their potential energy curves (PEC), energy related properties, spectroscopic parameters, and spin–orbit couplings. Dipole moment curves (DMC) and transition dipole moment curves (TDMC) of several low-lying electronic states of FeH+ and FeH2+ are also introduced. The ground state of FeH+ is a single-reference X5Δ (6σ2123) with an adiabatic D0 of ∼52 kcal mol−1, which is in agreement with the experimental value. The states with the largest adiabatic binding energies of FeH2+ (4Π and 4Δ) are multireference in nature with an approximate D0 of 22 kcal mol−1. We used CCSD(T) μ of the FeH+(X5Δ) to assess the density functional theory (DFT) errors associated with a series of functionals that span multiple rungs of Jacob's ladder of density functional approximation (DFA) and observed a general trend of improving μ when moving to more expensive functionals at the higher rungs. We expect weak spectral bands to be produced from the low-lying electronic states of FeH2+ and FeH+ due to their lower transition μ values. Lastly, we present results for the total internal partition function sums (TIPS) of FeH+ and FeH2+, which have not been presented in the literature before.


I. Introduction

The diatomic iron hydride cation FeH+ is predicted to be abundant in cool stellar atmospheres.1 However, due to the deficiency of available laboratory spectroscopic data on FeH+, its astronomical presence is yet be observed. Aiming to guide and motivate experimental analysis of FeH+ so far, a series of theoretical and computational attempts have been made specifically for gaining insight into its spectral features.

The first observation of the FeH+ goes back to the 1979 Mysov et al.'s mass spectroscopic fragment analysis study of (CH3C5H4)2Fe.2 Five years later, Halle, Klein, and Beauchamp analyzed the thresholds of the Fe+ + H2 and Fe+ + D2 reactions using ion beam apparatus and obtained the D0 of FeH+ (59 ± 5 kcal mol−1).3 In 1986, Schilling et al.4 performed an ab initio generalized valence bond plus configuration interaction study and assigned a 5Δ ground state with a 47.0 kcal mol−1D0 to FeH+ which is significantly lower than the experimental value obtained by Halle et al.3 In the same year, Elkind and Armentrout carried out a guided ion beam mass spectrometric study and reported a D0 of 48.9 ± 1.4 kcal mol−1 for FeH+[thin space (1/6-em)]5 which is in agreement with Schilling et al.'s work.4 Furthermore, they intuitively projected low-lying 5Π and 5Σ+ electronic states for FeH+ with σ3π3δ2 and σ4π2δ2 electronic configurations, respectively. In 1987 Schilling et al., conducted another theoretical study and provided theoretical evidence for the existence of the 5Π and 5Σ+ excited states for FeH+ lying 2.1 and 10.0 kcal mol−1 above.6 In the same year, Lars et al., carried out modified coupled-pair functional (MCPF) calculations to predict re (1.603 Å), ωe (1805 cm−1), and μ (2.41 D) values of FeH+(5Δ).7 They further estimated the D0 of FeH+ to be 52.3 kcal mol−1, which is 2 kcal mol−1 greater than the upper bound of the D0 reported by Elkind and Armentrout.5 Two years later, Sodupe, Lluch, and Oliva studied the PEC originating from the Fe+(6D) + H(2S) fragments using the restricted open Hartree–Fock (ROHF) and configuration interaction levels.8 In line with the previous reports, their ROHF calculations predicted a 5Δ ground state for FeH+. However, they found that the inclusion of the electron correlation leads to a 5Π ground state for FeH+. According to their potential energy profile, all the septet-spin electronic states originating from the ground state fragments are repulsive in nature. This observation was further corroborated by a study reported by Langhoff and Bauschlicher in 1991.9 Specifically, Langhoff and Bauschlicher carried out a theoretical spectroscopic study for FeH+ utilizing CASSCF (complete active space self-consistent field), MRCI, and MCPF levels of theory.9 The CASSCF order of the states that they observed was X5Δ, A5Π, B5Σ+, a3Σ, b3Φ, c3Π, and d3Δ. They reported the MRCI X5Δ → A5Π and X5Δ → a3Σ transition energies of FeH+ to be 669 and 10[thin space (1/6-em)]277 cm−1, respectively. Moreover, their study predicted a D0 of 50.2 kcal mol−1 for FeH+. In 2019, Cheng and DeYonker analyzed the low-lying X5Δ, A5Π, B5Σ+, a3Σ, b3Φ, c3Π, and d3Δ states of FeH+ using MRCI and coupled cluster levels of theories.10 This is clearly the most complete work reported for FeH+ so far. Their work utilized a highly accurate focal point approach to calculate the X5Δ → A5Π (600 cm−1) and X5Δ → a3Σ (10[thin space (1/6-em)]081 cm−1) transition energies and ionization energy (IE) of FeH (7.4851 eV).10 Furthermore, this work reported a series of spectroscopic constants for the aforementioned states. The most recent study related to FeH+ was reported in 2022 by the Beyer group.11 Here they performed infrared multiple photon dissociation (IRMPD) spectroscopic analysis for Ar2FeH+ aiming to guide future experimental spectroscopic studies of FeH+.

To the best of our knowledge, experimental spectroscopic analysis has not been conducted for FeH2+ before. We were only able to locate one WFT based study for this system which was reported by Wilson, Marsden, Nagy-Felsobuki in 2003.12 This study predicted a 4Δ ground state for FeH2+ with De (dissociation energy), re, and ωe values of 21.68 kcal mol−1, 1.998 Å, and 830 cm−1, respectively under the MRCI+Q level of theory.

In the present work we have utilized ab initio MRCI13–15 method and MRCI+Q16 correction to analyze the Fe+ + H and Fe2+ + H reactions and to investigate the ground and electronically excited states of FeH+ and FeH2+ species. The implemented MRCI is indeed capable of providing accurate results for both multireference and single-reference electronic states of highly correlated transition metal-based species such as FeH+ and FeH2+. Especially, this level of theory is ideal for efficiently producing full PEC for a large number of electronic states of diatomic molecules. On the other hand, the approximate quadruple substitution effect provided by MRCI+Q16 is often being used to gain more accurate results and reach experimental observations.

Here, we report 27 and 6 MRCI PEC of FeH+ and FeH2+, respectively. Under MRCI,13–15 MRCI+Q,16 and CCSD(T)17 levels, their equilibrium electronic configurations, various energy related properties, and a set of spectroscopic parameters are reported. At the MRCI level, the spin–orbit effects of FeH+ and FeH2+ were also evaluated. Furthermore, MRCI DMC and TDMC corresponding to several low-lying electronic states of FeH+ and FeH2+ are introduced. The CCSD(T) μ analyses were also performed for low-lying single-reference electronic states. The μ of the single-reference FeH+(X5Δ) were also analyzed with 17 functionals that span multiple families of DFA18 [i.e., semi-local generalized gradient approximation (GGA), meta-GGA (MGGA), global GGA hybrid, MGGA hybrid, range-separated hybrid (RSH), double hybrid (DH)]. Finally, we used the MRCI PEC of FeH+ and FeH2+ to calculate their TIPS values. We believe that the findings of this work will serve as a guide for future theoretical studies of similar transition metal-based diatomic species and for experimental analysis and identification of FeH+ and FeH2+ in the interstellar space.

II. Computational details

The internally contracted MRCI (≡ MRCISD)13–15 and CCSD(T)17 calculations were performed using the MOLPRO 2023.219–21 code. First, full MRCI PEC of FeH+ and FeH2+ were produced using the quadruple-ζ quality correlation consistent aug-cc-pVQZ basis set for both atoms.22,23 Specifically, the PEC arising from Fe+(6D) + H(2S), Fe+(4F) + H(2S), Fe+(4D) + H(2S), Fe+(4P) + H(2S), and Fe+(2G) + H(2S) were considered to study the low-lying electronic states of FeH+. On the other hand, the PEC of Fe2+(5D) + H(2S) were analyzed to investigate the electronic states of FeH2+. CASSCF24–27 wavefunctions were provided for all MRCI calculations. The CASSCF wave function of FeH+ were produced by allocating 8 electrons in 12 orbitals [CAS(8,12)]. At the dissociation limit, the active orbitals are the pure atomic orbitals of the five 3d, five 4d, and 4s of Fe and the 1s of H. The same set of orbitals were provided to build CASSCF active space of FeH2+ with 7 active electrons [CAS(7,12)]. Under the utilized C2v Abelian point-group, the active orbitals are 6a1 (3dz2, 3dx2y2, 4dz2, 4dx2y2, and 4s of Fe and 1s of H), 2b1 (3dxz and 4dxz of Fe), 2b2 (3dyz and 4dyz of Fe), and 2a2 (3dxy and 4dxy of Fe). At the MRCI level, the single and double electron substitutions from active to virtual orbitals were allowed. The ad hoc Davidson correction (MRCI+Q) was applied as a size-extensivity correction.16 The MRCI and MRCI+Q potential wells of FeH+ and FeH2+ were utilized to solve the rovibrational Schrödinger equation numerically and obtain their harmonic vibrational frequencies (ωe), anharmonicities (ωexe), equilibrium rotational constants (Be), anharmonic correction to the rotational constants (αe), and centrifugal distortion constants ([D with combining macron]e). At the MRCI level, DMC and TDMC of several low-lying electronic states of FeH+ and FeH2+ were also produced using the same basis set and active spaces. To evaluate the relativistic effects on the energetics and the spectroscopic constants, potential energy scans were performed at the MRCI and MRCI+Q levels for the 7 and 6 lowest electronic states of FeH+ and FeH2+, respectively. For these calculations, the aug-cc-pVQZ-DK basis set and the second-order Douglas–Kroll–Hess Hamiltonian were used. Hereafter these calculations are denoted by AQZ-DK-MRCI and AQZ-DK-MRCI+Q. At the AQZ-DK-MRCI level, spin–orbit coupling effects were evaluated by implementing the Breit-Pauli Hamiltonian (more information on the spin–orbit analysis is given in the Discussion section).

The CCSD(T) calculations were performed for a few single-reference electronic states of FeH+ and FeH2+ utilizing the restricted Hartree–Fock (RHF) wave functions. For CCSD(T) calculations, 3 types of correlation consistent basis sets were used: (1) aug-cc-pVQZ/Fe,H, (2) aug-cc-pV5Z/Fe,H, (3) aug-cc-pwCV5Z/Fe aug-cc-pV5Z/H.22,23 Hereafter these CCSD(T) calculations are denoted by AQZ-CCSD(T), A5Z-CCSD(T), and c-A5Z-CCSD(T), respectively. Note that the 3s2 and 3p6 core electrons of Fe were correlated at the c-A5Z-CCSD(T).22 The CCSD(T) potential energy scans performed around the equilibrium distance region of the electronic states were utilized to calculate their De, re, ωe, ωexe, Be, αe, and [D with combining macron]e values. Combined with the findings from a previous work of ours, the IE of FeH was calculated at the CCSD(T) level.28 The CCSD(T) μ values of the low-lying single-reference electronic states of FeH+ and FeH2+ were also computed with the finite-field method. For these CCSD(T) μ calculations, a field (f) of 0.01 a.u. was applied to the positive and negative directions of FeH+ and FeH2+ and the calculated E(f) and E(−f) energies were provided for the μ = [E(f) − E(−f)]/2f equation. The μ of the single-reference FeH+(X5Δ) was also calculated with DFT using a series of functionals belongs to different families of DFA; semi-local generalized gradient approximation (GGA: BP86,29,30 BLYP,31,32 PBE33), meta-GGA (MGGA: TPSS,34 MN15-L35), global GGA hybrid (B3LYP,36–38 B3P86,29,36 B3PW91,36,39 PBE040), MGGA hybrid (TPSSh,34 M06-2X,41 MN1542), range-separated hybrid (RSH: LRC-ωPBE,43 CAM-B3LYP,44 ωB97X45), and double hybrid (DH: PBE0-DH,46 DSDPBEP8647,48). The DFT μ values were calculated at the previously reported DFT re values of FeH+(X5Δ) combined with the aug-cc-pVQZ basis set.28 In all cases, the default origins (center of the mass) were used for the dipole moment calculations. For DFT calculations, Gaussian 1649 software was used.

The TIPS, Q, of a species can be evaluated via

image file: d4cp03296a-t1.tif
where gj is the degeneracy or statistical weight (including the nuclear spin) of level j, kB is the Boltzmann constant, T is the temperature, and Ej,0 is the excitation energy of level j from the ground state. For the evaluation of the FeH+ and FeH2+ TIPS, rovibrational energy levels were calculated by solving the rovibrational Schrödinger equation,50 where all bound rovibrational levels50 within the 27 electronic states of FeH+ and 6 electronic states of FeH2+ were included in the sum.

III. Results and discussion

III.A. FeH+

The implementation of large active spaces that include additional diffuse type d-functions, rather than traditional valence orbitals-based spaces has been tested to provide more accurate energetics and electronic structures for group 8 transition metal-based diatomic systems (i.e., FeH, FeO2+, RuO2+).28,51,52 Similarly, in the current study we have utilized large active spaces that are made of the five 3d, five 4d, and the 4s, atomic orbitals of Fe and the 1s atomic orbital of H for all of our multireference calculations. Using this active space, first we studied several CASSCF PEC of FeH+ that are originating from various asymptotes of Fe+ and H.

The ground state of Fe+ is a 6D with 3d64s1 valence electron configuration.53 The transfer of the 4s1 electron to 3d orbitals produces its first excited state 4F (at ∼5.35–8.91 kcal mol−1).53 The second excited state of Fe+ (4D) has an electron arrangement similar to the ground state, but with a lower spin due to the electron coupling (4D; at 22.75–25.29 kcal mol−1).53 The next state of Fe+ is a 4P with 3d7 configuration (at 38.53–39.76 kcal mol−1).53 The fourth excited state of Fe+ is indeed the first doublet-spin electronic state of Fe+ (2G; 3d7) lying at 45.30–46.80 kcal mol−1.53 The interactions of all these electronic states of Fe+ with the H(2S) ground state were selected to investigate the low-lying electronic states of FeH+. The reactions between Fe+(6D) + H(2S), Fe+(4F) + H(2S), Fe+(4D) + H(2S), Fe+(4P) + H(2S), and Fe+(2G) + H(2S) give rise to 7,5+, Π, Δ], 5,3, Π, Δ, Φ], 5,3+, Π, Δ], 5,3, Π], and 3,1+, Π, Δ, Φ, Γ] states.54,55 We used the CASSCF PEC of the aforementioned asymptotes to identify the lowest energy electronic states of FeH+; specifically its most stable 27 states were studied under the MRCI level of theory. Our calculated MRCI potential energy profile is given in the Fig. 1. Since we did not consider the interaction of excited states of H with the low-lying electronic states of Fe+ (due to the excitation energy of H atom being significantly higher), the energies of the fragments at the right side of the potential energy profile correspond to the excitation energies of Fe+. As expected, at the dissociation limit, the first 4 excitation energies of Fe+ at the MRCI level are ∼4, 22, 37, and 46 kcal mol−1, which are in reasonable agreement with the experimental values.53


image file: d4cp03296a-f1.tif
Fig. 1 MRCI/aug-cc-pVQZ PEC of FeH+ as a function of Fe+⋯H distance [r(Fe+⋯H), Å]. The relative energies are with respect to the total energy of Fe+(6D)⋯H(2S) when they are at 200 Å separation, which is set to 0 kcal mol−1. The dotted, solid, and dashed PEC correspond to the quintet, triplet, and singlet spins, respectively. The Δ, Π, Σ+, Σ, Φ, Γ, and H states are shown in red, blue, pink, green, black, cyan, and brown, respectively.

All 3 quintet-spin electronic states originating from the Fe+(6D) + H(2S) are strongly attractive and become the first 3 electronic states of FeH+ (i.e., X5Δ, A5Π, and B5Σ+). Note that within the studied energy range, the septet-spin PEC of the same asymptote are repulsive and do not form minima, and hence are not illustrated in Fig. 1. This is consistent with the previous analysis of the PEC of FeH+.8,9,56 The next 4 states of FeH+ are triplet in spin (i.e., a3Σ, b3Φ, c3Π, d3Δ) and dissociate to the second lowest energy asymptote Fe+(4F) + H(2S). The quintet-spin states of the same fragments are mildly attractive and form shallow minima around 2–2.3 Å. Interestingly, the ordering of these quintet-spin states (i.e., 15Σ, 15Φ, 25Π, 25Δ) of Fe+(4F) + H(2S) is the same as the ordering of its triplet-spin states. Beyond this point, the electronic spectrum of FeH+ is quite complicated with a series of closely arranged electronic states dissociating to Fe+(4D) + H(2S), Fe+(4P) + H(2S), Fe+(2G) + H(2S), and more high energy fragments (Fig. 1).

The equilibrium electronic configurations of the studied 27 electronic states of FeH+ are reported in the Table 1. The contours of the occupying molecular orbitals are given in ESI, Fig. S1. Notice that the first 3 electronic states of FeH+ are dominantly single-reference in nature. The X5Δ has the 6σ2123 configuration and the attachment of an electron to its 7σ orbital produces the dominant electronic configuration of the ground state of FeH (X4Δ).28 The first excited state of FeH+ is formed by transferring an electron from the doubly occupied 1δ orbital to a 3π orbital (6σ2132). Similar to the ground state, placing an electron in the singly occupied 7σ of FeH+(A5Π) gives rise to the first excited state of FeH (A4Π).28 On the other hand, the electronic structures of B5Σ+ of FeH+ and c6Σ+ of FeH (fifth excited state of FeH) are closely related except for the additional electron occupying in the 8σ of FeH(c6Σ+).28 The third excited state of FeH+ (i.e., a3Σ) has a major configuration of 6σ242 but bears a small contribution of 6σ224 as well. The next 3 states are chiefly multireference in nature (i.e., b3Φ, c3Π, d3Δ). Similarly, all other studied states are multireference except for the slightly bound 15Σ and 25Δ states (Table 1 and Fig. 1).

Table 1 Dominant electronic configurations of the 27 electronic states of FeH+ at their corresponding equilibrium distances
Statea Coefficientb Configurationc
a The corresponding C2v symmetries of Δ, Π, Φ, Γ, and H are listed in parenthesis. b All the CI coefficients that are larger than 0.3 of corresponding natural orbital representations are listed. c β and α-spin electrons are specified with and without bars over the spatial orbital, respectively.
X5Δ (A1) 0.97 27σ3πxy(1δx2y2)2xy
A5Π (B1) 0.97 27σ3πx2y(1δx2y2)1δxy
B5Σ+ 0.97 22xy(1δx2y2)1δxy
a3Σ 0.87 2x2y2(1δx2y2)1δxy
−0.31 2xy(1δx2y2)2xy2
b3Φ (B1) 0.65 2xy2(1δx2y2)1δxy2
0.65 2x2y(1δx2y2)2xy
c3Π (B1) −0.52 2xy2(1δx2y2)1δxy2
0.52 2x2y(1δx2y2)2xy
−0.49 image file: d4cp03296a-t2.tif
d3Δ (A1) 0.77 image file: d4cp03296a-t3.tif
−0.34 image file: d4cp03296a-t4.tif
−0.34 image file: d4cp03296a-t5.tif
15Σ 0.90 6σ7σ3πx2y2(1δx2y2)1δxy
0.33 6σ7σ3πxy(1δx2y2)2xy2
15Φ (B1) −0.68 6σ7σ3πxy2(1δx2y2)1δxy2
0.68 6σ7σ3πx2y(1δx2y2)2xy
25Π (B1) 0.50 6σ7σ3πxy2(1δx2y2)1δxy2
0.50 6σ7σ3πx2y(1δx2y2)2xy
0.64 6σ7σ2x2y(1δx2y2)1δxy
25Δ (A1) 0.93 6σ7σ2xy(1δx2y2)2xy
13H (B1) 0.48 image file: d4cp03296a-t6.tif
0.48 27σ3πxy2(1δx2y2)2
−0.48 27σ3πxy2xy2
−0.48 image file: d4cp03296a-t7.tif
23Π (B1) 0.53 image file: d4cp03296a-t8.tif
−0.42 image file: d4cp03296a-t9.tif
0.40 27σ3πxy2(1δx2y2)2
0.40 27σ3πxy2xy2
11Γ (A1) −0.67 2x2y2(1δx2y2)2
0.67 2x2y2xy2
11Σ 0.57 2x2y2(1δx2y2)2
0.57 2x2y2xy2
13Γ (A1) 0.46 27σ3πy2(1δx2y2)1δxy2
−0.46 27σ3πx2(1δx2y2)1δxy2
−0.46 image file: d4cp03296a-t10.tif
0.46 image file: d4cp03296a-t11.tif
23Δ (A2) 0.68 image file: d4cp03296a-t12.tif
0.51 27σ3πx2y2xy
23Σ −0.41 image file: d4cp03296a-t13.tif
0.41 image file: d4cp03296a-t14.tif
0.41 27σ3πx2(1δx2y2)2xy
−0.41 27σ3πy2(1δx2y2)2xy
33Π (B1) 0.53 image file: d4cp03296a-t15.tif
0.53 27σ3πx(1δx2y2)2xy2
−0.31 image file: d4cp03296a-t16.tif
−0.31 image file: d4cp03296a-t17.tif
33Σ 0.57 2xy(1δx2y2)2xy2
−0.40 22xyxy2
−0.40 22xy(1δx2y2)2
11Π (B1) −0.43 image file: d4cp03296a-t18.tif
0.43 image file: d4cp03296a-t19.tif
0.43 image file: d4cp03296a-t20.tif
−0.43 image file: d4cp03296a-t21.tif
23Φ (B1) 0.41 image file: d4cp03296a-t22.tif
−0.41 27σ3πxy2(1δx2y2)2
0.41 27σ3πxy2xy2
−0.41 image file: d4cp03296a-t23.tif
33Δ (A1) 0.47 27σ3πy2(1δx2y2)1δxy2
0.47 27σ3πx2(1δx2y2)1δxy2
−0.39 image file: d4cp03296a-t24.tif
0.39 image file: d4cp03296a-t25.tif
35Δ (A1) 0.70 image file: d4cp03296a-t26.tif
0.32 image file: d4cp03296a-t27.tif
35Π (B1) 0.71 image file: d4cp03296a-t28.tif
0.31 6σ7σ2xy2(1δx2y2)1δxy
25Σ+ 0.73 image file: d4cp03296a-t29.tif
0.30 22xy(1δx2y2)1δxy
11Φ (B1) 0.36 image file: d4cp03296a-t30.tif
−0.36 image file: d4cp03296a-t31.tif
0.36 image file: d4cp03296a-t32.tif
−0.36 image file: d4cp03296a-t33.tif


The predominantly single-reference X5Δ, A5Π, and B5Σ+ states of FeH+ provide us with the opportunity of performing single-reference CCSD(T) calculations for them. The CCSD(T) calculations were performed on top of the RHF wave functions that were produced for their dominant electronic configurations listed in the Table 1. The results of the CCSD(T) calculations carried out with AQZ and A5Z basis sets are listed in the Table 2 along with the AQZ-MRCI and AQZ-MRCI+Q values of FeH+. Note that the larger A5Z basis set was only utilized with the CCSD(T) method since CCSD(T) level is relatively less expensive compared to the MRCI level. Furthermore, due to the less demanding nature of CCSD(T) (compared to MRCI), we evaluated the core electron correlation effects on various properties of FeH+ by unfreezing the 3s2 and 3p6 core electrons of Fe+ with the application of proper aug-cc-pwCV5Z basis set of Fe and the results are reported in the Table 2.

Table 2 Adiabatic dissociation energy with respect to the Fe+(6D) + H(2S) fragments De (kcal mol−1), bond length re (Å), excitation energy Te (cm−1), harmonic vibrational frequency ωe (cm−1), anharmonicity ωexe (cm−1), equilibrium rotational constant Be (cm−1), anharmonic correction to the rotational constant αe (cm−1), centrifugal distortion constant [D with combining macron]e (cm−1), and dipole moment (μ) at the equilibrium distance of the 27 low-lying electronic states of FeH+
State Level of theory D e r e T e ω e ω e x e B e α e [D with combining macron] e μ
X5Δ AQZ-MRCI 53.27 1.586 1844 34.1 6.765 0.1794 0.000385 −2.31
AQZ-MRCI+Q 55.24 1.587 1862 30.6 6.758 0.1736 0.000357
AQZ-DK-MRCI 52.23 1.581 1849 34.9 6.818 0.1878 0.000370 −2.26
AQZ-DK-MRCI+Q 54.22 1.582 1870 32.6 6.802 0.1864 0.000359
AQZ-CCSD(T) 53.90 1.590 1852 30.5 6.752 0.1884 0.000352 −2.50
A5Z-CCSD(T) 54.18 1.589 1852 30.5 6.743 0.1889 0.000356 −2.50
c-A5Z-CCSD(T) 54.44 1.578 1865 30.5 6.834 0.1703 0.000365 −2.34
MRCI10 1.5944 1836
MRCI+Q10 1.5891 1848
CCSDT10 1.5882 1850.4 32.2 6.7508 0.1761 0.000359
CCSDTQ10 1.5882 1849.8 32.3 6.7524 0.1765 0.000360
FPA10 1.5736 1874.2 31.9 6.8766 0.1798 0.000370
A5Π AQZ-MRCI 51.07 1.569 768 1820 34.7 6.906 0.1888 0.000401 −2.14
AQZ-MRCI+Q 53.18 1.570 723 1843 30.9 6.911 0.1852 0.000386
AQZ-DK-MRCI 50.66 1.563 550 1829 33.7 6.974 0.1984 0.000404 −2.08
AQZ-DK-MRCI+Q 52.77 1.563 507 1857 33.5 6.975 0.1972 0.000396
AQZ-CCSD(T) 51.87 1.575 710 1826 31.1 6.895 0.1755 0.000386 −2.34
A5Z-CCSD(T) 52.12 1.574 718 1827 31.4 6.878 0.1744 0.000387 −2.34
c-A5Z-CCSD(T) 52.10 1.564 819 1845 30.8 6.857 0.1784 0.000396 −2.19
MRCI10 1.5780 1821
MRCI+Q10 1.5711 1837
CCSDT10 1.5699 1835.9 32.7 6.9088 0.1861 0.000391
CCSDTQ10 1.5697 1835.0 32.8 6.9105 0.1866 0.000392
FPA10 1.5558 601 1859.4 31.7 7.0351 0.1898 0.000402
B5Σ+ AQZ-MRCI 40.69 1.637 4401 1709 37.3 6.347 0.1879 0.000352 −2.23
AQZ-MRCI+Q 43.12 1.633 4239 1722 32.4 6.389 0.1807 0.000358
AQZ-DK-MRCI 41.64 1.620 3703 1712 37.7 6.491 0.2052 0.000372 −2.10
AQZ-DK-MRCI+Q 44.09 1.615 3540 1750 33.4 6.527 0.2057 0.000363
AQZ-CCSD(T) 42.75 1.635 3902 1727 33.3 6.368 0.1797 0.000346 −2.48
A5Z-CCSD(T) 42.84 1.635 3965 1727 33.4 6.370 0.1820 0.000347 −2.45
c-A5Z-CCSD(T) 42.28 1.630 4253 1730 32.7 6.411 0.1838 0.000352 −2.31
a3Σ AQZ-MRCI 25.19 1.497 9821 1938 71.1 7.703 0.4284 0.000624 −1.40
AQZ-MRCI+Q 26.34 1.497 10[thin space (1/6-em)]109 1897 69.7 7.603 0.3068 0.000488
AQZ-DK-MRCI 20.92 1.489 10[thin space (1/6-em)]949 1959 61.8 7.680 0.3714 0.000561 −1.49
AQZ-DK-MRCI+Q 22.13 1.489 11[thin space (1/6-em)]223 1974 58.7 7.679 0.3622 0.000547
MRCI10 1.4907 1922
MRCI+Q10 1.4944 1901
CCSDT10 1.4831 2000.3 47.6 7.7419 0.2522 0.000464
CCSDTQ10 1.4862 1977.1 49.6 7.7093 0.2598 0.000469
FPA10 1.4821 10[thin space (1/6-em)]081 1965.0 51.0 7.7530 0.2745 0.000481
b3Φ AQZ-MRCI 22.29 1.517 10[thin space (1/6-em)]835 1886 92.0 7.515 0.4764 0.000653 −1.34
AQZ-MRCI+Q 23.29 1.517 11[thin space (1/6-em)]175 1844 81.9 7.395 0.4341 0.000475
AQZ-DK-MRCI 18.22 1.509 11[thin space (1/6-em)]896 1906 87.4 7.574 0.4006 0.000578 −1.45
AQZ-DK-MRCI+Q 19.27 1.509 12[thin space (1/6-em)]222 1922 84.6 7.566 0.3915 0.000569
c3Π AQZ-MRCI 19.26 1.533 11[thin space (1/6-em)]896 1824 99.1 7.399 0.5214 0.000635 −1.37
AQZ-MRCI+Q 20.30 1.532 12[thin space (1/6-em)]222 1769 83.4 7.253 0.3544 0.000488
AQZ-DK-MRCI 15.48 1.520 12[thin space (1/6-em)]853 1839 92.5 7.473 0.4255 0.000600 −1.46
AQZ-DK-MRCI+Q 16.58 1.520 13[thin space (1/6-em)]162 1866 89.1 7.469 0.4139 0.000599
d3Δ AQZ-MRCI 10.21 1.599 15[thin space (1/6-em)]063 1581 108.0 6.637 0.4950 0.000694 −1.16
AQZ-MRCI+Q 11.10 1.598 15[thin space (1/6-em)]438 1513 97.7 6.670 0.3950 0.000518
AQZ-DK-MRCI 7.15 1.574 15[thin space (1/6-em)]765 1617 111.5 6.876 0.5075 0.000712 −1.28
AQZ-DK-MRCI+Q 8.13 1.573 16[thin space (1/6-em)]120 1658 96.3 6.883 0.4891 0.000723
15Σ AQZ-MRCI 2.05 2.079 17[thin space (1/6-em)]915
AQZ-MRCI+Q 2.52 2.061 18[thin space (1/6-em)]442
15Φ AQZ-MRCI 1.47 2.149 18[thin space (1/6-em)]118
AQZ-MRCI+Q 1.86 2.133 18[thin space (1/6-em)]671
25Π AQZ-MRCI 0.71 2.171 18[thin space (1/6-em)]384
AQZ-MRCI+Q 1.13 2.149 18[thin space (1/6-em)]926
25Δ AQZ-MRCI 2.297 18[thin space (1/6-em)]862
AQZ-MRCI+Q 2.267 19[thin space (1/6-em)]424
13H AQZ-MRCI 1.534 20[thin space (1/6-em)]288 1824 66.9 7.399 0.5214 0.000735
AQZ-MRCI+Q 1.532 20[thin space (1/6-em)]527 1811 71.5 7.259 0.5462 0.000813
23Π AQZ-MRCI 1.546 20[thin space (1/6-em)]591 1959 22.4 7.121 0.4947 0.000376
AQZ-MRCI+Q 1.545 21[thin space (1/6-em)]002 1898 31.4 7.132 0.4183 0.000413
11Γ AQZ-MRCI 1.473 21[thin space (1/6-em)]215 1928 28.0 7.803 0.2190 0.000482
AQZ-MRCI+Q 1.472 21[thin space (1/6-em)]600 2016 29.7 7.855 0.2223 0.000477
11Σ AQZ-MRCI 1.485 21[thin space (1/6-em)]330 2085 62.2 7.735 0.2492 0.000457
AQZ-MRCI+Q 1.485 21[thin space (1/6-em)]722 2014 54.7 7.768 0.2580 0.000463
13Γ AQZ-MRCI 1.581 21[thin space (1/6-em)]526 1793 28.9 6.866 0.1694 0.000415
AQZ-MRCI+Q 1.580 21[thin space (1/6-em)]939 1755 30.1 6.825 0.1607 0.000413
23Δ AQZ-MRCI 1.580 21[thin space (1/6-em)]540 1746 38.9 6.821 0.2613 0.000406
AQZ-MRCI+Q 1.581 21[thin space (1/6-em)]954 1795 41.9 6.813 0.2814 0.000393
23Σ AQZ-MRCI 1.562 21[thin space (1/6-em)]733 1807 31.5 6.892 0.2360 0.000512
AQZ-MRCI+Q 1.561 22[thin space (1/6-em)]114 1801 29.7 6.884 0.1976 0.000489
33Π AQZ-MRCI 1.581 22[thin space (1/6-em)]430 1823 36.0 6.826 0.1877 0.000411
AQZ-MRCI+Q 1.580 22[thin space (1/6-em)]809 1831 36.1 6.839 0.1862 0.000404
33Σ AQZ-MRCI 1.598 22[thin space (1/6-em)]629 1769 55.9 6.684 0.2353 0.000326
AQZ-MRCI+Q 1.597 22[thin space (1/6-em)]919 1780 52.5 6.680 0.2188 0.000376
11Π AQZ-MRCI 1.508 22[thin space (1/6-em)]978 1918 46.4 7.482 0.2334 0.000457
AQZ-MRCI+Q 1.507 23[thin space (1/6-em)]379 1924 46.6 7.501 0.2448 0.000456
23Φ AQZ-MRCI 1.570 23[thin space (1/6-em)]521 1686 38.6 6.867 0.2403 0.000433
AQZ-MRCI+Q 1.568 23[thin space (1/6-em)]885 1744 36.2 6.839 0.2161 0.000404
33Δ AQZ-MRCI 1.591 24[thin space (1/6-em)]552 1761 35.6 6.725 0.1956 0.000392
AQZ-MRCI+Q 1.589 24[thin space (1/6-em)]897 1771 36.5 6.745 0.1960 0.000391
35Δ AQZ-MRCI 2.649 24[thin space (1/6-em)]745
AQZ-MRCI+Q 2.603 25[thin space (1/6-em)]168
35Π AQZ-MRCI 2.613 24[thin space (1/6-em)]794
AQZ-MRCI+Q 2.569 25[thin space (1/6-em)]232
25Σ+ AQZ-MRCI 2.707 24[thin space (1/6-em)]938
AQZ-MRCI+Q 2.678 25[thin space (1/6-em)]404
11Φ AQZ-MRCI 1.522 25[thin space (1/6-em)]851 1884 28.9 7.311 0.2061 0.000440
AQZ-MRCI+Q 1.520 26[thin space (1/6-em)]242 1831 28.6 7.366 0.2087 0.000440


The FeH+(X5Δ) is ∼8 kcal mol−1 more stable than FeH(X4Δ) [i.e., the adiabatic De of FeH(X4Δ) at c-A5Z-CCSD(T) level is 46.06 kcal mol−1].28 The AQZ-MRCI level predicted a De of 53.27 kcal mol−1 for the FeH+(X5Δ). MRCI+Q increased the De by ∼2 kcal mol−1 compared to MRCI. The AQZ-DK-MRCI and AQZ-DK-MRCI+Q Des are only ∼1 kcal mol−1 lower compared to the AQZ-MRCI and AQZ-MRCI+Q for FeH+(X5Δ). Under all utilized CCSD(T) levels, the Des of FeH+ are ∼54 kcal mol−1. The zero-point energy corrected binding energy of FeH+ under our largest non-relativistic level of theory c-A5Z-CCSD(T) is 51.79 kcal mol−1. This value is ∼1.5 kcal mol−1 greater than the upper bound of the D0 value of the experimental study by Elkind and Armentrout (i.e., 48.9 ± 1.4 kcal mol−1).5 The AQZ-DK-MRCI+Q D0 is closer to the c-A5Z-CCSD(T) D0 (i.e., 51.57 versus 51.79 kcal mol−1). The AQZ-DK-MRCI D0 is 1.95 kcal mol−1 lower compared to the AQZ-DK-MRCI+Q D0 of FeH+(X5Δ). Similar to the ground state, all Des predicted by the 3 CCSD(T) approaches are in between their AQZ-MRCI and AQZ-MRCI+Q values for both A5Π and B5Σ+. Overall, 10 electronic states of FeH+ bear positive Des compared to the ground state fragments [i.e., Fe+(6D) + H(2S) fragments].

The c-A5Z-CCSD(T) predicted re of FeH+ (X5Δ) is 1.578 Å which is slightly shorter compared to the res of core electron correlation disregarded approaches (∼1.59 Å). The same pattern was observed for the A5Π and B5Σ+ states as well. The observation of the tendency of core electron correlation to shorten the res is in agreement with our past experiences of transition metal-based diatomic species.57–61 This is due to the electron excitation from core-to-virtual orbitals which further exposes the nuclear charge of Fe+ to a favorable attraction with the valence electron of H. The implementation of relativistic effects decreased the res of FeH+ slightly. Specifically, the discrepancies between the AQZ-MRCI/AQZ-MRCI+Q versus AQZ-DK-MRCI/AQZ-DK-MRCI+Q are less than 0.03 Å. Among the chemically bound states, the longest re was observed for the B5Σ+ state, which is the only state to host 2 electrons in the 7σ orbital (note: the states with shallow minima are disregarded). The discrepancy between AQZ-MRCI and AQZ-MRCI+Q Tes is less than 570 cm−1 (Table 2). For all states, the AQZ-MRCI+Q predicted Tes are higher than the AQZ-MRCI (by 230–570 cm−1) except for the first 2 excited state of FeH+ in which the AQZ-MRCI+Q predicted 45 and 162 cm−1 lower Tes respectively compared to the AQZ-MRCI. Importantly, both AQZ-MRCI and AQZ-MRCI+Q provided the same ordering for all the electronic states reported in the present work. Among the available relativistic data, the discrepancy between AQZ-DK-MRCI and AQZ-DK-MRCI+Q Tes is less than 400 cm−1.

Overall, our results are in good agreement with the findings reported for the first few states of FeH+ by Cheng and DeYonker (Table 2 and ref. 10). The IE of FeH reported by Cheng and DeYonker under a focal point analysis (i.e., 7.4851 eV) is also in good agreement with our adiabatic IEs [i.e., 7.244, 7.263, and 7.425 eV at AQZ-CCSD(T), A5Z-CCSD(T), and c-A5Z-CCSD(T) levels, respectively].10

The spin–orbit coupling effects of FeH+ were evaluated by including the X5Δ, A5Π, B5Σ+, a3Σ, b3Φ, c3Π, and d3Δ electronic states of FeH+ in the spin–orbit matrix. These electronic states produce Ω = 4, 3, 2, 1, 0+, and 0 (X5Δ), Ω = 3, 2, 1, 1, 0+, and 0 (A5Π), Ω = 2, 1, and 0+ (B5Σ+), Ω = 1 and 0 (a3Σ), Ω = 4, 3, and 2 (b3Φ), Ω = 2, 1, 0+, and 0 (c3Π), and Ω = 3, 2, and 1 (d3Δ). The spin–orbit curves of these states are depicted in the Fig. 2. Their re, Te, ωe and ΛS compositions are given in the Table 3. The Ω products of the X5Δ span within 0–703 cm−1. The ground spin–orbit state of FeH2+ is indeed the Ω = 4 component of the X5Δ. The re and the ωe of the Ω = 4 ground state are slightly different from those of the parent X5Δ electronic state (Tables 2 and 3). The Ω = 4 ground spin–orbit state is 413 cm−1 more stabilized compared to the X5Δ of FeH+.


image file: d4cp03296a-f2.tif
Fig. 2 AQZ-DK-MRCI spin–orbit coupling curves resulting from X5Δ, A5Π, B5Σ+, a3Σ, b3Φ, c3Π, and d3Δ electronic states of FeH+ as a function of Fe+⋯H distance [r(Fe+⋯H), Å]. The relative energies are referenced with respect to the Ω = 4 ground state minimum of FeH+. The Ω = 4, Ω = 3, Ω = 2, Ω = 1, Ω = 0+, and Ω = 0 curves are shown in green, black, red, blue, cyan, and pink, respectively. See Fig. 1 for the PEC of their parent electronic states.
Table 3 Bond length re (Å), excitation energy Te (cm−1), harmonic vibrational frequency ωe (cm−1), and % ΛS composition of several low-lying spin–orbit states of FeH+ at AQZ-DK-MRCI level
Ω r e T e ω e ΛS compositiona
a Only components that are equal or larger than 1% are listed.
4 1.586 0 1859 100% X5Δ
3 1.584 139 1868 89% X5Δ + 11% A5Π
2 1.583 288 1800 82% X5Δ + 18% A5Π
1 1.582 479 1797 75% X5Δ + 25% A5Π
0+ 1.581 688 1792 68% X5Δ + 31% A5Π
0 1.581 703 1794 73% X5Δ + 27% A5Π
3 1.570 812 1844 89% A5Π + 11% X5Δ
2 1.572 949 1851 81% A5Π + 18% X5Δ
1 1.573 1063 1861 76% A5Π + 23% X5Δ
1 1.568 1130 1784 98% A5Π + 2% X5Δ
0+ 1.573 1134 1884 67% A5Π + 32% X5Δ + 1% B5Σ+
0 1.573 1174 1872 73% A5Π + 27% X5Δ
2 1.623 4005 1725 100% B5Σ+
1 1.622 4033 1724 99% B5Σ+ + 1% A5Π
0+ 1.622 4042 1728 98% B5Σ+ + 2% A5Π
0 1.492 11[thin space (1/6-em)]577 1930 95% a3Σ + 5% c3Π
1 1.492 11[thin space (1/6-em)]628 1938 97% a3Σ + 3% c3Π
4 1.511 12[thin space (1/6-em)]060 1893 100% b3Φ
3 1.511 12[thin space (1/6-em)]574 1894 100% b3Φ
2 1.511 13[thin space (1/6-em)]093 1890 100% b3Φ
2 1.522 13[thin space (1/6-em)]366 1828 100% c3Π
1 1.522 13[thin space (1/6-em)]654 1842 96% c3Π + 3% a3Σ
0 1.523 13[thin space (1/6-em)]818 1834 100% c3Π
0+ 1.521 13[thin space (1/6-em)]941 1847 95% c3Π + 5% a3Σ
3 1.577 16[thin space (1/6-em)]136 1627 100% d3Δ
2 1.577 16[thin space (1/6-em)]514 1624 99% d3Δ
1 1.577 16[thin space (1/6-em)]877 1619 100% d3Δ


Accurate μ values are vital for calculating radiative characteristics, spectra, and opacities of molecules. Aiming to aid such future studies, here we report MRCI and CCSD(T) μ values of several low-lying electronic states of FeH+ (Table 2). Among all states (at equilibrium distances) the largest μ was observed for the ground state of FeH+ [∼2.3 D at AQZ-MRCI, AQZ-DK-MRCI, and c-A5Z-CCSD(T)] (Table 2). This value is ∼0.1 D smaller than the MCPF μ reported by Lars et al. in 1987 (i.e., 2.41 D).7 The relativistic effects caused a minor change in μ (Table 2). Specifically, the largest difference between AQZ-MRCI versus AQZ-DK-MRCI was observed for the d3Δ state and it is only 0.12 D. Upon comparison of non-relativistic analysis, AQZ-MRCI μ versus c-A5Z-CCSD(T) μ of each X5Δ, A5Π, and B5Σ+ are in better agreement and the discrepancies are less than 0.1 D. The AQZ-CCSD(T) and A5Z-CCSD(T) are very close to each other (Table 2) but they are 0.1–0.2 D larger compared to c-A5Z-CCSD(T) μ values.

Here we further report μ of FeH+ (X5Δ) under a series of exchange correlation functionals that span multiple rungs of Jacob's ladder of DFA aiming to assess its density functional theory errors. Specifically, we used GGAs (BP86,29,30 BLYP,31,32 PBE33), MGGAs (TPSS,34 MN15-L35), global GGA hybrids (B3LYP,36–38 B3P86,29,36 B3PW91,36,39 PBE040), MGGA hybrids (TPSSh,34 M06-2X,41 MN1542), RSHs (LRC-ωPBE,43 CAM-B3LYP,44 ωB97X45), and DHs (PBE0-DH,46 DSDPBEP8647,48). We utilized the AQZ-CCSD(T) total μ (2.50 D) of FeH+ (X5Δ) to assess DFT errors since DFT calculations were also performed under the AQZ basis set. We see a general trend of improvement of μ, when moving from lower to higher rungs of DFA (ESI, Fig. S2 and Table S1). Compared to CCSD(T), the more expensive DHs (PBE0-DH and DSDPBEP86) overestimated μ by ∼5 and 13% (ESI, Table S1). All 3 functionals of RSH predicted μ values with less than 9% of errors. The MGGA hybrid M06-2X μ is almost identical to the AQZ-CCSD(T) μ value (2.53 versus 2.50 D) and this is the best performing functional for μ of FeH+(X5Δ) among all 17 DFAs. The errors of global GGA hybrids span in between 9–17%. The MGGA MN15-L is a clear outlier of the linear-like μ improving trend going from GGA to DHs. However, the MN15-L μ is closer to the AQZ-CCSD(T) μ than for any functional of GGA. The largest deviation of DFT μ compared to the AQZ-CCSD(T) was observed for the least expensive GGAs with approximate errors of 30% (ESI, Table S1). Overall, our general expectation that the more expensive functionals from the higher rungs of the Jacob's ladder of DFA would perform better compared to the ones at the lower rungs holds true for the μ of FeH+(X5Δ).18

The DMC of the first 7 electronic states of FeH+ obtained at the MRCI level are illustrated in the Fig. 3a. Among the focused range, the largest total μ was observed for the B5Σ+ (2.5 D) around 1.3 Å. The DMC minima of X5Δ and A5Π were observed at ∼1.5 Å with −2.4 and −2.2 D, respectively. The DMC of all 4 triplet-spin states (i.e., b3Φ, c3Π, d3Δ, and a3Σ) are qualitatively and quantitatively similar throughout the scale and they reach the minimization around the 1.3–1.4 Å. Note that for all states, the DMC minima were observed at slightly shorter Fe+⋯H distances compared to their equilibrium distances (Fig. 1 and 3). The TDMC arising from the lowest 7 states of FeH+ are given in Fig. 3b. Among the studied quintet-spin states (i.e., X5Δ, A5Π, and B5Σ+) the X5Δ ↔ A5Π and B5Σ+ ↔ A5Π transitions are permitted, whereas the transition between X5Δ ↔ B5Σ+ is forbidden. The largest transition μ values of X5Δ ↔ A5Π and B5Σ+ ↔ A5Π were observed at approximately 1.6 and 1.3 Å and they are only 0.08 and 0.11 D, respectively. The d3Δ ↔ c3Π, a3Σ ↔ c3Π, and d3Δ ↔ b3Φ transitions are allowed for the studied triplet-spin states of FeH+ and the corresponding transition μ values are increasing with the compression of the Fe+⋯H distance.


image file: d4cp03296a-f3.tif
Fig. 3 (a) MRCI/aug-cc-pVQZ DMC of FeH+ as a function of Fe+⋯H distance [r(Fe+⋯H), Å]. The solid and dotted DMC correspond to the triplet and quintet spins, respectively. The Δ, Π, Σ+, Σ, and Φ states are shown in red, blue, pink, green, and black, respectively. (b) MRCI/aug-cc-pVQZ TDMC resulting from X5Δ ↔ A5Π, B5Σ+ ↔ A5Π, d3Δ ↔ c3Π, a3Σ ↔ c3Π, and d3Δ ↔ b3Φ of FeH+ as a function of Fe+⋯H distance [r(Fe+⋯H), Å].

III.B. FeH2+

The removal of the 4s1 valence electron from the Fe+ (6D; [Ar]3d64s1) produces the ground electronic state of Fe2+ (5D; [Ar]3d6).53 The low energy electronic spectrum of Fe2+ is much less dense compared to the spectrum of Fe+. For example, the first excited state of Fe2+ (3P; [Ar]3d6) lies 55.48–60.64 kcal mol−1 above, whereas Fe+ populates 5 excited states within 0–55 kcal mol−1 range.53 Since the excitation energies of Fe2+ are relatively high in energy, in the present work we have only considered the reaction between the ground state of Fe2+(5D) and H(2S). According to the Wigner–Witmer rules, this combination produces 6,4+, Π, Δ] electronic states. Here all of these states were analyzed under the MRCI level of theory.54,55 Similar to FeH+, the CASSCF active space used for MRCI calculations of FeH2+ was constructed from the five 3d, five 4d, and the 4s orbitals of Fe and the 1s orbital of H [CAS(7,12)]. The full PEC of the 6 electronic states of FeH2+ studied are given in the Fig. 4.
image file: d4cp03296a-f4.tif
Fig. 4 MRCI/aug-cc-pVQZ PEC of FeH2+ as a function of Fe2+⋯H distance [r(Fe2+⋯H), Å]. The relative energies are with respect to the total energy of Fe2+(5D) + H(2S) when they are at 200 Å separation, which is set to 0 kcal mol−1. The solid and dotted PEC correspond to the quartet and sextet spins, respectively. The Δ, Π, and Σ+ states are shown in red, blue, and green, respectively.

All PEC are attractive in nature with ∼14–23 kcal mol−1De. According to the MRCI potential energy profile, the ground state of FeH2+ is a 4Π state, followed very closely by a 4Δ state (Fig. 4). The two most stable states of FeH2+ (4Π and 4Δ) are multireference in nature (Table 4). The electronic configurations of the two main components of the 4Π state are 6σ232 (55%) and 7σ232 (18%). Notice that the major configuration of 4Π state (6σ232) can be produced by eliminating the 7σ1 electron from the A5Π of FeH+ (6σ2132). The MRCI and MRCI+Q predicted IEs of this process are 17.0 and 17.2 eV, respectively. On the other hand, the dominant configuration of the 4Δ (6σ223) is the 7σ1 electron ionized product of the FeH+ (X5Δ; 6σ2123). The next 3 states of FeH2+ carry single-reference electronic configurations and hence those were further analyzed with CCSD(T) levels of theory. All our CCSD(T) and MRCI numerical findings of FeH2+ are listed in Table 5.

Table 4 Dominant electronic configurations at the equilibrium distances of the 6 low-lying electronic states of FeH2+
Statea Coefficientb Configurationc
a The corresponding A1 (of Δ) and B1 (of Π) components under C2v symmetry are listed. b All the CI coefficients that are larger than 0.3 of corresponding natural orbital representations are given. c β and α-spin electrons are specified with and without bars over the spatial orbital, respectively.
4Π 0.74 2x2y(1δx2y2)1δxy
−0.43 2x2y(1δx2y2)1δxy
4Δ 0.73 2xy(1δx2y2)2xy
−0.44 2xy(1δx2y2)2xy
6Σ+ 0.99 27σ3πxy(1δx2y2)1δxy
6Δ 0.99 6σ7σ3πxy(1δx2y2)2xy
6Π 0.99 6σ7σ3πx2y (1δx2y2)1δxy
4Σ+ 0.83 image file: d4cp03296a-t34.tif
−0.32 image file: d4cp03296a-t35.tif


Table 5 Adiabatic dissociation energy with respect to the Fe2+(5D) + H(2S) fragments De (kcal mol−1), bond length re (Å), excitation energy Te (cm−1), harmonic vibrational frequency ωe (cm−1), anharmonicity ωexe (cm−1), equilibrium rotational constant Be (cm−1), anharmonic correction to the rotational constant αe (cm−1), centrifugal distortion constant [D with combining macron]e (cm−1), and dipole moment (μ) at equilibrium distance of the 6 low-lying electronic states of FeH2+
State Level of theory D e r e T e ω e ω e x e B e α e [D with combining macron] e μ
4Π AQZ-MRCI 22.85 1.923 841 16.3 4.604 0.1566 0.000552 1.21
AQZ-MRCI+Q 23.13 1.918 846 16.2 4.639 0.1574 0.000529
AQZ-DK-MRCI 23.47 1.908 848 15.8 4.690 0.1566 0.000538 1.22
AQZ-DK-MRCI+Q 23.76 1.902 852 15.7 4.711 0.1157 0.000562
4Δ AQZ-MRCI 22.47 1.989 135 839 17.0 4.314 0.1369 0.000455 1.36
AQZ-MRCI+Q 22.68 1.986 156 841 17.0 4.329 0.1367 0.000453
AQZ-DK-MRCI 22.95 1.977 180 842 16.5 4.368 0.1366 0.000450 1.36
AQZ-DK-MRCI+Q 23.18 1.973 203 845 16.5 4.384 0.1364 0.000467
MRCI+Q12 21.68 1.998 830
6Σ+ AQZ-MRCI 19.90 1.999 1033 719 4.9 4.263 0.1503 0.000567 1.15
AQZ-MRCI+Q 20.15 1.994 1042 727 4.4 4.297 0.1512 0.000574
AQZ-DK-MRCI 21.60 1.913 654 712 11.1 4.585 0.2079 0.000768 1.05
AQZ-DK-MRCI+Q 21.89 1.909 653 719 14.3 4.675 0.2141 0.000781
AQZ-CCSD(T) 19.95 2.016 728 6.1 4.195 0.1144 0.000582 1.20
A5Z-CCSD(T) 19.95 2.018 734 6.2 4.186 0.1129 0.000582 1.21
c-A5Z-CCSD(T) 20.84 2.005 747 5.6 4.240 0.1225 0.000539 1.25
6Δ AQZ-MRCI 18.92 2.203 1376 836 21.3 3.514 0.1370 0.000247 1.38
AQZ-MRCI+Q 19.04 2.202 1430 836 21.2 3.516 0.1337 0.000248
AQZ-DK-MRCI 19.64 2.190 1339 891 26.3 3.551 0.1180 0.000225 1.40
AQZ-DK-MRCI+Q 19.77 2.197 1396 887 26.3 3.630 0.1129 0.000223
AQZ-CCSD(T) 19.06 2.202 837 22.6 3.512 0.1350 0.000287 1.42
A5Z-CCSD(T) 19.07 2.202 837 22.5 3.513 0.1343 0.000262 1.42
c-A5Z-CCSD(T) 20.01 2.188 850 22.2 3.556 0.1331 0.000258 1.48
6Π AQZ-MRCI 18.17 2.187 1638 819 21.7 3.560 0.1428 0.000246 1.34
AQZ-MRCI+Q 18.31 2.185 1687 820 21.2 3.570 0.1405 0.000269
AQZ-DK-MRCI 18.96 2.171 1575 834 21.8 3.613 0.1375 0.000270 1.38
AQZ-DK-MRCI+Q 19.11 2.169 1624 832 21.0 3.618 0.1144 0.000275
AQZ-CCSD(T) 18.33 2.186 821 22.7 3.562 0.1404 0.000274 1.40
A5Z-CCSD(T) 18.35 2.186 821 22.6 3.563 0.1391 0.000246 1.40
c-A5Z-CCSD(T) 19.21 2.173 833 22.3 3.601 0.1372 0.000247 1.37
4Σ+ AQZ-MRCI 14.18 2.246 3031 712 22.4 3.388 0.1484 0.000315 1.17
AQZ-MRCI+Q 14.34 2.242 3073 716 22.1 3.389 0.1460 0.000309
AQZ-DK-MRCI 14.72 2.225 3060 719 21.2 3.450 0.1450 0.000324 1.22
AQZ-DK-MRCI+Q 14.89 2.220 3101 724 21.4 3.465 0.1448 0.000325


The De of FeH2+(4Π) under AQZ-MRCI and AQZ-MRCI+Q levels are 22.85 and 23.13 kcal mol−1. The zero-point energy corrected AQZ-MRCI and AQZ-MRCI+Q D0s of FeH(4Π) are 21.66 and 21.94 kcal mol−1. Under both AQZ-MRCI and AQZ-MRCI+Q levels the 4Δ state lies only 0.4 kcal mol−1 above the 4Π (Table 5) (i.e., the AQZ-MRCI and AQZ-MRCI+Q D0 of 4Δ are 21.29 and 21.49, respectively). Our MRCI+Q is only 1 kcal mol−1 larger than the MRCI+Q De of 4Δ reported by Wilson et al. in 2003 (22.68 versus 21.68 kcal mol−1).12 Our AQZ-DK-MRCI and AQZ-DK-MRCI+Q D0 values of FeH2+(4Π) are 22.27 and 22.57 kcal mol−1, respectively. The introduction of the relativistic effects only increased the D0 of FeH2+(4Π) by ∼1 kcal mol−1. Importantly, since the energy difference between the 4Π and 4Δ states is within the margin of error of the basis set and the method, it is difficult to assign a true ground state for FeH2+. We performed AQZ-CCSD(T), A5Z-CCSD(T), and c-A5Z-CCSD(T) calculations for the single-reference 6Σ+, 6Δ, and 6Π states of FeH2+. For all these 3 states, the AQZ-CCSD(T) Des and A5Z-CCSD(T) Des are almost identical to each other (Table 5). As expected, the electron excitation from core-to-virtual orbitals [i.e., c-A5Z-CCSD(T)] relaxes (or stabilizes) the electronic states increasing the De values approximately by 0.9 kcal mol−1. The bond lengths of the electronic states of FeH2+ are significantly longer compared to the res of the states of FeH+ which translate to the lower Des of FeH2+ compared to FeH+ (compare res and Des given in Tables 2 and 5). This also means that an apparent measured IE of FeH+ will likely be at higher energies (due to the Franck–Condon overlap). The MRCI+Q re of the 4Δ state reported by Wilson et al., is 0.01 Å longer compared to our MRCI+Q value.12 For all states, MRCI+Q predicted slightly shorter res compared to MRCI (by ∼0.001–0.005 Å). Similar to the FeH+ case, the relativistic effects on the res of the states of the FeH2+ are minor (Table 5). The c-A5Z-CCSD(T) res of FeH2+ are shorter compared to the A5Z-CCSD(T) res similar to the FeH+ case. The AQZ-MRCI+Q predicted slightly higher Tes compared to the AQZ-MRCI Tes (0–60 cm−1). The largest discrepancy between the AQZ-DK-MRCI/AQZ-DK-MRCI+Q versus AQZ-MRCI/AQZ-MRCI+Q was observed for the 6Σ+ state which is ∼400 cm−1, whereas in all other cases it is less than 65 cm−1. Finally, we observed that all AQZ-MRCI+Q, AQZ-MRCI, AQZ-DK-MRCI, and AQZ-DK-MRCI+Q levels’ predictions on spectroscopic constants agree well with each other (i.e., ωe, ωexe, Be, αe, and [D with combining macron]e).

To investigate the spin–orbit effects of FeH2+, we have included 4Π, 4Δ, 6Σ+, 6Δ, 6Π, and 4Σ+ electronic states in the spin–orbit matrix. The spin–orbit coupling produces the Ω = 1/2, 1/2, 3/2, and 5/2 (from 4Π), Ω = 1/2, 3/2, 5/2, and 7/2 (from 4Δ), Ω = 1/2, 3/2, and 5/2 (from 6Σ+), Ω = 1/2, 1/2, 3/2, 5/2, 7/2, and 9/2 (from 6Δ), Ω = 1/2, 1/2, 3/2, 3/2, 5/2, and 7/2 (from 6Π), and Ω = 1/2 and 3/2 (from 4Σ+). The spin–orbit curves of the Ω states are given in the Fig. 5 and the corresponding re, Te, ωe, and ΛS compositions are listed in the Table 6. The Ω states of the ground 4Π electronic states span within 0–525 cm−1 whereas those of the first excited 4Δ extend from 150 to 1021 cm−1. The ground spin–orbit state of the FeH2+ is an Ω = 5/2 which is stabilized over its parent 4Π state by 353 cm−1. As expected, the Ω = 5/2 ground spin–orbit state bears substantial composition of 4Δ (20%) due to the proximity of the 4Π and 4Δ states. Similarly, notable mixings were observed for many Ω states which clearly highlights the importance of the spin–orbit coupling effects of the FeH2+ system (Table 6).


image file: d4cp03296a-f5.tif
Fig. 5 AQZ-DK-MRCI spin–orbit coupling curves resulting from 4Π, 4Δ, 6Σ+, 6Δ, 6Π, and 4Σ+ electronic states of FeH2+ as a function of Fe2+⋯H distance [r(Fe2+⋯H), Å]. The relative energies are referenced with respect to the Ω = 5/2 ground state minimum of FeH2+. The Ω = 1/2, Ω = 3/2, Ω = 5/2, Ω = 7/2, and Ω = 9/2 curves are shown in blue, red, green, cyan, and black, respectively. See Fig. 4 for the PEC of their parent 4Π, 4Δ, 6Σ+, 6Δ, 6Π, and 4Σ+ states.
Table 6 Bond length re (Å), excitation energy Te (cm−1), harmonic vibrational frequency ωe (cm−1), and % ΛS composition of several low-lying spin–orbit states of FeH2+ at AQZ-DK-MRCI level
Ω r e T e ω e ΛS compositiona
a Only components that are equal or larger than 1% are listed.
5/2 1.922 0 810 78% 4Π + 20% 4Δ + 2% 6Σ+
3/2 1.919 123 823 82% 4Π + 16% 4Δ + 2% 6Σ+
7/2 1.958 150 858 100% 4Δ
1/2 1.917 299 808 88% 4Π + 10% 4Δ + 2% 6Σ+
1/2 1.909 524 867 98% 4Π + 2% 6Σ+
5/2 1.966 538 913 72% 4Δ + 25% 4Π + 3% 6Σ+
3/2 1.964 789 833 75% 4Δ + 17% 4Π + 8% 6Σ+
1/2 1.947 921 626 85% 6Σ+ + 9% 4Δ + 5% 6Π
3/2 1.942 964 724 85% 6Σ+ + 6% 4Π + 5% 6Π + 4% 4Δ
5/2 1.929 1007 675 93% 6Σ+ + 4% 4Π + 2% 6Π
1/2 1.967 1021 862 79% 4Δ + 14% 4Π + 6% 6Σ+
9/2 2.197 1250 904 100% 6Δ
7/2 2.187 1341 916 80% 6Δ + 20% 6Π
5/2 2.150 1539 1031 65% 6Δ + 23% 6Σ+ + 10% 6Π + 1% 4Π
3/2 2.164 1659 917 64% 6Δ + 23% 6Π + 12% 6Σ+ + 1% 4Π
1/2 2.173 1778 896 53% 6Δ + 42% 6Π + 4% 6Σ+ + 1% 4Π
7/2 2.174 1818 825 78% 6Π + 21% 6Δ + 1% 4Δ
1/2 2.169 1978 869 61% 6Δ + 30% 6Π + 8% 6Σ+ + 1% 4Π
5/2 2.161 2035 924 66% 6Π + 23% 6Δ + 10% 6Σ+ + 1% 4Δ
3/2 2.174 2112 822 76% 6Π + 22% 6Δ
1/2 2.179 2238 743 53% 6Π + 45% 6Δ + 2% 6Σ+
3/2 2.153 2247 826 77% 6Π + 15% 6Σ+ + 8% 6Δ
1/2 2.162 2358 830 57% 6Π + 31% 6Δ + 11% 6Σ+
3/2 2.220 3449 667 98% 4Σ+ + 1% 6Π + 1% 4Π
1/2 2.215 3487 732 97% 4Σ+ + 3% 4Π


The AQZ-MRCI and AQZ-DK-MRCI μ values of the 4Π and 4Δ states at their res are 1.2 and 1.4 D, respectively. Among all states, the largest and smallest μ values were observed for the 6Δ and 6Σ+ states, respectively (Table 5). The relativistic effects on the μ values of FeH2+ are minor, where the largest difference was observed for the 6Σ+ state which is only 0.1 D. Similar to the FeH+ case, the μ values predicted by AQZ-CCSD(T) levels are larger than the AQZ-MRCI μ. The calculated AQZ-MRCI DMC of the 6 low-lying states of FeH2+ are shown in Fig. 6a. Similar to FeH+, the μ values of FeH2+ increase moving to shorter internuclear distances and shift towards the negative direction. The maxima of the DMC were observed around the 2.4–2.5 Å. Only 4Δ ↔ 4Π, 6Δ ↔ 6Π, 6Σ+6Π, and 4Σ+4Π transitions are allowed for the studied states of FeH2+. The TDMC corresponding to these transitions are illustrated in the Fig. 6b. Among these transitions, the smallest transition μ values were observed for the Δ ↔ Π. Comparatively, the Σ+ ↔ Π transition μ values are significant. Especially, the 4Σ+4Π transition μ values increase exponentially moving to shorter internuclear distances.


image file: d4cp03296a-f6.tif
Fig. 6 (a) AQZ-MRCI DMC of FeH2+ as a function of Fe2+⋯H distance [r(Fe2+⋯H), Å]. The solid and dotted DMC correspond to the quartet and sextet spins, respectively. The Δ, Π, and Σ+ states are shown in red, blue, and green, respectively. (b) AQZ-MRCI TDMC resulting from 4Δ ↔ 4Π, 6Δ ↔ 6Π, 6Σ+6Π, and 4Σ+4Π of FeH2+ as a function of Fe2+⋯H distance [r(Fe2+⋯H), Å].

III.C. TIPS of FeH+ and FeH2+

In the TIPS calculations of the of FeH+ and FeH2+, we included all bound rovibrational levels50 that were allowed from the respective PEC (see Fig. 1 and 4). Here the FeH+ 33Σ, 13H, and 33Δ PEC were extrapolated by fitting each PEC with an extended Morse PEC, noting that these states are products of the Fe+(a2H) + H(2S) reaction. For the states included within the FeH+ and FeH2+ models, we estimate that the present TIPS results are accurate (within ∼0.05%) up to approximately 5000 K and 1000 K, respectively. However, the errors associated with the accuracy of the PEC and solution of the rovibrational Schrödinger equation are not taken into account. We expect that the present TIPS calculation will have reasonable errors at low temperatures ≲1500 K, where the accuracy of the method, spin–orbit coupling and non-adiabatic effects can be relatively important to the TIPS. This is the first presentation of the FeH+ and FeH2+ TIPS in the literature, hence we provide results for the TIPS beyond the expected range of accuracy.

The following TIPS fit function62

image file: d4cp03296a-t36.tif
can represent the TIPS over the temperature range 10–30[thin space (1/6-em)]000 K (and perhaps even over a broader range). The a coefficients were optimized to minimize the maximum error of the fit function, noting that we fit the TIPS over a broader range of temperatures than we expect the TIPS to be accurate for in equations of state calculations. The FeH+ and FeH2+N = 17 and N = 19 fit coefficients are given in ESI, Tables S2 and S3, where the errors of the fit functions are less than approximately 0.06% and 0.07%, respectively over the 10–30[thin space (1/6-em)]000 K range. Note this error is only due to the fit (see above).

IV. Conclusions

The MRCI and CCSD(T) WFT calculations were performed with large correlation consistent basis sets to analyze the ground and excited electronic states of FeH+ and FeH2+. Multireference calculations were constructed using a bigger active space made of the five 3d, five 4d, and the 4s atomic orbitals of Fe and the 1s atomic orbital of H. We introduced 27 and 6 MRCI PEC for FeH+ and FeH2+, respectively. The ωe, ωexe, Be, αe, [D with combining macron]e spectroscopic parameters and re, Te, and μ values of FeH+ and FeH2+ were obtained using the CCSD(T), MRCI, and MRCI+Q levels of theory. At the MRCI level, spin–orbit coupling effect of FeH+ and FeH2+ were also tested. FeH+ has 10 bound electronic states with respect to the Fe+(6D) + H(2S) fragments, but 3 of them are relatively weakly bound with less than 3 kcal mol−1De. The ground state of FeH+ is a X5Δ with a D0 51.8 kcal mol−1. This value is in reasonable agreement with the previously reported experimental D0 value of FeH+ (i.e., 48.9 ± 1.4 kcal mol−1). The FeH+(X5Δ) bears the 6σ2123 electronic configuration which can be produced by ionizing an electron from the 7σ2 of the dominant configuration of the FeH(X4Δ). The calculated IE of this process is 7.425 eV. Among the studied electronic states, the largest μ was observed for the ground state of FeH+(X5Δ). For all states we observed a good agreement between AQZ-MRCI μ versus c-A5Z-CCSD(T) μ. The μ values of single-reference FeH+(X5Δ) were calculated with a series of functionals that span multiple rungs of Jacob's ladder of DFA and compared with the highly reliable CCSD(T) value obtained with the finite-field method. In agreement with our expectation, we observed a general trend of improving μ going from lower to higher rungs of Jacob's ladder of DFA. Furthermore, we introduced the MRCI DMC and TDMC corresponding to several low-lying electronic states of FeH+ and FeH2+. Compared to the FeH+(X5Δ), the FeH2+(4Π) is ∼30 kcal mol−1 less strongly bound. The two most stable electronic states of FeH2+ (i.e., 4Π and 4Δ) are multireference in nature and bound by ∼23 kcal mol−1 with respect to Fe2+(5D) + H(2S) dissociation. Lower μ values were observed for the low-lying electronic states of FeH2+ compared to those of FeH+. The transition μ values for both FeH+ and FeH2+ are relatively small and hence we expect those transitions to produce weak bands in the corresponding spectra. Finally, we provided a fit function and coefficients for calculation of the FeH+ and FeH2+ TIPS. These are the only available TIPS data in the literature for these molecules.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Aeronautics and Space Administration (NASA) ROSES-NRA NNH20ZDA001N-XRP, Los Alamos National Laboratory (LANL) Laboratory Directed Research and Development projects 20240039DR and 20240737PRD1. This research used resources provided by the Los Alamos National Laboratory Institutional Computing Program, which is supported by the U.S. Department of Energy National Nuclear Security Administration under Contract No. 89233218CNA000001. Reviewers are thanked for their useful comments on improving this work.

References

  1. H. R. Johnson and A. J. Sauval, A&AS, 1982, 49, 77–87 CAS .
  2. E. I. Mysov, I. R. Lyatifov, R. B. Materikova and N. S. Kochetkova, J. Organomet. Chem., 1979, 169, 301–308 CrossRef CAS .
  3. L. F. Halle, F. S. Klein and J. L. Beauchamp, J. Am. Chem. Soc., 1984, 106, 2543–2549 CrossRef CAS .
  4. J. B. Schilling, W. A. Goddard and J. L. Beauchamp, J. Am. Chem. Soc., 1986, 108, 582–584 CrossRef CAS .
  5. J. L. Elkind and P. B. Armentrout, J. Phys. Chem., 1986, 90, 5736–5745 CrossRef CAS .
  6. J. B. Schilling, W. A. Goddard and J. L. Beauchamp, J. Phys. Chem., 1987, 91, 5616–5623 CrossRef CAS .
  7. L. G. M. Pettersson, C. W. Bauschlicher, S. R. Langhoff and H. Partridge, J. Chem. Phys., 1987, 87, 481–492 CrossRef CAS .
  8. M. Sodupe, J. M. Lluch, A. Oliva, F. Illas and J. Rubio, J. Chem. Phys., 1989, 90, 6436–6442 CrossRef CAS .
  9. S. R. Langhoff and C. W. Bauschlicher, Jr., Astrophys. J., 1991, 375, 843–845 CrossRef CAS .
  10. Q. Cheng and N. J. DeYonker, J. Chem. Phys., 2019, 150, 234304 CrossRef PubMed .
  11. S. Jin, J. Heller, C. van der Linde, M. Oncak and M. K. Beyer, J. Phys. Chem. Lett., 2022, 13, 5867–5872 CrossRef CAS PubMed .
  12. D. J. D. Wilson, C. J. Marsden and E. I. von Nagy-Felsobuki, Phys. Chem. Chem. Phys., 2003, 5, 252–258 RSC .
  13. H.-J. Werner and P. J. Knowles, J. Chem. Phys., 1988, 89, 5803–5814 CrossRef CAS .
  14. P. J. Knowles and H.-J. Werner, Chem. Phys. Lett., 1988, 145, 514–522 CrossRef CAS .
  15. K. R. Shamasundar, G. Knizia and H. J. Werner, J. Chem. Phys., 2011, 135, 054101 CrossRef CAS PubMed .
  16. S. R. Langhoff and E. R. Davidson, Int. J. Quantum Chem., 1974, 8, 61–72 CrossRef CAS .
  17. K. Raghavachari, G. W. Trucks, J. A. Pople and M. Head-Gordon, Chem. Phys. Lett., 1989, 157, 479–483 CrossRef CAS .
  18. J. P. Perdew and K. Schmidt, AIP Conf. Proc., 2001, 577, 1–20 CrossRef CAS .
  19. H. J. Werner, P. J. Knowles, G. Knizia, F. R. Manby and M. Schütz, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2011, 2, 242–253 Search PubMed .
  20. H. J. Werner, P. J. Knowles, F. R. Manby, J. A. Black, K. Doll, A. Hesselmann, D. Kats, A. Kohn, T. Korona, D. A. Kreplin, Q. Ma, T. F. Miller, 3rd, A. Mitrushchenkov, K. A. Peterson, I. Polyak, G. Rauhut and M. Sibaev, J. Chem. Phys., 2020, 152, 144107 CrossRef CAS PubMed .
  21. H.-J. Werner, P. J. Knowles, et al., MOLPRO, version 2023.2, a package of ab initio programs, see https://www.molpro.net Search PubMed .
  22. N. B. Balabanov and K. A. Peterson, J. Chem. Phys., 2005, 123, 64107 CrossRef PubMed .
  23. R. A. Kendall, T. H. Dunning and R. J. Harrison, J. Chem. Phys., 1992, 96, 6796–6806 CrossRef CAS .
  24. H.-J. Werner and P. J. Knowles, J. Chem. Phys., 1985, 82, 5053–5063 CrossRef CAS .
  25. P. J. Knowles and H.-J. Werner, Chem. Phys. Lett., 1985, 115, 259–267 CrossRef CAS .
  26. D. A. Kreplin, P. J. Knowles and H. J. Werner, J. Chem. Phys., 2019, 150, 194106 CrossRef PubMed .
  27. D. A. Kreplin, P. J. Knowles and H. J. Werner, J. Chem. Phys., 2020, 152, 074102 CrossRef CAS PubMed .
  28. I. R. Ariyarathna, J. A. Leiding, A. J. Neukirch and M. C. Zammit, J. Phys. Chem. A, 2024, 128, 9412–9425 CrossRef CAS PubMed .
  29. J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8822–8824 CrossRef PubMed .
  30. A. D. Becke, Phys. Rev. A, 1988, 38, 3098–3100 CrossRef CAS PubMed .
  31. B. Miehlich, A. Savin, H. Stoll and H. Preuss, Chem. Phys. Lett., 1989, 157, 200–206 CrossRef CAS .
  32. F. J. Devlin, J. W. Finley, P. J. Stephens and M. J. Frisch, J. Phys. Chem., 1995, 99, 16883–16902 CrossRef CAS .
  33. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed .
  34. J. Tao, J. P. Perdew, V. N. Staroverov and G. E. Scuseria, Phys. Rev. Lett., 2003, 91, 146401 CrossRef PubMed .
  35. H. S. Yu, X. He and D. G. Truhlar, J. Chem. Theory Comput., 2016, 12, 1280–1293 CrossRef CAS PubMed .
  36. A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS .
  37. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS PubMed .
  38. P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627 CrossRef CAS .
  39. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 6671–6687 CrossRef CAS PubMed .
  40. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS .
  41. Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2007, 120, 215–241 Search PubMed .
  42. H. S. Yu, X. He, S. L. Li and D. G. Truhlar, Chem. Sci., 2016, 7, 5032–5051 RSC .
  43. M. A. Rohrdanz, K. M. Martins and J. M. Herbert, J. Chem. Phys., 2009, 130, 054112 CrossRef PubMed .
  44. T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS .
  45. J. D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106 CrossRef PubMed .
  46. E. Bremond and C. Adamo, J. Chem. Phys., 2011, 135, 024106 CrossRef PubMed .
  47. S. Kozuch and J. M. Martin, Phys. Chem. Chem. Phys., 2011, 13, 20104–20107 RSC .
  48. S. Kozuch and J. M. Martin, J. Comput. Chem., 2013, 34, 2327–2344 CrossRef CAS PubMed .
  49. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16, Gaussian Inc., Wallingford CT, 2016 Search PubMed .
  50. M. C. Zammit, J. A. Leiding, J. Colgan, W. Even, C. J. Fontes and E. Timmermans, J. Phys. B: At., Mol. Opt. Phys., 2022, 55, 184002 CrossRef CAS .
  51. J. K. Kirkland, S. N. Khan, B. Casale, E. Miliordos and K. D. Vogiatzis, Phys. Chem. Chem. Phys., 2018, 20, 28786–28795 RSC .
  52. N. M. S. Almeida, I. R. Ariyarathna and E. Miliordos, J. Phys. Chem. A, 2019, 123, 9336–9344 CrossRef CAS PubMed .
  53. A. Kramida, Y. Ralchenko and J. Reader, NIST Atomic Spectra Database (Version 5.3), National Institute of Standards and Technology, Gaithersburg, MD, 2015, https://physics.nist.gov/asd Search PubMed .
  54. R. S. Mulliken, Rev. Mod. Phys., 1932, 4, 1–86 CrossRef CAS .
  55. E. Wigner and E. E. Witmer, Z. Phys., 1928, 51, 859–886 CrossRef CAS .
  56. K. K. Irikura, W. A. Goddard and J. L. Beauchamp, Int. J. Mass Spectrom. Ion Processes, 1990, 99, 213–222 CrossRef CAS .
  57. I. R. Ariyarathna and E. Miliordos, Phys. Chem. Chem. Phys., 2018, 20, 12278–12287 RSC .
  58. I. R. Ariyarathna, C. Duan and H. J. Kulik, J. Chem. Phys., 2022, 156, 184113 CrossRef CAS PubMed .
  59. I. R. Ariyarathna and E. Miliordos, J. Quant. Spectrosc. Radiat. Transfer, 2020, 255, 107265 CrossRef CAS .
  60. I. R. Ariyarathna, Phys. Chem. Chem. Phys., 2024, 26, 21099–21109 RSC .
  61. I. R. Ariyarathna, Phys. Chem. Chem. Phys., 2024, 26, 22858–22869 RSC .
  62. A. J. Sauval and J. B. Tatum, Astrophys. J., Suppl. Ser., 1984, 56, 193–209 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: Fig. S1 illustrates the molecular orbitals of FeH+; Fig. S2 illustrates the μ of FeH+(15Δ) under various functionals of DFT; Table S1 lists the total DFT μ of FeH+(15Δ) and % DFT errors compared to CCSD(T) μ; Tables S2 and S3 list the TIPS fit coefficients of FeH+ and FeH2+; Table S4 lists the absolute energies of the electronic states of the FeH+; Table S5 lists the absolute energies and the spectroscopic parameters of FeH+ at CAS(8,7); Table S6 lists the absolute energies of the electronic states of the FeH2+. See DOI: https://doi.org/10.1039/d4cp03296a

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.