Next Article in Journal
A Comparative Evaluation of the Photosensitizing Efficiency of Porphyrins, Chlorins and Isobacteriochlorins toward Melanoma Cancer Cells
Next Article in Special Issue
Organoboron Complexes as Thermally Activated Delayed Fluorescence (TADF) Materials for Organic Light-Emitting Diodes (OLEDs): A Computational Study
Previous Article in Journal
Research Progress and Future Trends of Low Temperature Plasma Application in Food Industry: A Review
Previous Article in Special Issue
Shaped Microwave Field in a Three-Level Closed Loop Dense Atomic System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ab Initio Calculations on the Ground and Excited Electronic States of Thorium–Ammonia, Thorium–Aza-Crown, and Thorium–Crown Ether Complexes

Department of Chemistry and Biochemistry, Auburn University, Auburn, AL 36849-5312, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(12), 4712; https://doi.org/10.3390/molecules28124712
Submission received: 26 May 2023 / Revised: 9 June 2023 / Accepted: 9 June 2023 / Published: 12 June 2023

Abstract

:
Positively charged metal–ammonia complexes are known to host peripheral, diffuse electrons around their molecular skeleton. The resulting neutral species form materials known as expanded or liquid metals. Alkali, alkaline earth, and transition metals have been investigated previously in experimental and theoretical studies of both the gas and condensed phase. This work is the first ab initio exploration of an f-block metal–ammonia complex. The ground and excited states are calculated for Th0–3+ complexes with ammonia, crown ethers, and aza-crown ethers. For Th3+ complexes, the one valence electron Th populates the metal’s 6d or 7f orbitals. For Th0–2+, the additional electrons prefer occupation of the outer s- and p-type orbitals of the complex, except Th(NH3)10, which uniquely places all four electrons in outer orbitals of the complex. Although thorium coordinates up to ten ammonia ligands, octa-coordinated complexes are more stable. Crown ether complexes have a similar electronic spectrum to ammonia complexes, but excitations of electrons in the outer orbitals of the complex are higher in energy. Aza-crown ethers disfavor the orbitals perpendicular to the crowns, attributed to the N-H bonds pointing along the plane of the crowns.

1. Introduction

Isolated (gas phase) metal–ammonia complexes have been shown to host one or more diffuse electrons in their periphery [1,2]. Such complexes are called solvated electron precursors (SEPs) and the diffuse electrons reside at hydrogenic-type orbitals, which follow the energy order observed for the nuclear or Jellium shell models [1]. Specifically, the lowest-energy outer orbital is of s-type (dubbed 1s), followed by the 1p, 1d, 2s, 2p, and 1f orbitals [1]. This shell structure is universal and is independent of the central metal, including alkali metals, alkaline earth metals, boron, and transition metals [2,3,4,5]. The transition metal–ammonia complexes retain inner-valence d-orbitals of the metal [3,6,7,8]. Spectroscopic studies exist in the literature for lithium, sodium, magnesium, calcium, aluminum, vanadium, chromium, nickel, cobalt, copper, and silver [7,9,10,11,12,13,14,15,16,17,18,19,20]. In this work, we provide the first theoretical investigation of an f-block metal SEP, focusing on thorium–ammonia complexes.
Materials composed of SEPs (liquid or expanded metals) have been synthesized and characterized in detail for lithium [21,22], but have also been reported for multiple metals, including the lanthanides europium and ytterbium with the composition Eu(NH3)6 and Yb(NH3)6 [23]. Recently, such materials have been proposed as redox catalysts [24] and candidates for quantum information applications [25].
The first four ionization energies (IEs) of thorium (6.3067, 11.9, 20.0, and 28.8 eV) are remarkably lower than those of transition metals or other f-block elements [26]. For example, the same IE values for Eu, Yb, and V are in the ranges of 5.7–6.7, 11.3–14.6, 24.9–29.3, and 42.7–46.7 eV. Note that although the first two IEs are both comparable, the third IE of thorium is about 5 eV lower, while its fourth IE is more than 15 eV lower and of the same order of the third IE of other metals. As will be explained later, this fact in combination with the large coordination numbers for its first solvation shell (up to ten) [27,28] render ammonia ligands capable of stabilizing highly oxidized metal centers and displacing multiple electrons of thorium to the periphery of the thorium–ammonia complex.
As shown later, a Th(NH3)10 is feasible with a Th4+ center and four diffuse electrons. This is the largest number of diffuse electrons observed so far. However, the most stable structure, Th(NH3)8, has a Th3+ center and three diffuse electrons. To see if this electronic structure is characteristic of ammonia coordination only, we used two aza-crown ether ligands with four and five nitrogen atoms each, thus retaining the number of nitrogen atoms anchored to the metal. Such complexes have been studied before for lithium, sodium, and magnesium [29]. Finally, we replaced the aza-crown ethers with the corresponding crown ethers (NH groups replaced by O atoms) to observe the effect of the coordinating atom and the presence of N−H bonds in the electronic structure of the complex.
In the next section, we detail the computational methods employed currently. Then we discuss our findings, and finally summarize our results.

2. Computational Details

The geometry optimizations were performed at the density functional theory (DFT) level using the CAM-B3LYP functional combined with the cc-pVDZ, cc-pVDZ, aug-cc-pVDZ, and cc-pVDZ-PP basis sets for carbon, nitrogen, hydrogen, and thorium centers, respectively [30,31,32,33]. The latter basis set is supplemented with the relative pseudopotential [34]. The employed functional was shown to provide accurate geometries of the MP2 and CCSD(T) level for other metal–ammonia complexes [35]. Gaussian 16 was invoked for these calculations [36]. Every optimized structure bears real harmonic vibrational frequencies; geometries and energies are given in the Supplementary Materials (SM).
The Th(NH3)84+, Th(NH3)104+, Th(12C4N)24+, Th(15C5N)24+, Th(12C4O)24+ crown ether structures were used for the subsequent multi-reference calculations: 12C4X/15C5X denote aza-crown ethers with 12/15 the total number of non-hydrogen atoms and 4/5 the number of nitrogen (X=N) or oxygen (X=O) atoms. These species are closed shell, and they adopt the highest possible symmetry compared to the trications, dications, monocations, or neutral counterparts.
The active space for the CASSCF (complete active space self-consistent field) calculations generally includes both inner (6d, 7f) and outer (1s, 1p, 1d) diffuse orbitals, but has been adjusted to balance the computational cost based on the population of the various orbitals. For example, the inner orbitals are less and less populated when more electrons are added. The exact active space has been optimized by multiple trial-and-error attempts for each molecular species and is provided below accordingly. The subsequent CASPT2 (CASSCF + second-order perturbation theory) [37] calculations included the dynamic correlation from the ammonia/ammine/oxygen lone pairs as well. CASPT2 calculations have been shown to be sensitive to the used active space [38]. CASPT2 calculations with similar active spaces have been shown to agree perfectly with electron propagator techniques and EOM-CCSD calculations for other metal–ammonia complexes [1,3,39]. Due to the high computational cost, only the s and p functions of nitrogen and carbon centers from the cc-pVDZ sets are included. This is expected to have a minimal effect on computed excitation energies (<0.1 eV), as excitations occur only within the Th valence space and the peripheral orbitals of the complex, which are described predominantly by the hydrogen atom basis functions [40]. The MOLPRO suite of codes [41] is used, specifically, the internally contracted version of CASPT2 (CASPT2c) [42]. A level shift value of 0.2 a.u. and IPEA shift of 0.25 a.u. were used to resolve linear dependence issues [43].

3. Results and Discussion

We first optimized the geometries for the Th(NH3)n4+ and Th(NH3)n3+ species for n = 1–10. We considered only the isomers where all ammonia ligands are coordinated to thorium. These systems have simple electronic structure (no or one unpaired electron) and are described properly with single determinantal methods such as DFT. The optimized geometries for trications are shown in Figure 1, along with the singly occupied molecular orbital (SOMO). for each structure. The ground state of Th3+ is 2F(7f1) and it stays in the 2F state only for one ammonia ligand. After coordination of additional ammonias results in the population of a ~6dz2-type SOMO up to n = 9. For Th(NH3)103+, the SOMO becomes a ~6dxz type.
The sequential dissociation energy De for the ammonia ligands, Th(NH3)n4+,3+,0 → Th(NH3)n−14+,3+,0 + NH3 at CAM-B3LYP (De = E[Th(NH3)n−14+,3+,0] + E[NH3] − E[Th(NH3)n4+,3+,0], where E[X] denotes the equilibrium energy of species X), is plotted with respect to n in Figure 2. We faced insurmountable technical/convergence issues for several cationic and dicationic species, likely due to their complex electronic structure (see below), and thus these species are not included in the figure. The ground state is a singlet and doublet for 4+ and 3+ charges, respectively, and is a singlet for 1 ≤ n ≤ 4 and triplet for 5 ≤ n ≤ 10 in the neutral complexes (see Supplementary Materials). The binding energy drops sharply with increase in n for the +4 charge (on average by 18 kcal/mol per ammonia ligand). For all species, but more evidently for the trications, there is a sudden drop going from n = 8 to n = 9. For the neutral, the De range is 15 ± 6 kcal/mol for 1 ≤ n ≤ 8, which becomes 7.3 and 5.7 kcal/mol for n = 9 and 10, and there is a slight increase in De from n = 7 to n = 8 (9.8 to 13.0 kcal/mol). Therefore, we believe that the most prominent structure of a thorium expanded metal will be the octacoordinated Th(NH3)8, unlike the hexacoordinate Eu(NH3)6 and Yb(NH3)6 [23].
Using the Th(NH3)84+ structure (D4d actual point group, C2v computational point group), we performed CASSCF and CASPT2 calculations for all species with charges from 3+ to 0 in order to elucidate their electronic structure and explain the convergence issues in DFT for the intermediate charges (1+ and 2+). The geometry and CASSCF active orbitals for Th(NH3)83+ are depicted in Figure 3 and include the inner 6d and 7f orbitals of thorium and the outer 1s, 1p, and 1d of the whole complex (1e/21 orbitals). The energies for the 1s1, 1p1, 1d1, 6d1, and 7f1 states are listed in Table 1. The lowest energy states, X ˜ 2A1, 12E2, and 12E3, correspond to 6d orbitals of thorium. The ground state has a (6dz2)1 configuration and is well separated from the other 6d states and higher excited states by ≥1.3 eV. The first few 7f states (12B2 and 12E1) appear next at ~2.0 eV, followed closely by the first electronic state with an outer electron (22A1; 1s1). At ~1.0 eV higher are the 1p1 states (22B2, 22E1) and one more degenerate 7f state (22E2). The last 7f state is at ~3.5 eV. All outer 1d states lie between 4.2 and 4.9 eV. Note that the CASSCF and CASPT2 excitation energies differ by less than 0.17 eV.
The addition of one more electron results in a highly multi-reference wave function, explaining the convergence issues with DFT. CASSCF calculations (2e/15 orbitals) of the dication includes two electrons in fifteen orbitals (1s, 1p, 1d, 6dz2, and five orbitals of 7f/6d character). The results show the ground state X ˜ 1A1 of Th(NH3)82+ has two major electronic configurations, X ˜ A   1 1 0.84   1 s 2 0.49   6 d z 2 2 , and it is just 0.06 eV lower than the A ˜ A   3 1 state, A ˜ A   3 1 0.99   1 s 1 6 d z 2 1 . The singlet state can be seen as a mixture of a closed-shell singlet 1s2 (49%) and an open-shell singlet 1s1(6dz2)1 (51%). The percentages are estimated from the 0.84 and 0.49 coefficients [44].
The next electronic states (1,3B2, 1,3E1) lie between 0.72 and 0.93 eV and include all combinations of 1s11p1 and (6dz2)11p1. The first 1,3E2 states follow at 1.28 eV with configurations 1s1(f/d)1 and (6dz2)11(f/d)1, where f/d refers to the orbitals produced from mixing 7f/6d orbital functions. We calculated ten more states, where electrons are promoted to outer 1d orbitals and Th 7f orbitals: 1.33 (1A1), 1.43 (1A1), 1.75 (1A1), 1.76 (3E3), 1.80 (1E1), 1.82 (3E2), 1.83 (3A1), 1.85 (1E2), 1.89 (3E1), 1.90 (1E3) eV.
Upon adding one more electron to form Th(NH3)81+, the ground state X ˜ 2A1 still has two major configurations involving the 1s and 6dz2 orbitals, X ˜ A   2 1 0.84   6 d z 2 1 1 s 2 0.34   6 d z 2 2 1 s 1 , comprising 71% and 12% of the wave function, respectively. Next, four quasi-degenerate states (two quartet and two doublet states) lie at 0.53 ± 0.03 eV (see Table 2). The quartets are single reference states with 95% of the wave function produced from some (6dz2)11s11p1 configuration. The same percentages for the doublet states are 45% (12E1) and 41% (12B2) after adding the contribution from all three Slater determinants with different spin-up/spin-down combinations of (6dz2)11s11p1. The next-largest term (34% for 12E1 and 37% for 12B2) corresponds to the 1s21p1 configuration, which has three diffuse outer electrons, and finally a 15% portion (for both 12E1 and 12B2) belongs to (6dz2)21p1, where there is only one outer diffuse electron.
The above results were obtained with a 3e/10 orbital active space, with the 10 orbitals being the 6dz2, 1s, 1p, and 1d. According to these calculations, the higher-energy states are extremely multi-reference lying above 1.0 eV at the CASSCF level of theory. Compared to the dicationic and tricationic species, Th(NH3)81+ has a higher density of low-lying electronic states with five states in the first 0.6 eV. Only one or two states are present in this energy range for Th(NH3)83+ and Th(NH3)82+. Finally, as noted for the trication, the CASSCF and CASPT2 excitation energies are also in perfect agreement here (within 0.04 eV).
Moving to the neutral species (4e/10 orbitals), the fourth valence electron occupies a 1p orbital, resulting in the (6dz2)11s21p1 configuration of the X ˜ 3E1 ground state, which is 72% of the wave function. This state is comparable to the addition of a 1p electron to (6dz2)11s2, the major component of ground state X ˜ 2A1 in Th(NH3)81+. The other component of X ˜ 2A1 is a (6dz2)21s1; addition of a 1p electron to this configuration constitutes only 2% of the ground state for Th(NH3)80. The 1px,y orbitals are populated first and the corresponding singlet and triplet states (1,3E1) are practically degenerate (see Table 2). The states 11,3B2 pertain to occupation of 1pz and are higher by <0.1 eV.
The next batch of electronic states, 13,5A2 and 13,5E3, have a (6dz2)11s11p2 character by 72% (S = 1) and 92% (S = 2). In every case, the 1p2 electrons couple into a triplet spin state. All lie in the range between 0.30 and 0.39 eV (see Table 2). Coupling of the 1p2 electrons into a singlet spin multiplicity generates the largest portion (from 42% to 58%) of the last six states of Table 2. Resembling the 1D state of carbon, there are five 1p2 components belonging to the E2, E3 and A1 irreducible representations. The second-largest contribution to the wave function of these states (26–34%) pertains to (6dz2)11s21d1, which also has five components of the same irreducible representations. The excitation energies for these six states are 0.54−0.63 eV. Overall, the neutral species have the most “dense” electronic spectrum, with 14 states present within 0.63 eV. As in the cation, no 6d or 7f orbital (excluding 6dz2) is occupied within the states studied, as additional electrons (relative to Th3+) favor occupation of the outer orbitals. The active space used is the same as in Th(NH3)81+ (4e/10 orbitals).
In all octacoordinated thorium complexes, there is one inner electron in 6dz2, which is perturbed by the molecular skeleton in order to avoid all Th-N coordination bonds (see Figure 3). This orbital has a substantial metallic/non-bonding character. The other 6d orbitals have some σTh−N* anti-bonding character and are higher in energy by at least 1.3 eV (see Table 1). As such, the addition of further ammonia ligands is expected to destabilize the 6dz2 orbital, as these must approach along the z-direction, inducing a similar anti-bonding character. This is evidenced by our study of Th(NH3)10, where the addition of two ammonia ligands results in the promotion of the 6dz2 electron to an outer 1s or 1p orbital (see Figure 4).
Specifically, we performed multi-reference calculations for all Th(NH3)103+,2+,1+,0 species (4e/10 orbitals, 3e/14 orbitals, 2e/9 orbitals, 1e/9 orbitals) and found the ground states of Th(NH3)101+,0 have no inner electrons. Instead, they adopt 1s11p2 (S = 3/2) and 1s21p2 (S = 1) configurations. The ground states of Th(NH3)102+,3+ retain a metallic electron; this electron occupies an orbital and is composed of a mixture of 6d/7f orbitals in order to minimize its amplitude along the Th-N bonds. However, the 6d/7f → 1s1 excitation for Th(NH3)103+ occurs at 0.72 eV and for Th(NH3)102+ at 0.14 eV compared to 2.05 eV for Th(NH3)83+ These results illustrate how the decacoordinate complex destabilizes the metallic electronic states and favors the promotion of electrons to the outer orbitals. The CASSCF active space used for each system included (number of electrons/number of orbitals) 1/10, 2/14, 3/9, and 4/9 for Th(NH3)103+,2+,1+,0, respectively; at CASPT2, correlation of all NH3 lone pairs was also included and the geometries used were of C2 symmetry. Further, it appears that the displacement of the 6dz2 electron to the periphery of the complex reduces the binding energy of the ninth and tenth ammonia ligands (see Figure 2).
In an attempt to identify more stable complexes, we then studied the Th(12C4N)2 and Th(15C5N)2 aza-crown ethers (shown in Figure 5), where the “top” and “bottom” ammonia ligands of Figure 4 are connected via a hydrocarbon bridge with two carbon atoms (tetradentate ligands). Binding energies of the aza-crown ligands were calculated in two ways: (1) by dissociating the two ligands from Th, but keeping them fixed in the optimum geometry of the complex, and (2) by optimizing the dissociated ligands. For the constrained dissociation, the binding energy per Th-N bond in Th(12C4N)24+,3+,0 is 114.0, 68.8, and 13.7 kcal/mol, respectively. The corresponding binding energies of the unconstrained geometry are 108.9, 63.7, and 8.6 kcal/mol. The drop in binding energy illustrates a steric strain introduced to the aza-crown rings during Th coordination of the order of ~5 kcal/mol. At CAM-B3LYP, the binding energy per Th-N bond in Th(NH3)84+,3+,0 is 103.5, 63.1, and 13.6 kcal/mol. From this, we see the energy of the 4+ and 3+ aza-crowns is comparable to that of the Th(NH3)n0 complexes. However, the polydentate binding results in a more stable complex, as the removal of one aza-crown ligand is significantly more difficult, e.g., 4 × 8.6 = 34.4 kcal/mol. A similar situation is found for the Th(15C5N)24+,3+,0 complexes: the per Th-N bond energies are 89.2, 53.2, and 12.2 kcal/mol for Th(NH3)104+,3+,0. The corresponding values for Th(15C5N)24+,3+,0 are 93.0, 55.0, and 7.1 kcal/mol, respectively. Removal of one 15C5N ligand requires a minimum of 28.4 kcal/mol (4 × 7.1). In this case, we were able to obtain the CAM-B3LYP binding energies for all charges from 4+ to 0, which are listed in Table 3. This increased stability facilitates the formation of the neutral, decacoordinate Th complex, where it was unfavorable with ammonia ligands. This begs the question: How does the electronic structure change when replacing ammonia with aza-crown ethers, and does Th(15C5N)2 have four peripheral electrons like Th(NH3)10?
To answer these questions, we performed CASSCF calculations. These were performed using the Th(12C4N)24+ and Th(15C5N)24+ optimized geometries, as they bear the highest possible symmetry (C2 point group; see Figure 5 for the C2 axis). Under this symmetry, the 1px and 1py orbitals remain quasi-degenerate, but the 1pz is destabilized considerably and is not populated in the low-lying electronic states (see Figure 5 for the z-axis). The same happens for the 1d orbitals, where only the 1dxy and 1dx2−y2 orbitals participate in the low-lying states; all orbitals with amplitudes along the z-axis shift to higher energies, as has been seen for lithium–crown ether complexes [45]. Therefore, the active space has been adjusted to exclude these orbitals. Specifically, the active space used for the aza-crown ethers is (number of electrons/number of orbitals): 1/13, 2/8, 3/6, 4/6 and 1/15, 2/9, 3/6, 4/6 for Th(12C4N)23+,2+,1+,0 and Th(15C5N)23+,2+,1+,0, respectively.
The active orbitals and excitation energies for Th(12C4N)23+ are given in Figure 5 and Table 1, respectively. Similarly to Th(NH3)83+, the ground state retains its (6dz2)1 character and is still followed by the (6dxy)1/(6dx2−y2)1 states at around 1.04/1.02 eV (~0.3 eV lower). However, the inner 6dxz/yz and outer 1pz orbitals are no longer populated in the low-lying states, while the 1s outer orbital is polarized towards the xy plane. There are three reasons for these observed differences: (1) the Th4+ charge is less screened along the xy plane, (2) all N-H bonds (known to solvate electrons) point in the ±x/y directions, and (3) many of the C-H bonds (known to “repel” diffuse electrons) [5,25] point in the ±z directions. These effects shift the energy of the 1s1 state by ~0.7 eV from 2.05 to 2.76 eV. The 1px,y1 states are practically unaffected, and among the 1f orbitals, only the energy of the (7fx(x2−3y2))1 and (7fy(3x2−y2))1 states changes considerably, shifting from 3.59 to 1.86 ± 0.02 eV.
Going from Th(12C4N)23+ to Th(15C5N)23+, the overall picture remains the same, but now the inner 6dxz/yz orbitals are populated and are highly mixed with the 7f ones. It appears that the larger ring provides more space along the xz/yz diagonals, lowering the energy of the 6dxz/yz1 states. This is the reason that the active space of Th(15C5N)23+ includes two more orbitals (see above). These nine states (seven 7f1 + two 6dxz/yz1) cover an energy range between 1.48 and 3.08 eV. The (6dxy)1/(6dx2−y2)1 states are located at 1.18/1.22 eV, and the outer 1s1 and 1px,y1 states move to 3.23 and 3.44/3.52 eV (~0.4 and ~0.2 eV higher). In all the three complexes Th(NH3)83+, Th(12C4N)23+, and Th(15C5N)23+ CASPT2 (vs. CASSCF) stabilizes the 1s1 and 1p1 states by about 0.3 eV.
The ground state of the dicationic thorium–aza-crown–ether complexes is a multi-reference singlet electronic state involving a mixture of 1s2, (6dz2)2, and 1s1(6dz2)1 configurations. This is followed by the single reference triplet 1s1(6dz2)1 state. These states are 0.07 and 0.20 eV apart for Th(12C4N)22+ and Th(15C5N)22+, respectively. This mirrors the first two states of Th(NH3)82+. In Th(NH3)82+, the next states are six states with 1s11p1 and (6dz2)11p1 character, but for the aza-crown ether complexes, only four of these states remain below 1.0 eV (0.28–0.38 eV for the tetra- and 0.59–0.71 eV for the penta-amine ligands), as the 1pz orbital is highly destabilized.
The low-lying electronic states of the aza-crown monocationic complexes are identical to those of Th(NH3)8+ (see Table 2) if we exclude the 2,4B2 states, which populate the 1pz orbital. The degeneracy of the 2,4E1 states is lifted due to the lower symmetry: the two components of the doublet states are at 0.02/0.26 and 0.09/0.43 eV for Th(12C4N)2+ and Th(15C5N)2+, respectively, and those of the quartet states are at 0.04/0.29 and 0.14/0.57 eV.
Finally, for the neutral aza-crown complexes, only the states corresponding to X ˜ 3E1, X ˜ ΄ 1E1, 13A2, 15A2, 11E2, and 13E2 of Th(NH3)8 (see Table 2) “survive”, since the 1pz, 1dz2, and 1dxz/yz outer orbitals do not contribute. There are five nearly degenerate states within 0.03/0.13 eV for Th(12C4N)2/Th(15C5N)2. These are very multi-reference states (largest coefficient 0.48) with (6dz2)11s21px,y1 and (6dz2)11s11px11py1 characters, with the 3B always the ground state, followed by the 1A (at 0.00/0.09 eV), 1B (at 0.01/0.10 eV), 3A (at 0.02/0.13 eV), and 5B (at 0.03/0.12 eV) states. Notice the 5B [(6dz2)11s11px11py1] state is now closer to the ground state, since it has two electrons in the 1p orbitals, which are stabilized over 1s (see discussion on Th(NH3)83+ above and Table 1). The next four states are also highly multi-reference, involving the (6dz2)11s11p2 and (6dz2)11s2(1dxy,x2−y2)1 configurations, and are located at 0.74–0.77/0.24–0.35 eV higher than the ground state.
To answer our earlier questions: the replacement of ammonia ligands with aza-crown ethers has a significant impact on the electronic structure and Th(15C5N)2 has only three peripheral electrons, unlike Th(NH3)10, which has four. While Th(NH3)10 hosts four outer electrons (1s21p2) in the ground state, Th(15C5N)2 favors the (6dz2)11sm1px,yn (m + n = 3) configurations. The main reason is that two ammonia ligands are placed along the z-axis, which disfavors the presence of a 6dz2 electron and promotes this electron to the outer 1s orbital (see Figure 4). The two neutral octa-coordination complexes, Th(NH3)8 and Th(12C4N)2, have three outer electrons (1s21p1) and one inner 6dz2 electron in their ground states. For the three reasons discussed above and symmetry lowering (D4d to C2), only the 1px,y and 1dxy,x2−y2 (unlike 1pz and 1dz2,xz,yz) participate in the low-lying states.
Finally, we performed calculations for the crown ether Th(12C4O)2q and Th(15C5O)2q (q = +4, +3, +2, +1, 0) complexes. Such complexes are the building units of electrides and our goal is to provide insights for electrides made of oxygen- and nitrogen-based crown ethers. Our CAM-B3LYP geometries and energies are listed in the Supplementary Materials and the binding energies are reported in Table 3. The per-bond binding energies are identical for the two kinds of ethers within the accuracy of our calculations.
Regarding the excited states of Th(12C4O)23+, the electronic structure more closely resembles Th(NH3)83+ than Th(12C4N)23+ owing to the greater equivalence of the z-direction to the x-,y-, as demonstrated by the relative energy of 1pz1 to 1px,y1 (3.36, 3.60, 3.6 eV). The absence of the N-H bonds along the xy plane appears to make the space more isotropic. Calculated excitation energies are listed in Table 1 (1e/20 orbitals). The excited 6d states are higher in energy than those of Th(NH3)83+ and the 7f orbitals are highly mixed. The outer 1s and 1p states are higher than the ammonia and aza-crown complexes, but the 1d states are in the same energy range as Th(NH3)83+.
The addition of an electron (CASSCF active space 2e/12 orbitals) leads to two singlet states of (6dz2)2 and (6dz2)11s1 mixed character at 0.00 and 0.43 eV, while the triplet (6dz2)11s1 state is located in between (0.21 eV). The next six states correspond to (6dz2)11p1. The three triplets have energies of 0.31, 0.43, 0.55 eV and the three singlets 0.43, 0.51, 0.56 eV. The addition of further electrons (CASSCF active spaces 3e/6 orbitals, 4e/6 orbitals) yields highly multi-reference wave functions. The ground state of Th(12C4O)21+ is mainly (6dz2)11s2 followed by the (6dz2)11s11p1 states lying between 0.04 and 0.18 eV, while the ground state of Th(12C4O)2 is (6dz2)11s21p1 followed by (6dz2)11s11p2 states spanning an energy range of 0.05–0.25 eV. In every case, the low-lying states have only one inner electron and up to three outer electrons, similarly to the other complexes.

4. Conclusions

This work is the first systematic investigation of thorium SEPs. We used high-level electronic structure methods to study the thorium–ammonia, thorium–aza-crown, and thorium–crown ether complexes. We found that the octa-coordinate complexes are more stable, suggesting that the stoichiometry of thorium expanded metals will be Th(NH3)8. In all cases, except Th(NH3)10, there is a Th3+ center and zero/one (1s1)/two (1s2)/three (1s21p1) outer electrons for 3+/2+/1+/0 charged complexes. In the case of Th(NH3)10 there is a Th4+ center with four diffuse electrons. The nature and energetics of the outer orbitals change considerably with the ligand type. Ammonia makes isotropic structures keeping the near degeneracy of the 1p and 1d orbitals. The aza-crown ethers have N−H bonds parallel to the crowns (xy dimension) and thus disfavor orbitals lying along the z-axis. The replacement of NH with O (plain metal–crown ether complexes) restores partially the isotropic environment, but increases the energy of the 1s1 and 1p1 states of the trications. For all species, although the 1s1/1p1 states are higher in energy than 6d1/7f1 states, they are populated first when at least a second electron is added. The wave functions become extremely multi-determinantal when two or more active electrons are present, and thus multi-reference methods are sine qua non for these systems. More work on other lanthanide/actinide complexes in the gas and condensed phases is in progress.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124712/s1, Figure S1: Structures of aza-crown ether 12C4N: (a) fully optimized, (b) fixed at the geometry in the Th(12C4N)24+ complex.; Table S1: Energies (hartrees) for the Th(NH3)n = 0–104+,3+,0 species. The spin for the charged species is singlet (4+) and doublet (3+), while for the neutral species both the lowest singlet (S = 0) and triplet (S = 1) states are reported.; Table S2: Cartesian coordinates (Å) for the Th(NH3)n = 1–104+,3+,0 species. The spin for the charged species is singlet (4+) and doublet (3+), while for the neutral species both the lowest singlet (S = 0) and triplet (S = 1) states are reported.; Table S3: Energies (hartrees) for the Th(12C4X)2q and Th(15C5X)2q species (q = 4+, 3+, 2+, 1+, 0 and X = N, O). Different spin states are reported (spin is given in parenthesis).; Table S4: Cartesian coordinates (Å) for the Th(12C4X)2q and Th(15C5X)2q species (q = 4+, 3+, 2+, 1+, 0 and X = N, O). Different spin states are reported (spin is given in parenthesis).

Author Contributions

Conceptualization, E.M.; methodology, E.M. and B.A.J.; software, E.M., B.A.J. and Z.L.; validation, E.M. and B.A.J.; investigation, E.M., B.A.J. and Z.L.; resources, E.M.; data curation, E.M.; writing—original draft preparation, E.M.; writing—review and editing, E.M. and B.A.J.; visualization, E.M.; supervision, E.M.; project administration, E.M. and B.A.J.; funding acquisition, E.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the United States Department of Energy–Basic Energy Sciences through the Chemical Sciences, Geosciences, and Biosciences (CSGB) subdivision for Heavy Elements Chemistry with Award DE-SC0019177, and the National Science Foundation under grant CHE-1940456.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

DFT/CAM-B3LYP geometries and energies for all structures are listed in Tables S1–S4 of the Supplementary Materials File.

Acknowledgments

The authors are indebted to Auburn University (AU) for financial support. EM is especially grateful to the donors of the James E. Land endowment, and Anne Elizabeth Gorden (Texas Tech University) for the valuable discussions. This work was completed with resources provided by the Auburn University Hopper and Easley Clusters. The authors would like to acknowledge the contributions of Jared Stinson for his assistance on the Th–crown ether complexes.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ariyarathna, I.R.; Khan, S.N.; Pawłowski, F.; Ortiz, J.V.; Miliordos, E. Aufbau Rules for Solvated Electron Precursors: Be(NH3)40,± Complexes and Beyond. J. Phys. Chem. Lett. 2018, 9, 84–88. [Google Scholar] [CrossRef] [Green Version]
  2. Ariyarathna, I.R.; Pawłowski, F.; Ortiz, J.V.; Miliordos, E. Molecules mimicking atoms: Monomers and dimers of alkali metal solvated electron precursors. Phys. Chem. Chem. Phys. 2018, 20, 24186–24191. [Google Scholar] [CrossRef]
  3. Almeida, N.M.S.; Pawłowski, F.; Ortiz, J.V.; Miliordos, E. Transition-metal solvated-electron precursors: Diffuse and 3d electrons in V(NH3)0,±6. Phys. Chem. Chem. Phys. 2019, 21, 7090–7097. [Google Scholar] [CrossRef]
  4. Ariyarathna, I.R.; Almeida, N.M.S.; Miliordos, E. Stability and Electronic Features of Calcium Hexa-, Hepta-, and Octa-Coordinated Ammonia Complexes: A First-Principles Study. J. Phys. Chem. A 2019, 123, 6744–6750. [Google Scholar] [CrossRef]
  5. Jordan, Z.; Khan, S.N.; Jackson, B.A.; Miliordos, E. Can boron form coordination complexes with diffuse electrons? Evidence for linked solvated electron precursors. Electron. Struct. 2022, 4, 015001. [Google Scholar] [CrossRef]
  6. Almeida, N.M.S.; Miliordos, E. Electronic and structural features of octa-coordinated yttrium–ammonia complexes: The first neutral solvated electron precursor with eight ligands and three outer electrons. Phys. Chem. Chem. Phys. 2019, 21, 7098–7104. [Google Scholar] [CrossRef]
  7. Jackson, B.A.; Miliordos, E. Electronic and geometric structure of cationic and neutral chromium and molybdenum ammonia complexes. J. Chem. Phys. 2021, 155, 014303. [Google Scholar] [CrossRef]
  8. Khan, S.N.; Miliordos, E. Scandium in Neutral and Positively Charged Ammonia Complexes: Balancing between Sc2+ and Sc3+. J. Phys. Chem. A 2020, 124, 4400–4412. [Google Scholar] [CrossRef]
  9. Takasu, R.; Misaizu, F.; Hashimoto, K.; Fuke, K. Microscopic Solvation Process of Alkali Atoms in Finite Clusters:  Photoelectron and Photoionization Studies of M(NH3)n and M(H2O)n (M = Li, Li-, Na-). J. Phys. Chem. A 1997, 101, 3078–3087. [Google Scholar] [CrossRef]
  10. Brockhaus, P.; Hertel, I.V.; Schulz, C.P. Electronically excited states in size-selected solvated alkali metal atoms. III. Depletion spectroscopy of Na(NH3)n-clusters. J. Chem. Phys. 1999, 110, 393–402. [Google Scholar] [CrossRef]
  11. Lee, J.I.; Sperry, D.C.; Farrar, J.M. Spectroscopy and reactivity of size-selected Mg+-ammonia clusters. J. Chem. Phys. 2004, 121, 8375–8384. [Google Scholar] [CrossRef] [Green Version]
  12. Mune, Y.; Ohashi, K.; Iino, T.; Inokuchi, Y.; Judai, K.; Nishi, N.; Sekiya, H. Infrared photodissociation spectroscopy of [Al(NH3)n]+ (n = 1–5): Solvation structures and insertion reactions of Al+ into NH3. Chem. Phys. Lett. 2006, 419, 201–206. [Google Scholar] [CrossRef] [Green Version]
  13. Inoue, K.; Ohashi, K.; Iino, T.; Judai, K.; Nishi, N.; Sekiya, H. Coordination and solvation of copper ion: Infrared photodissociation spectroscopy of Cu+(NH3)n (n = 3–8). Phys. Chem. Chem. Phys. 2007, 9, 4793–4802. [Google Scholar] [CrossRef]
  14. Salter, T.E.; Ellis, A.M. Structures of Small Li(NH3)n and Li(NH3)n+ Clusters (n = 1−5):  Evidence from Combined Photoionization Efficiency Measurements and ab Initio Calculations. J. Phys. Chem. A 2007, 111, 4922–4926. [Google Scholar] [CrossRef] [PubMed]
  15. Inoue, K.; Ohashi, K.; Iino, T.; Sasaki, J.; Judai, K.; Nishi, N.; Sekiya, H. Coordination structures of the silver ion: Infrared photodissociation spectroscopy of Ag+(NH3)n (n = 3–8). Phys. Chem. Chem. Phys. 2008, 10, 3052–3062. [Google Scholar] [CrossRef]
  16. Ohashi, K.; Inoue, K.; Iino, T.; Sasaki, J.; Judai, K.; Nishi, N.; Sekiya, H. A molecular picture of metal ion solvation: Infrared spectroscopy of Cu+(NH3)n and Ag+(NH3)n in the gas phase. J. Mol. Liq. 2009, 147, 71–76. [Google Scholar] [CrossRef]
  17. Imamura, T.; Ohashi, K.; Sasaki, J.; Inoue, K.; Furukawa, K.; Judai, K.; Nishi, N.; Sekiya, H. Infrared photodissociation spectroscopy of Co+(NH3)n and Ni+(NH3)n: Preference for tetrahedral or square-planar coordination. Phys. Chem. Chem. Phys. 2010, 12, 11647–11656. [Google Scholar] [CrossRef]
  18. Koga, N.; Ohashi, K.; Furukawa, K.; Imamura, T.; Judai, K.; Nishi, N.; Sekiya, H. Coordination and Solvation of V+ with Ammonia Molecules: Infrared Photodissociation Spectroscopy of V+(NH3)n (n=4–8). Chem. Phys. Lett. 2012, 539–540, 1–6. [Google Scholar] [CrossRef]
  19. Albaqami, M.D.; Ellis, A.M. Infrared spectroscopy of Ca(NH3)n complexes. Chem. Phys. Lett. 2018, 706, 736–740. [Google Scholar] [CrossRef]
  20. Kozubal, J.; Heck, T.R.; Metz, R.B. Vibrational Spectroscopy of Cr+(NH3)n (n = 1–6) Reveals Coordination and Hydrogen-Bonding Motifs. J. Phys. Chem. A 2019, 123, 4929–4936. [Google Scholar] [CrossRef]
  21. Seel, A.G.; Swan, H.; Bowron, D.T.; Wasse, J.C.; Weller, T.; Edwards, P.P.; Howard, C.A.; Skipper, N.T. Electron Solvation and the Unique Liquid Structure of a Mixed-Amine Expanded Metal: The Saturated Li–NH3–MeNH2 System. Angew. Chem. Int. Ed. 2017, 56, 1561–1565. [Google Scholar] [CrossRef] [Green Version]
  22. Seel, A.G.; Zurek, E.; Ramirez-Cuesta, A.J.; Ryan, K.R.; Lodge, M.T.J.; Edwards, P.P. Low energy structural dynamics and constrained libration of Li(NH3)4, the lowest melting point metal. Chem. Comm. 2014, 50, 10778–10781. [Google Scholar] [CrossRef]
  23. Glaunsinger, W.S.; Von Dreele, R.B.; Marzke, R.F.; Hanson, R.C.; Chieux, P.; Damay, P.; Catterall, R. Structures and properties of metal-ammonia compounds on the trail of a new ammonia geometry. J. Phys. Chem. 1984, 88, 3860–3877. [Google Scholar] [CrossRef]
  24. Jackson, B.A.; Miliordos, E. Simultaneous CO2 capture and functionalization: Solvated electron precursors as novel catalysts. Chem. Comm. 2022, 58, 1310–1313. [Google Scholar] [CrossRef]
  25. Jackson, B.A.; Dale, S.G.; Camarasa-Gómez, M.; Miliordos, E. Introducing Novel Materials with Diffuse Electrons for Applications in Redox Catalysis and Quantum Computing via Theoretical Calculations. J. Phys. Chem. C 2023, 127, 9295–9308. [Google Scholar] [CrossRef]
  26. Haynes, W.M. CRC Handbook of Chemistry and Physics, 93rd ed.; Taylor & Francis: Boca Raton, FL, USA, 2012. [Google Scholar]
  27. Tutson, C.D.; Gorden, A.E.V. Thorium coordination: A comprehensive review based on coordination number. Coord. Chem. Rev. 2017, 333, 27–43. [Google Scholar] [CrossRef] [Green Version]
  28. Blanchard, F.; Rivenet, M.; Vigier, N.; Hablot, I.; Grandjean, S.; Abraham, F. Solid State Chemistry of Ten-Fold Coordinate Thorium(IV) Complexes with Oxalates in the Presence of Ammonium and Hydrazinium Ions. Cryst. Growth Des. 2018, 18, 4593–4601. [Google Scholar] [CrossRef]
  29. Ariyarathna, I.R. Superatomic Chelates: The Cases of Metal Aza-Crown Ethers and Cryptands. Inorg. Chem. 2022, 61, 579–585. [Google Scholar] [CrossRef] [PubMed]
  30. Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  31. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  32. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. IV. Calculation of static electrical response properties. J. Chem. Phys. 1994, 100, 2975–2988. [Google Scholar] [CrossRef] [Green Version]
  33. Peterson, K.A. Correlation consistent basis sets for actinides. I. The Th and U atoms. J. Chem. Phys. 2015, 142, 074105. [Google Scholar] [CrossRef] [PubMed]
  34. Weigand, A.; Cao, X.; Hangele, T.; Dolg, M. Relativistic Small-Core Pseudopotentials for Actinium, Thorium, and Protactinium. J. Phys. Chem. A 2014, 118, 2519–2530. [Google Scholar] [CrossRef] [PubMed]
  35. Ariyarathna, I.R.; Pawłowski, F.; Ortiz, J.V.; Miliordos, E. Aufbau Principle for Diffuse Electrons of Double-Shell Metal Ammonia Complexes: The Case of M(NH3)4@12NH3, M = Li, Be+, B2+. J. Phys. Chem. A 2020, 124, 505–512. [Google Scholar] [CrossRef] [PubMed]
  36. Frisch, M.J.; Schlegel, G.W.T.H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; Li, X.; et al. Gaussian 16; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  37. Ghigo, G.; Roos, B.O.; Malmqvist, P.A. A Modified Definition of the Zeroth-Order Hamiltonian in Multiconfigurational Perturbation Theory (CASPT2). Chem. Phys. Lett. 2004, 396, 142–149. [Google Scholar] [CrossRef]
  38. Azizi, Z.; Roos, B.O.; Veryazov, V. How accurate is the CASPT2 method? Phys. Chem. Chem. Phys. 2006, 8, 2727–2732. [Google Scholar] [CrossRef]
  39. Ariyarathna, I.R.; Miliordos, E. Geometric and electronic structure analysis of calcium water complexes with one and two solvation shells. Phys. Chem. Chem. Phys. 2020, 22, 22426–22435. [Google Scholar] [CrossRef]
  40. Jackson, B.A.; Miliordos, E. The nature of supermolecular bonds: Investigating hydrocarbon linked beryllium solvated electron precursors. J. Chem. Phys. 2022, 156, 194302. [Google Scholar] [CrossRef]
  41. Werner, H.-J.; Knowles, P.J.; Knizia, G.; Manby, F.R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; Lindh, R.; et al. MOLPRO, Version 2015.1, a Package of ab Initio Programs. 2015. Available online: https://www.molpro.net/ (accessed on 5 June 2023).
  42. Celani, P.; Werner, H.-J. Multireference perturbation theory for large restricted and selected active space reference wave functions. J. Chem. Phys. 2000, 112, 5546–5557. [Google Scholar] [CrossRef]
  43. Roos, B.O.; Andersson, K. Multiconfigurational Perturbation-Theory with Level Shift—The Cr2 Potential Revisited. Chem. Phys. Lett. 1995, 245, 215–223. [Google Scholar] [CrossRef]
  44. Miliordos, E.; Ruedenberg, K.; Xantheas, S.S. Unusual Inorganic Biradicals: A Theoretical Analysis. Angew. Chem. Int. Ed. 2013, 52, 5736–5739. [Google Scholar] [CrossRef] [PubMed]
  45. Ariyarathna, I.R.; Miliordos, E. Ground and excited states analysis of alkali metal ethylenediamine and crown ether complexes. Phys. Chem. Chem. Phys. 2021, 23, 20298–20306. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Optimized geometries and contours for the singly occupied orbital of Th(NH3)n3+, n = 0–10.
Figure 1. Optimized geometries and contours for the singly occupied orbital of Th(NH3)n3+, n = 0–10.
Molecules 28 04712 g001
Figure 2. Dissociation energy (De) for ammonia ligands in the Th(NH3)n4+,3+,0 complexes corresponding to the Th(NH3)n4+,3+,0 → Th(NH3)n−14+,3+,0 + NH3 dissociation process.
Figure 2. Dissociation energy (De) for ammonia ligands in the Th(NH3)n4+,3+,0 complexes corresponding to the Th(NH3)n4+,3+,0 → Th(NH3)n−14+,3+,0 + NH3 dissociation process.
Molecules 28 04712 g002
Figure 3. Geometry of Th(NH3)84+ and CASSCF active orbitals for Th(NH3)83+; 1s, 1p, 1d correspond to peripheral outer orbitals, and 6d, 7f to inner thorium orbitals.
Figure 3. Geometry of Th(NH3)84+ and CASSCF active orbitals for Th(NH3)83+; 1s, 1p, 1d correspond to peripheral outer orbitals, and 6d, 7f to inner thorium orbitals.
Molecules 28 04712 g003
Figure 4. The addition of two ammonia ligands (marked with red nitrogen atoms) to the Th(NH3)83+,2+,1+,0 species occurs along the z-direction, resulting in the promotion of the 6dz2 electron to an outer 1s or 1p orbital.
Figure 4. The addition of two ammonia ligands (marked with red nitrogen atoms) to the Th(NH3)83+,2+,1+,0 species occurs along the z-direction, resulting in the promotion of the 6dz2 electron to an outer 1s or 1p orbital.
Molecules 28 04712 g004
Figure 5. Geometries of Th(12C4N)24+, Th(15C5N)24+, and CASSCF active orbitals for Th(12C4N)23+; 1s and 1p correspond to peripheral outer orbitals, and 6d, 7f to inner thorium orbitals.
Figure 5. Geometries of Th(12C4N)24+, Th(15C5N)24+, and CASSCF active orbitals for Th(12C4N)23+; 1s and 1p correspond to peripheral outer orbitals, and 6d, 7f to inner thorium orbitals.
Molecules 28 04712 g005
Table 1. Electronic configurations (EC) and excitation energies ΔΕ (eV) for the lowest lying electronic states of Th(NH3)83+, Th(12C4N)23+ and Th(12C4O)23+ at the CASSCF and CASPT2 levels of theory.
Table 1. Electronic configurations (EC) and excitation energies ΔΕ (eV) for the lowest lying electronic states of Th(NH3)83+, Th(12C4N)23+ and Th(12C4O)23+ at the CASSCF and CASPT2 levels of theory.
StateECCASSCFCASPT2CASSCFCASPT2CASSCFCASPT2
ΔΕ[Th(NH3)83+]ΔΕ[Th(12C4N)23+] aΔΕ[Th(12C4O)23+] a
X ˜ 2A1(6dz2)10.000.000.000.000.000.00
12E2(6dxy)1/(6dx2−y2)11.351.281.03/1.051.04/1.021.61/1.291.57/1.27
12E3(6dxz)1/(6dyz)11.921.79 4.36/4.004.28/3.88
12B2(7fz3)12.081.942.032.03N/C bN/C b
12E1(7fxz2)1/(7fyz2)12.182.012.13/2.112.11/2.11N/C bN/C b
22A1(1s)12.152.053.032.763.153.06
22B2(1pz)13.193.10 3.733.60
22E1(1px)1/(1py)13.243.153.50/3.363.23/3.103.42/3.713.36/3.64
22E2(7fxyz)1/(7fz(x2−y2))13.303.183.24/3.263.22/3.25N/C bN/C b
22E3(7fx(x2−3y2))1/(7fy(3x2−y2))13.593.511.89/1.831.88/1.84N/C bN/C b
32A1(1dz2)14.244.20 3.953.82
32E2(1dx2−y2)1/(1dxy)14.464.40 4.29/4.26
32E3(1dxz)1/(1dyz)14.894.86
a The degenerate states split into the 2A/2B components (C2 point group; see text). b The assignment was not clear, since the 7f orbitals are highly mixed. The CASSCF/CASPT2 energies are 1.94, 2.09, 2.35, 2.37, 2.53, 3.55, 3.72/1.93, 2.09, 2.31, 2.34, 2.49, 3.49, 3.67 eV.
Table 2. Electronic configurations (EC) and excitation energies ΔΕ (eV) for the lowest-lying electronic states of Th(NH3)81+ and Th(NH3)80 at the CASSCF and CASPT2 levels of theory.
Table 2. Electronic configurations (EC) and excitation energies ΔΕ (eV) for the lowest-lying electronic states of Th(NH3)81+ and Th(NH3)80 at the CASSCF and CASPT2 levels of theory.
StateElectron ConfigurationΔE (CASSCF)ΔE (CASPT2)
Th(NH3)81+
X ˜ 2A1a(6dz2)11s2/(6dz2)21s10.000.00
12E1 a(6dz2)11s11px,y1/1s21px,y1/(6dz2)21px,y10.470.50
14B2(6dz2)11s11pz10.480.51
14E1(6dz2)11s11px,y10.490.53
12B2a1s21pz1/(6dz2)21pz1/(6dz2)11s11pz10.530.56
Th(NH3)80
X ˜ 3E1(6dz2)11s21px,y10.000.00
X ˜ 1E1(6dz2)11s21px,y10.010.00
13B2(6dz2)11s21pz10.040.04
11B2(6dz2)11s21pz10.110.09
13A2(6dz2)11s11px11py10.280.30
15A2(6dz2)11s11px11py10.270.31
15E3(6dz2)11s11pz11px,y10.320.36
13E3(6dz2)11s11pz11px,y10.380.39
11E2a,b(6dz2)11s11p2/(6dz2)11s2(1dxy,x2−y2)10.560.54
13E2a,b(6dz2)11s11p2/(6dz2)11s2(1dxy,x2-y2)10.550.56
11E3a,b(6dz2)11s11p2/(6dz2)11s2(1dxz,yz)10.630.60
23E3a,b(6dz2)11s11p2/(6dz2)11s2(1dxz,yz)10.630.62
13A1a,b(6dz2)11s11p2/(6dz2)11s2(1dz2)10.630.63
11A1a,b(6dz2)11s11p2/(6dz2)11s2(1dz2)10.650.63
a See text for percentages of each configuration. b The 1p2 configuration is in a singlet spin multiplicity; see text for more details.
Table 3. CAM-B3LYP binding energies per Th-N bond for thorium ammonia, thorium aza-crown ethers, and thorium crown ethers for various charges q.
Table 3. CAM-B3LYP binding energies per Th-N bond for thorium ammonia, thorium aza-crown ethers, and thorium crown ethers for various charges q.
qTh(NH3)8qTh(12C4N)2qTh(12C4O)2qTh(NH3)10qTh(15C5N)2qTh(15C5O)2q
+4103.5108.9108.089.293.092.9
+363.163.764.153.255.054.9
+2N/A38.838.634.136.836.9
+1N/A19.318.719.115.414.2
013.68.66.112.27.15.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, Z.; Jackson, B.A.; Miliordos, E. Ab Initio Calculations on the Ground and Excited Electronic States of Thorium–Ammonia, Thorium–Aza-Crown, and Thorium–Crown Ether Complexes. Molecules 2023, 28, 4712. https://doi.org/10.3390/molecules28124712

AMA Style

Lu Z, Jackson BA, Miliordos E. Ab Initio Calculations on the Ground and Excited Electronic States of Thorium–Ammonia, Thorium–Aza-Crown, and Thorium–Crown Ether Complexes. Molecules. 2023; 28(12):4712. https://doi.org/10.3390/molecules28124712

Chicago/Turabian Style

Lu, Zhongyuan, Benjamin A. Jackson, and Evangelos Miliordos. 2023. "Ab Initio Calculations on the Ground and Excited Electronic States of Thorium–Ammonia, Thorium–Aza-Crown, and Thorium–Crown Ether Complexes" Molecules 28, no. 12: 4712. https://doi.org/10.3390/molecules28124712

Article Metrics

Back to TopTop