The van der Waals interactions in systems involving superheavy elements: the case of oganesson (Z = 118)

Luiz Guilherme Machado de Macedo a, Charles Alberto Brito Negrão b, Rhuiago Mendes de Oliveira c, Rafael Ferreira de Menezes d, Fernando Pirani ef and Ricardo Gargano *d
aFederal University of São João del Rei, Campus Centro Oeste Dona Lindu (CCO/UFSJ), 35501-296, Divinópolis, MG, Brazil
bPrograma de Pós-Graduação em Química (PPGQ), Universidade Federal do Pará, Belém, 66075-110, Brazil
cInstituto Federal de Educação, Ciência e Tecnologia do Maranhão, Campus Bacabal, MA 65700-000, Brazil
dInstituto de Física, Universidade de Brasília, Campus Darcy Ribeiro, Brasília, DF, Brazil. E-mail: gargano@unb.br
eDipartimento di Chimica, Biologia e Biotecnologie, Universitá degli studi di Perugia, via Elce di Sotto 8, Perugia, Italy
fDipartimento di Ingegneria Civile ed Ambientale, Università di Perugia, via Duranti 93, 06125 Perugia, Italy

Received 24th September 2022 , Accepted 28th November 2022

First published on 29th November 2022


Abstract

This work presents a study involving dimers composed of He, Ne, Ar, Kr, Xe, Rn, and Og noble gases with oganesson, a super-heavy closed–shell element (Z = 118). He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, Rn–Og, and Og–Og ground state electronic potential energy curves were calculated based on the 4-component (4c) Dirac–Coulomb Hamiltonian and were counterpoise corrected. For the 4c calculations, the electron correlation was taken into account using the same methodology (MP2-srLDA) and basis set quality (Dyall's acv3z and Dunning's aug-cc-PVTZ). All calculations included quantum electrodynamics effects at the Gaunt interaction level. For all the aforementioned dimers the vibration energies, spectroscopic constants (ωe, ωexe, ωeye, αe, and γe), and lifetime as a function of the temperature (which ranged from 200 to 500 K) were also calculated. The obtained results suggest that the inclusion of quantum electrodynamics effects reduces the value of the dissociation energy of all hetero-nuclear molecules with a percentage contribution ranging from 0.48% (for the He–Og dimer) to 9.63% (for the Rn–Og dimer). The lifetime calculations indicate that the Og–He dimer is close to the edge of instability and that Ng–Og dimers are relatively less stable when the Gaunt correction is considered. Exploiting scaling laws that adopt the polarizability of involved partners as scaling factors, it has also been demonstrated that in such systems the interaction is of van der Waals nature (size repulsion plus dispersion attraction) and this permitted an estimation of dissociation energy and equilibrium distance in the Og–Og dimer. This further information has been exploited to evaluate the rovibrational levels in this symmetric dimer and to cast light on the macroscopic properties of condensed phases concerning the complete noble gas family, emphasizing some anomalies of Og.


1 Introduction

Research involving relativistic quantum chemistry of super-heavy atoms is attracting a lot of attention from the scientific community with the emergence of new elements in the periodic table. For instance, Eliav et al.1 calculated the electron affinity of oganesson (Og), the most recently discovered noble gas, to be Z = 118,2 by using the relativistic coupled-cluster method based on the Dirac–Coulomb–Breit Hamiltonian. The obtained value was 0.056 eV (with an estimated error of 0.01 eV) when relativistic corrections were included. However, no Og electron affinity was found when non-relativistic or uncorrelated calculations were performed.1 This fact shows that relativistic theoretical chemistry plays an extremely relevant role, often being the exclusive chemical information source useful in describing the properties of systems involving super-heavy elements.3

Since few isotopes of the transactinide elements can be obtained, what is known about their properties comes mostly from theory. A theoretical study becomes very important, especially when the chemical/physical properties of the heaviest elements usually cannot be predicted from extrapolations of known characteristics in the groups or periods of the Periodic Table.3 In fact, the higher the nuclear charge of an atom, the faster the speed of the innermost electrons, which reach speeds corresponding to a considerable fraction of the light speed, and substantially affect the atomic behaviour. Thus, the relativistic mass of the electrons increases and causes the contraction and stabilization of the orbitals. This fact allows other outermost orbitals to expand and destabilize. Therefore, theoretical calculations for super-heavy elements need to include the special relativity theory for a correct description of the atomic properties of the heaviest elements.4

In the systems involving super-heavy atoms, the effects of quantum electrodynamics (QED) must be included in the rationalization of their behavior. For instance, the ionization potential of the lawrencium atom has recently been determined and the obtained result agreed with the experimental value (4.96 eV) only when QED effects were included. On the other hand, the understanding of the nature of the main components determining non-covalent interactions is very important since it allows us to rationalize and predict a large number of chemical/physical phenomena.5–9 In the literature, however, there are a limited number of experimental/theoretical studies available that focus on characterization of the weak interaction, vibrational energies, spectroscopy, and lifetime of the dimers involving Ng and super-heavy atoms.

Based on these facts, the present paper reports on the characterization of the interaction in Ng–Og (Ng = He, Ne, Ar, Kr, Xe, Rn, and Og) systems considering relativistic effects at the 4-component level and with the inclusion of QED effects at the Gaunt interaction level and basis set superposition error (BSSE).10 More precisely, our work aimed to obtain the Ng–Og ground state electronic potential energy curves (PECs) by applying a combination of the second-order Moller–Plesset long-range theory (MP2-srDFT) with the DFT short-range approach within the framework of the 4-component relativistic method. These PECs have then been used to calculate the vibrational energies, spectroscopic constants, and lifetime of the Ng–Og dimers.

A further effort, addressed to better define the nature, strength and range of the leading interaction components, determining binding energy and equilibrium distance in systems involving Og, has also been performed fully exploiting all the obtained theoretical results. In particular, it has been demonstrated that the obtained interaction potentials are of van der Waals type, depending exclusively on the polarizability value of the involved partners,11 which is a collective property of all electrons (mostly the external ones and partially the internal ones) defining both size repulsion and dispersion attraction. For the Og atom, a polarizability value consistent with that recently proposed by Jerabek et al.12 has been adopted. The use of correlation formulas,11 providing the potential well features of van der Waals interactions in terms of polarizability, also allowed the prediction of binding features and of the strength of the long-range attraction in the Og2 dimer. Exploiting this information, the dependence of melting (Tmp) and boiling (Tbp) temperature on the two-body interaction components has been analyzed for the complete family of noble gas atoms. With respect to the general trend of all other noble gas atoms, some deviations of Og have been appropriately emphasized. In particular, an increased role of multi-body interaction contributions extends in an anomalous way the stability of the oganesson liquid phase, probably due to the highest relativistic effects, as recently proposed by Smits et al.14

2 Methodologies

The determination for any system of energies and rovibrational spectroscopic constants, involves an accurate fitting of radial dependence of the interaction energy, indicated as the potential energy curve (PEC). This fitting is necessary to characterize the dissociation energy, De, the equilibrium distance, Re, and the shape of the potential well. For these purposes, two analytical forms were used. The first option to fit the PEC of the studied systems was the adoption Improved Lennard Jones (ILJ) function. This model, defined in terms of only three parameters, has been extensively used in several molecular systems and it has proved to be very efficient to describe van der Waals-type interactions.15–19 For neutral-neutral systems, the ILJ form assumes the following expression:20–22
 
image file: d2cp04456k-t1.tif(1)
where image file: d2cp04456k-t2.tif and the β parameter describes the softness/hardness of the elements involved in the molecular system. It is of relevance to stress that the shape of the potential well, obtained from the ILJ formulation, is of great significance for the present study and is in agreement with that provided by other important potential models.23–25 A detailed probe of the ILJ formulation, applied to the noble gas interactions and performed on high-resolution scattering experiments, has been discussed by Pirani et al.20 In particular, the analysis of the velocity dependence of the integral cross section represented a critical test of the radial dependence of the long-range attraction strength, arising from the critical balance of appropriately damped long-range dispersion coefficients. At the same time, the observed interference effects have been exploited to control basic features of the potential well. Moreover, for a He–He molecule a direct comparison of the ILJ potential formulation with results of ab initio calculations and of other more sophisticated potential models26 has been discussed in ref. 15 and 27.

The second analytical form was the extended-Rydberg28 function, which has been used successfully to describe various molecular systems29–31 and it is described according to the equation:

 
image file: d2cp04456k-t3.tif(2)
where ci are adjustable coefficients. Once the analytical forms that fit the electronic energies are known, one can solve the nuclear Schrödinger equation (within the Born–Oppenheimer approximation) and determine the vibrational energies of the molecules under study. In the present study, this equation was solved employing the Discrete Variable Representation method.32 Once the vibrational energies are known, the rovibrational spectroscopic constants, such as ωe, ωexe, ωeye, αe, and γe, can be calculated using the following expressions:29
 
image file: d2cp04456k-t4.tif(3)
where Ev,j represents the rovibrational energy of a level (ν, j), and ν and j represent the vibrational and rotational quantum numbers, respectively.

As far as we know, there is no experimental data for the systems studied here. Thus, it is necessary to determine the spectroscopic constants using another methodology to ensure the reliability of the current results. The Dunham method,33 which depends on the derivatives of the potential energy curves in the equilibrium position, was chosen as a second option to calculate the rovibrational spectroscopic constants.

To determine the lifetime as a function of temperature for each Ng–Og dimer within Slater theory,34,35 the following equation was employed:

 
image file: d2cp04456k-t5.tif(4)
where Rg represents the universal gas constant, T is the temperature, E0,0 is the ground state rovibrational energy, ωe is the vibrational harmonic spectroscopic constant, and De is the dissociation energy.

3 Computational details

All calculations were performed using the DIrac1736 program package, were based on 4-component (4c) Dirac-Coulomb Hamiltonian37,38 and were counterpoise corrected.10 In the 4c calculations, the (SS|SS) two electron integrals were neglected and replaced by a charge39 using Dirac's default settings. Since the Dirac program can include the Gaunt interaction only at the Dirac–Hartree–Fock level, it was treated additively.40,41

For the 4c calculations, the electron correlation was taken into account using the long-range Moller–Plesset perturbation Theory MP2-lrDFT42 and its short-range srLDA version43,44 since it has a better performance than MP2-srPBE for rare gas dimers.42 For all systems, at least 60 different distances were calculated from the repulsion region (from around 2.5–3.4 Å) up to the dissociation limit. The electrons correlated were the valence (n − 1)dnsnp and the energy threshold for inclusion of virtual orbitals for MP2-srLDA calculations was set to +20 hartrees due to computational limitations. For He, Ne and Ar atoms we employed the aug-CC-PVTZ basis sets of Dunning and coworkers.45–48 For heavier atoms (Kr, Xe, Rn and Og), we employed the acv3z basis set of Dyall.49,50

Finally, the obtained electronic energies were fitted with both the ILJ model and extended-Rydberg analytical form (with ten coefficients) via Powell's method.51 In the fitting process, the equilibrium distance and the dissociation energy were kept fixed. With the potential energy curves in hand, calculations of rovibrational energies, spectroscopic constants and the lifetime were carried out for each system.

4 Results and discussion

We found it appropriate to present the obtained results with their discussion in four subsections in order to better emphasize their relevance and to exhibit their general implications.

4.1 Dissociation energy and equilibrium distance

Tables S1–S7 of the ESI show the calculated ground state electronic energies of the He–Og, Ar–Og, Kr–Og, Xe–Og, Rn–Og, and Og–Og dimers for several different internuclear distances R, considering the inclusion of quantum electrodynamics effects at the Gaunt interaction level (QEEGIL), without inclusion of QEEGIL, and with inclusion of QEEGIL + BSSE correction. Table 1 presents the dissociation energies (De) and equilibrium distances (Re) for each diatomic system. From this table, it is observed that the inclusion of quantum electrodynamics effects reduces the value of the dissociation energy of these molecules, which may be attributed to retardation since it could be significant in dispersion interactions.52 The percentage contribution of the quantum electrodynamics effect on the dissociation energy ranged from 0.56% (for the He–Og dimer) to 9.63% (for the Rn–Og dimer), that is, as the atomic number of the noble gases increases, the more significant is the contribution of the Gaunt correction. When the BSSE correction is considered, there is a decrease of the dissociation energy of the dimers. The greatest decrease (1.53 meV) occurred for the Og–Ne system. On the other hand, the equilibrium distance remains practically unchanged when the BSSE correction is taken into account. The De results found for the Og–Og system were compared with those available in the literature. The differences between the values determined by Gaunt + BSSE and predicted by correlation formulas with those obtained by Saue et al. were 0.0272 eV (0.6270 kcal mol−1) and 0.0236 eV (0.5447 kcal mol−1), respectively. When a comparison is made taking into account the results of Jerabek et al., these values are 0.02685 eV (0.6191 kcal mol−1) and 0.02394 eV (0.5526 kcal mol−1), respectively. It can also be inferred from Table 1 that De increases as the noble gas atomic mass also increases, but the same trend does not happen with the Re. The He–Og equilibrium distance is comparable (or even larger) with respect to that of the largest among all the systems. This feature can be explained by exploring the fact that the Re of a van der Waals complex is determined through the balance between the terms of repulsion (exchange) and of attraction (dominated by dispersion effects). Both these terms depend on the electronic polarizability of the interacting partners as a measure of their size and provides also the probability of induced dipole formation. Knowing that Og shows the larger size, being more polarizable than He, Ne, Ar, Kr, Xe, and Rn atoms, then it is expected that the size-repulsion (term of exchange) is practically the same for the He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, and Rn–Og molecule and this determines equilibrium distances of comparable value.
Table 1 He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, Rn–Og, and Og–Og dissociation energies (De) and internuclear equilibrium distances (Re) calculated without inclusion of quantum electrodynamics effects at the Gaunt interaction level (QEEGIL), with inclusion of QEEGIL, and with inclusion of QEEGIL + BSSE correction. The De, Re, and asymptotic dispersion coefficient (C6) predicted by correlation formulas (see Cambi et al.11) were determined using for Og the polarizability value of 8.0 Å3, a value consistent with that reported by Jerabek et al.12 and an effective number of electrons equal to 24 (see Cambi et al.11). The estimated uncertainty in the C6 is about 10–15% for lighter and 15–20% for heavier systems
Systems D e (meV) R e (Å) C6 (eV Å6) Source
He–Og 1.77 4.62 (Without Gaunt)
He–Og 1.76 4.62 (With Gaunt)
He–Og 1.61 4.55 (Gaunt + BSSE)
He–Og 2.91 4.37 20.2 (Predicted by correlation formulas)
Ne–Og 6.38 4.20 (Without Gaunt)
Ne–Og 6.39 4.20 (With Gaunt)
Ne–Og 4.86 4.30 (Gaunt + BSSE)
Ne–Og 6.73 4.33 44.3 (Predicted by correlation formulas)
Ar–Og 19.40 4.33 (Without Gaunt)
Ar–Og 19.25 4.33 (With Gaunt)
Ar–Og 19.00 4.30 (Gaunt + BSSE)
Ar–Og 20.13 4.40 145.9 (Predicted by correlation formulas)
Kr–Og 26.17 4.39 (Without Gaunt)
Kr–Og 25.57 4.40 (With Gaunt)
Kr–Og 24.98 4.40 (Gaunt + BSSE)
Kr–Og 27.34 4.46 215.8 (Predicted by correlation formulas)
Xe–Og 34.94 4.48 (Without Gaunt)
Xe–Og 33.44 4.49 (With Gaunt)
Xe–Og 32.76 4.50 (Gaunt + BSSE)
Xe–Og 35.53 4.56 320.7 (Predicted by correlation formulas)
Rn–Og 55.14 4.40 (Without Gaunt)
Rn–Og 51.34 4.40 (With Gaunt)
Rn–Og 50.50 4.40 (Gaunt + BSSE)
Rn–Og 43.44 4.63 426.7 (Predicted by correlation formulas)
Og–Og 104.58 4.25 (Gaunt + BSSE)
Og–Og 53.79 4.76 626.5 (Predicted by correlation formulas)
Og–Og 77.73 4.31 Jerabek et al.12
Og–Og 77.39 4.33 Saue et al.13


Indeed, all Re values are confined within 10% of their average value, and any small reduction, such as that observed when passing from He–Og to Ne–Og, and the general small increase along the Ng–Og family is due to the fine balance of size repulsion with dispersion attraction. On the other hand, the De value critically depends on the strength of the global attraction at Re which is related to the polarizability product of involved partners. Since the polarizability of the He atom is lower than that of the other noble gases, the associated De is the smallest one and its value increases regularly along the family by a factor larger than 20. Therefore, the observed behavior is typical of systems whose repulsion component, controlling Re, is mostly dominated by the size of the most polarizable atom (Og), while De varies along a homologous family (Ng–Og) according to the polarizability change of the Ng partner.

Fig. S1 and S2 (ESI) show the ab initio (with and without Gaunt interaction) and ILJ adjusted PEC of the He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, and Rn–Og molecules, while Fig. S3 and S4 (ESI) presents the ab initio (with and without Gaunt interaction) and Rydberg adjusted PEC of the same systems. Through these figures one notes that the Og–He electronic energies have small oscillations. Ab initio (with Gaunt + BSSE correction) and ILJ adjusted PEC of the He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, and Rn–Og dimers are shown in Fig. S5 (ESI), while the ab initio (with Gaunt + BSSE correction) and Rydberg adjusted PEC of the same complexes are represented in Fig. S6 (ESI). From these figures it is possible to verify that, even when the BSSE correction is taken into account, the small oscillations in the electronic energies of the Og–He system continue.

Tables 2 and 3 present the fitted values of the β ILJ parameter and the extended-Rydberg analytic coefficients, respectively. The β fitted values, depending on the softness of the interaction partners, ranged between 5.80 and 9.58. For this range, the ILJ analytical form works very well for various complexes formed by neutral–neutral and ion–neutral species, i.e., they provided an adequate description of the role of the van der Waals interaction component in the formation of weak hydrogen and halogen intermolecular bonds.53,54Tables 2 and 3 also show that the RMSD are very small, evidencing the good quality of the He–Og, Ar–Og, Kr–Og, Xe–Og, Rn–Og, and Og–Og electronic energy fits with Gaunt + BSSE correction and the Ne–Og electronic fits with only Gaunt interaction.

Table 2 Fitting values of the β ILJ parameter and root mean square deviation (RMSD)
Systems RMSD (Hartree) β parameter
He–Og 2.0976 × 10−6 9.5808
Ne–Og 2.7833 × 10−6 7.2151
Ar–Og 1.0030 × 10−5 8.0185
Kr–Og 1.2681 × 10−5 6.6153
Xe–Og 1.2135 × 10−5 6.9770
Rn–Og 6.3491 × 10−6 6.8909
Og–Og 1.0493 × 10−4 5.8012


Table 3 Adjustable parameters (cnn)) of the extended-Rydberg analytic form of degree 10 and root mean square deviation (RMSD)
c n n) He–Og Ne–Og Ar–Og Kr–Og Xe–Og Rn–Og Og–Og
C 1 1.47710955 1.38672315 1.79983909 3.39536153 2.77583010 1.70196661 1.75418289
C 2 −1.12088589 −0.81105960 −0.45115730 3.94296832 2.15135216 −0.24350070 −0.00576175
C 3 1.18081113 0.23626192 0.47373096 2.03562284 0.75799095 0.23677149 0.21663296
C 4 0.51099513 −0.25504175 0.23110898 0.77943221 0.19755360 0.01311309 0.05863362
C 5 −1.30890474 0.50599162 −0.35933640 2.41706278 0.97262535 0.00125084 0.06468921
C 6 0.13378930 −0.30124010 0.12987214 1.02490653 0.06621025 −0.02737260 −0.03353552
C 7 0.62722811 0.03474181 0.00452929 −1.44895344 −0.68265120 0.02038624 0.00327290
C 8 −0.38639035 0.02387602 −0.01121581 0.28371066 0.41795297 −0.00607865 0.00397981
C 9 0.08922130 −0.00815757 0.00216509 0.18217579 −0.08852640 0.00084895 −0.00112308
C 10 −0.00739282 0.00074162 −0.00012930 −0.04479104 0.00542335 −0.00004265 0.00010223
RMSD 1.5224 × 10−6 1.1589 × 10−6 8.5612 × 10−6 6.3886 × 10−6 1.2681 × 10−5 6.3491 × 10−6 1.0493 × 10−4


Fig. S7 (ESI) shows the Og–Og ab initio electronic energies (with Gaunt + BSSE correction) and those adjusted by both the ILJ and Rydberg forms. It can be seen from this figure that the ILJ analytical form does not adjust well the ab initio energies at the anharmonic region of the PEC. This fact suggests that the results of rovibrational energies, spectroscopic constants and the lifetime of the Og–Og molecule are more reliable when the Rydberg analytical function is used.

4.2 Nature of the interatomic interaction

As anticipated in Section 4.1, for all systems under investigation the potential energies are determined by a critical balance of size repulsion with dispersion attraction. This means that they exhibit the nature of typical van der Waals interactions. Further confirmation of this general aspect comes from the prediction of correlation formulas, representing both Re and De for all systems in terms of polarizability (α) of the involved partners.11 It is interesting to note that all correlation formulas here exploited, defined in terms of the electronic polarizability α of the interacting partners and first proposed 37 years ago,55 have been suggested on the phenomenological ground.11 Moreover, the one defining Re, showing a dependence on α1/7 values, has been recently confirmed by a refined quantum mechanical treatment.56 Therefore, the two approaches provide coincident-internally consistent results for symmetric and asymmetric noble gas dimers, respectively. Predicted values of the basic De and Re quantities are given in Table 1, where they are compared with the results of the present ab initio calculations. Considering the good agreement between the calculated and predicted values, the same correlation formulas have been exploited to estimate Re and De also for the Og–Og dimer. For all systems, the same correlation formulas are given and also the C6 coefficient that provides the leading term of the dispersion attraction due to induced dipole–induced dipole interaction.

Considering the high number of internal and external electrons, with associated relativistic effects, the fundamental properties of the heavy Og monomer are not easy to evaluate. Therefore, for the Og–Og dimer, the De value predicted by the correlation formulas and that obtained by the present calculations (see Table 1) can be reasonably taken as the lower and the upper limit of the dissociation energy. In particular, considering the mutual-combined uncertainty, they are in reasonable agreement, within ≈0.5 kcal mol−1 (1 kcal mol−1 = 43.32 meV) with theoretical estimates given in ref. 12 and 13 (only for De predicted by correlation formulas does the uncertainty amount to 20–30%). Note also that the long-range C6 attraction coefficients, evaluated for Og–Og inserting De and Re of Table 1 in the asymptotic behavior of ILJ, and defined as C6 = DeRe6, agree within 20% with the results of correlation formulas.

4.3 Some speculations on the macroscopic properties

We also researched a possible correlation between the macroscopic properties of noble gas atoms in their condensed phases and the intermolecular forces involved. In particular, the focus has been on the temperatures Tmp and Tbp of melting and boiling points of prototype van der Waals systems provided by noble gases, whose solid phase is usually classified as typical of close (compact)-packing. It is well known that for such systems the repulsive components are mainly responsible for melting while attractive components are responsible for boiling. For a family of systems, where many-body components directly relate to two-body contributions, it can be expected that the balance of their attraction and repulsion components affects the cohesive energy. Eqn (1) and (2) suggest that at R = Re both components scale approximately with De. Therefore, we have attempted to find a possible correlation between Tmp and Tbp with De. All involved values are given in Table 4, while Fig. 1 shows the obtained plot. For all systems from He to Rn, the reported Tmp and Tbp are experimental values while for Og they have been recently proposed by Smits et al.14 The first important feature to be emphasized is that for all systems, including the super-heavy Og, Tmb scales almost linearly with De. The second important feature concerns Tbp which follows an approximately linear dependence on De for all systems, except the super-heavy Og. In particular, for most systems, the existence zone of the liquid phase is confined in the range of a few K, while Og exhibits a range of liquid larger than 100 K. This anomalous behavior probably arises from the largest many-body contributions which are triggered by relativistic effects.
Table 4 Correlation between dissociation energy De, due to two bodies interaction provided by different sources, with temperatures of the melting point Tmp and the boiling point Tbp. Estimated uncertainty in De values starts from the fraction meV for Ne–Ne up to 1 meV for Xe–Xe (in all these systems accurate experimental information is available) and increases to 3–4 meV for Rn–Rn and to 5–6 meV for Og–Og. Note that except for Og all other noble gas Tmp and Tbp are taken from ref. 57
System T mp (K) T bp (K) D e (meV) Source
Ne–Ne 24.5 27.1 4.28 Predicted by correlation formulas
3.66 Exp.15
3.75 Calculations15
Ar–Ar 83.8 87.5 11.61 Predicted by correlation formulas
12.37 Exp.15
12.36 Calculations15
Kr–Kr 116.6 120.9 17.61 Predicted by correlation formulas
17.30 Exp.15
17.05 Calculations15
Xe–Xe 161.3 166.1 25.36 Predicted by correlation formulas
24.20 Exp.15
24.31 Calculations15
Rn–Rn 202.2 211.4 36.20 Predicted by correlation formulas
30.32 Calculations15
Og–Og 335 ± 30 350 ± 30 53.79 Predictions-two body extrapolation 14
325 ± 15 450 ± 10



image file: d2cp04456k-f1.tif
Fig. 1 Dependence, for the complete family of Ng, of Tmp and Tbp temperatures on the dissociation energy De generated by the two-body interaction. The horizontal error bars represent the uncertainty in the De value, which is larger for Rn and Og because of the absence of experimental information, which is instead available for all other noble gases.

4.4 Rovibrational levels of dimers

Vibrational energy values (not shown in this work) obtained for both Og–He ILJ and Rydberg PEC with Gaunt and without Gaunt interaction are very similar, suggesting that gaunt correction is not important for this particular system. The PEC's of both Ne–Og and Ar–Og with and without Gaunt corrections have very close vibrational energy values (a total of 6 and 16, respectively) indicating that the Gaunt correction is also not very important for these dimers. It was also found that the Kr–Og ILJ PEC with and without Gaunt correction have the same number of vibrational levels (a total of twenty-six levels). Moreover, the vibrational energies calculated with Gaunt correction are smaller than those calculated without Gaunt correction. The Xe–Og dimer presents thirty-seven and thirty-six vibrational levels with and without Gaunt corrections, respectively, while the Rn–Og system contains fifty-five and fifty-two levels. Also, for both systems, the values of Gaunt vibrational energies are smaller than without Gaunt correction. These features clearly show that the Gaunt correction becomes more important as the atomic number of atoms increases.

Tables 5 and 6 present the He–Og, Ar–Og, Kr–Og, Xe–Og, and Rn–Og pure vibrational (j = 0) and rovibrational (j = 1) energies for the ILJ PEC with Gaunt + BSSE correction. For the Og–Ne system, both the used ILJ and Rydberg PEC were obtained only with the Gaunt interaction for the reasons explained in Section 4.1. From these tables, it can be seen that the He–Og system has only a single vibrational level considering both the ILJ and Rydberg PEC with Gaunt + BSSE correction (Tables S8 and S9 of the ESI). Vibrational energies are very sensitive to PEC adjustments. In fact, small differences in adjustments can produce changes in the amount of vibrational levels within the well of each adjusted PEC. This fact was observed for some systems involved in the present study. For example, for the Og–Ar, Og–Kr, Og–Xe, and Og–Rn systems, 16, 26, 36, and 52 vibrational levels were obtained for the PEC ILJ, respectively, while for the PEC Rydberg these values were 14, 20, 27, and 50. Og–Og vibrational energies (80 levels) obtained via ILJ PEC adjusted from Gaunt + BSSE electronic energies are shown in Table 7, while the Rydberg PEC contains 107 vibrational levels (Table S10 of the ESI). With the values of Re and De estimated (correlation formulas) for the Og–Og dimer, it was also possible to construct the ILJ PEC of this system.

Table 5 He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og and Rn–Og pure vibrational energies calculated through ILJ potential energy curves
ν j He–Og Ne–Og Ar–Og Kr–Og Xe–Og Rn–Og
0 8.087589 7.557380 11.719211 9.145629 8.856483 9.512609
1 20.082742 33.622236 26.770172 26.089224 28.181499
2 29.220193 53.505590 43.507955 42.684451 46.377093
3 35.116381 71.388952 59.358726 58.642633 64.099541
4 38.161975 87.296729 74.322389 73.964313 81.349010
5 39.171675 101.260169 88.399103 88.650133 98.125690
6 113.320322 101.589424 102.700861 114.429800
7 123.532024 113.894481 116.117435 130.261593
8 131.968954 125.316218 128.901003 145.621367
9 138.729297 135.857706 141.052978 160.509470
10 143.940704 145.523552 152.575108 174.926314
11 147.762342 154.320420 163.469549 188.872385
12 150.382299 162.257693 173.738967 202.348253
13 152.011351 169.348290 183.386644 215.354596
14 152.876593 175.609621 192.416619 227.892208
15 153.227384 181.064657 200.833844 239.962023
16 185.742987 208.644363 251.565141
17 189.681710 215.855529 262.702847
18 192.925912 222.476230 273.376646
19 195.528566 228.517142 283.588290
20 197.549798 233.990987 293.339822
21 199.055775 238.912788 302.633612
22 0 200.117520 243.300080 311.472406
23 200.809866 247.173075 319.859375
24 201.210565 250.554727 327.798170
25 201.411283 253.470702 335.292986
26 255.949227 342.348622
27 258.020855 348.970548
28 259.718178 355.164974
29 261.075536 360.938917
30 262.128754 366.300264
31 262.914910 371.257832
32 263.472121 375.821406
33 263.839330 380.001781
34 264.056097 383.810769
35 264.169466 387.261196
36 390.366872
37 393.142551
38 395.603860
39 397.767234
40 399.649835
41 401.269475
42 402.644549
43 403.793971
44 404.737111
45 405.493744
46 406.083994
47 406.528272
48 406.847220
49 407.061634
50 407.192395
51 407.263120


Table 6 He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og and Rn–Og Rovibrational energies calculated through ILJ potential energy curves with the rotational quantum number j = 1
ν j He–Og Ne–Og Ar–Og Kr–Og Xe–Og Rn–Og
0 8.405667 7.649373 11.770146 9.172055 8.874696 9.526304
1 20.164716 33.671288 26.796031 26.107153 28.195047
2 29.290414 53.552666 43.533230 42.702090 46.390492
3 35.172352 71.433944 59.383401 58.659977 64.112788
4 8.200502 87.339514 74.346444 73.981355 81.362104
5 39.191075 101.300607 88.422519 88.666865 98.138628
6 113.358254 101.612177 102.717277 114.442579
7 123.567271 113.916547 116.133527 130.274212
8 132.001316 125.337570 128.916762 145.633822
9 138.758560 135.878315 141.068397 160.521759
10 143.966642 145.543384 152.590176 174.938435
11 147.784726 154.339439 163.484257 188.884333
12 150.400893 162.275861 173.753304 202.360027
13 152.025912 169.365564 183.400598 215.366192
14 152.886857 175.625958 192.430179 227.903622
15 153.233754 181.080007 200.846995 239.973253
16 185.757302 208.657093 251.576182
17 189.694938 215.867824 262.713695
18 192.938003 222.488073 273.387297
19 195.539469 228.528517 283.598740
20 197.559463 234.001879 293.350066
21 199.064152 238.923178 302.643646
22 1 200.124558 243.309952 311.482225
23 200.815512 247.182410 319.868974
24 201.214765 250.563508 327.807544
25 201.414366 253.478910 335.302129
26 255.956845 342.357528
27 258.027866 348.979211
28 259.724563 355.173389
29 261.081278 360.947077
30 262.133834 366.308164
31 262.919310 371.265464
32 263.475821 375.828765
33 263.842309 380.008860
34 264.058337 383.817561
35 264.171150 387.267694
36 390.373070
37 393.148442
38 395.609439
39 397.772493
40 399.654768
41 401.274076
42 402.648811
43 403.797887
44 404.740675
45 405.496948
46 406.086831
47 406.530736
48 406.849302
49 407.063326
50 407.193690
51 407.264090


Table 7 Og–Og Rovibrational energies calculated through ILJ potential energy curves
ν j Og–Og ν j Og–Og ν j Og–Og ν j Og–Og
0 12.501425 40 702.818996 0 12.514080 40 702.827310
1 37.221281 41 712.188881 1 37.233850 41 712.197050
2 61.563577 42 721.182957 2 61.576058 42 721.190979
3 85.527891 43 729.803055 3 85.540284 43 729.810928
4 109.113793 44 738.050734 4 109.126098 44 738.058455
5 132.320846 45 745.927666 5 132.333061 45 745.935235
6 155.148607 46 753.436238 6 155.160732 46 753.443653
7 177.596626 47 760.580076 7 177.608659 47 760.587334
8 199.664447 48 767.364268 8 199.676388 48 767.371365
9 221.351612 49 773.795154 9 221.363461 49 773.802085
10 242.657659 50 779.879770 10 242.669413 50 779.886529
11 263.582122 51 785.625209 11 263.593782 51 785.631790
12 284.124537 52 791.038169 12 284.136101 52 791.044568
13 304.284438 53 796.124843 13 304.295905 53 796.131057
14 324.061362 54 800.891129 14 324.072732 54 800.897155
15 0 343.454852 55 0 805.342998 15 1 343.466122 55 1 805.348837
16 362.464454 56 809.486880 16 362.475624 56 809.492530
17 381.089724 57 813.329935 17 381.100793 57 813.335396
18 399.330227 58 816.880194 18 399.341194 58 816.885466
19 417.185546 59 820.146578 19 417.196409 59 820.151660
20 434.655274 60 823.138828 20 434.666032 60 823.143718
21 451.739031 61 825.867378 21 451.749683 61 825.872074
22 468.436455 62 828.343216 22 468.447000 62 828.347716
23 484.747218 63 830.577753 23 484.757653 63 830.582053
24 500.671020 64 832.582698 24 500.681345 64 832.586794
25 516.207603 65 834.369970 25 516.217816 65 834.373860
26 531.356753 66 835.951643 26 531.366852 66 835.955321
27 546.118305 67 837.339899 27 546.128290 67 837.343362
28 560.492154 68 838.547020 28 560.502022 68 838.550263
29 574.478257 69 839.585378 29 574.488006 69 839.588399
30 588.076645 70 840.467439 30 588.086274 70 840.470234
31 601.287437 71 841.205765 31 601.296943 71 841.208331
32 614.110862 72 841.813012 32 614.120244 72 841.815344
33 626.547268 73 842.301919 33 626.556525 73 842.304015
34 638.597113 74 842.685305 34 638.606241 74 842.687162
35 650.260928 75 842.976041 35 650.269926 75 842.977655
36 661.539322 76 843.187028 36 661.548188 76 843.188397
37 672.433049 77 843.331171 37 672.441781 77 843.332291
38 682.943163 78 843.421348 38 682.951758 78 843.422215
39 693.071161 79 843.472547 39 693.079617 79 843.473215


Table 8 shows the calculated values for the rovibrational spectroscopic constants of all systems under study. An important fact that can be observed from this table is the good agreement of the results obtained with both methods, i.e., eqn (3) (named with the DVR method) and Dunham's method. With the exception of the He–Og dimer, there is a good agreement between the values of the rovibrational spectroscopic constants obtained with the ILJ and extended-Rydberg PEC. It can be seen that it was not possible to determine the spectroscopic constants for the He–Og system (for both ILJ and Rydberg PECs) viaeqn (3). The reason lies in the fact that this equation can only be used if there are at least 4 vibrational levels within the well of the potential energy curve. Table 8 shows also the Og–Og spectroscopic constant results available in the literature. When the results of ωe obtained with Dunham-Rydberg and correlation formulas are compared with Jerabek's results, the differences are 2.05 cm−1 and 5.45 cm−1, respectively. When we make the same comparisons for the ωexe vibrational constant, we have the following differences 0.001 cm−1 and 0.21 cm−1, respectively. When the ωe harmonic spectroscopic constants calculated with Dunham-Rydberg and Correlation formulas are compared with the value obtained by Saue, we have the following differences 2.70 cm−1 and 4.80 cm−1, respectively. From these comparisons, it can be noted that the values found in the present work for the Og–Og spectroscopic constants (mainly those obtained with Dunham-Rydberg) agree well with the results of Jerabek and Saue.

Table 8 Rovibrational spectroscopic constants for the Ng–Og (Ng = He, Ne, Ar, Kr, Xe, Rn, and Og) systems using ILJ and Rydberg potential energy curves
System Methods ω e (cm−1) ω e x e (cm−1) ω e y e (cm−1) α e (cm−1) γ e (cm−1)
He–Og DVR-ILJ
Dunham-ILJ 20.88670
DVR-Rydberg
Dunham-Rydberg 22.14117
DVR-ILJ 15.96782 1.73275 0.00403 4.14 × 10−3 −4.33 × 10−4
Ne–Og Dunham-ILJ 16.05381 1.80394 0.02444 4.40 × 10−3 −2.45 × 10−4
DVR-Rydberg 16.46278 1.93427 0.04158 3.65 × 10−3 −6.00 × 10−4
Dunham-Rydberg 15.75879 1.28925 −0.12236 3.93 × 10−3 −3.12 × 10−4
DVR-ILJ 23.94156 1.02459 0.00328 8.94 × 10−4 −2.35 × 10−5
Ar–Og Dunham-ILJ 23.93896 1.02207 0.00236 8.99 × 10−4 −1.94 × 10−5
DVR-Rydberg 24.71511 1.29686 0.02626 9.23 × 10−4 −2.55 × 10−5
Dunham-Rydberg 24.66753 3.39022 −1.61570 9.96 × 10−4 9.27 × 10−4
DVR-ILJ 18.51106 0.44319 −0.00004 2.76 × 10−4 −3.94 × 10−6
Kr–Og Dunham-ILJ 18.51139 0.44315 −0.00006 2.76 × 10−4 −3.47 × 10−6
DVR-Rydberg 19.50191 0.41346 −0.01355 1.72 × 10−4 −2.14 × 10−5
Dunham-Rydberg 19.47647 0.47659 −0.09941 1.70 × 10−4 3.35 × 10−5
DVR-ILJ 17.87071 0.31911 0.00008 1.39 × 10−4 −1.40 × 10−6
Xe–Og Dunham-ILJ 17.87086 0.31906 0.00006 1.39 × 10−4 −1.29 × 10−6
DVR-Rydberg 18.28701 0.27636 −0.00593 1.12 × 10−4 −4.67 × 10−6
Dunham-Rydberg 18.27687 0.31902 −0.11668 1.12 × 10−4 3.31 × 10−5
DVR-ILJ 19.14233 0.23676 0.00002 7.23 × 10−5 −4.97 × 10−7
Rn–Og Dunham-ILJ 19.14251 0.23674 0.00002 7.27 × 10−5 −4.67 × 10−7
DVR-Rydberg 19.16661 0.24821 0.00018 7.37 × 10−5 −5.68 × 10−7
Dunham-Rydberg 19.16455 0.28186 −0.10395 7.37 × 10−5 2.04 × 10−5
DVR-ILJ 25.09701 0.18846 −0.00007 4.33 × 10−5 −1.84 × 10−7
Og–Og Dunham-ILJ 25.09726 0.18846 −0.00007 4.29 × 10−5 −1.73 × 10−7
DVR-Rydberg 24.44471 0.20151 −0.00024 4.54 × 10−5 −2.68 × 10−7
Dunham-Rydberg 24.44251 0.21228 −0.05290 3.85 × 10−5 6.75 × 10−6
Og–Og (correlation formulas) DVR-ILJ 16.9382 0.1740 0.00001 4.41 × 10−5 2.48 × 10−7
Og–Og (correlation formulas) Dunham-ILJ 16.9384 0.1740 0.00001 4.44 × 10−5 2.37 × 10−7
Og–Og (Jerabek et al.12) 22.39 0.2132 0.00017 1.8 × 10−4
Og–Og (Saue et al.13) 21.74


Fig. 2 shows the behavior of lifetime as a function of temperature (which ranged from 200 to 500 K) for each molecule studied. This figure reveals that the lifetime determined by both ILJ and Rydberg PEC essentially has the same behavior for the entire temperature range considered. The He–Og lifetime obtained via Rydberg PEC and ILJ PEC are slightly above 1.0 picosecond and they are the ones with the lowest lifetime values between 200 and 500 K. All other systems have a lifetime above 1.0 picosecond within the temperature range of 200–500 K. Thus, according to Wolfgang's study,58 a lifetime of a complex above 1.0 picoseconds means that the well of the potential energy is deep enough to assure its stability and the interacting complex must be considered stable. Therefore, based on this condition, the Og–He system is close to the edge of instability and it is expected, as this system has only one vibrational level inside of the well of ILJ and Rydberg PEC. On the other hand, the Og–Og dimer has the longest lifetime in the entire temperature range considered, confirming its great stability.


image file: d2cp04456k-f2.tif
Fig. 2 Lifetime as a function of temperature for the Ng–Og (Ng = He, Ne, Ar, Kr, Xe, Rn, and Og) systems using ILJ PEC, (a) and (c), and Rydberg PEC, (b) and (d).

5 Conclusions

Using the relativistic 4-component Dirac–Fock method and with the inclusion of quantum electrodynamics effects at the Gaunt interaction level and with BSSE correction, this work presents the ground state electronic potential energy curves, the vibrational energies, spectroscopic constants, and lifetime of the He–Og, Ne–Og, Ar–Og, Kr–Og, Xe–Og, Rn–Og, and Og–Og dimers.

The inclusion of quantum electrodynamics effects reduces the value of the dissociation energy of the studied dimers, which can be attributed to the fact that interactions between electrons are no longer treated as instantaneous. Furthermore, the percentage contribution of the quantum electrodynamics effect on the dissociation energy ranged from 0.48% (for the He–Og dimer) to 9.63% (for the Rn–Og dimer). This fact is also reflected in the number of vibrational levels, mainly in the dimers composed of the heaviest noble gases.

The current results obtained for the Og–Og spectroscopic constants (mainly those obtained with Dunham-Rydberg) agree well with the theoretical results of Jerabek and Saue. The lifetime as a function of temperature (which ranged from 200 to 500 K) indicates that the Og–He molecule is close to the edge of instability as expected, because this dimer has only one vibrational level inside both ILJ and Rydberg PEC. It is expected that the results of this study can motivate future spectroscopy experiments involving the He, Ne, Ar, Kr, Xe, and Rn noble gases with the oganesson (Z = 118) super-heavy element.

In the present work, the ILJ function has been adopted to fit the ab initio points. However, the obtained De and Re are not only fitting parameters, but they exhibit an appropriate meaning since they are related to the fundamental physical properties of the interacting partners. This is confirmed (see Table 1) by the good agreement (except for Og–Og) between the fitting values and predictions of correlation formulas that relate the value of the basic potential parameters with the electronic polarizability of the involved partners. Moreover, ILJ provides an asymptotic attraction, where the leading C6 coefficient, given by C6 = DeRe6 scales along with the Ng–Og family as those predicted by correlation formulas, and for any system, their absolute values are within ≈20%.

An accurate analysis of basic intermolecular force components involved in all Ng–Og (Ng = He, Ne, Ar, Kr, Xe, and Rn) family of dimers confirmed the nature of van der Waals of the global interaction. This is also proved by the behavior of the equilibrium distance Re whose values, along the Ng–Og family, remain confined in the restricted range of Re = 4.50 ± 0.25 Å, while the dissociation energy De changes by a factor of 40± 20. This is typical behavior of systems whose repulsion component, controlling Re, is effectively dominated by the size of the most polarizable atom (Og), while De varies according to the polarizability change of the Ng partner. The use of correlation formulas, defined in terms of polarizability, which represents the basic physical properties useful to scale both size repulsion and dispersion attraction, permitted us to evaluate not only the C6 dispersion coefficient for Ng–Og systems but also the basic Re, De, and C6 interaction features of the Og–Og dimer. It has been also emphasized that with respect to the other Ng, Og exhibits an unexpected high Tbp value which determines a range of temperature, defining the liquid stability, significantly larger than 100 K. This anomalous behavior probably arises from non-conventional many-body attractive interaction contributions controlled by high relativistic effects.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors gratefully acknowledge the financial support from the Brazilian Research Councils: CAPES, CNPq and FAPDF. LGMM would like to acknowledge Desenvolvimento Científico e Tecnologico (CNPQ, under grant 408085/2021-5) and also Universidade Federal de Sao João del Rei (CCO-UFSJ).

Notes and references

  1. E. Eliav, U. Kaldor, Y. Ishikawa and P. Pyykkö, Element 118: The First Rare Gas with an Electron Affinity, Phys. Rev. Lett., 1996, 77, 5350–5352,  DOI:10.1103/PhysRevLett.77.5350.
  2. Y. T. Oganessian, V. K. Utyonkov, Y. V. Lobanov, F. S. Abdullin, A. N. Polyakov, R. N. Sagaidak, I. V. Shirokovsky, Y. S. Tsyganov, A. A. Voinov, G. G. Gulbekian, S. L. Bogomolov, B. N. Gikal, A. N. Mezentsev, S. Iliev, V. G. Subbotin, A. M. Sukhov, K. Subotic, V. I. Zagrebaev, G. K. Vostokin, M. G. Itkis, K. J. Moody, J. B. Patin, D. A. Shaughnessy, M. A. Stoyer, N. J. Stoyer, P. A. Wilk, J. M. Kenneally, J. H. Landrum, J. F. Wild and R. W. Lougheed, Synthesis of the isotopes of elements 118 and 116 in the Cf249 and Cm245 + Ca48 fusion reactions, Phys. Rev. C: Nucl. Phys., 2006, 74, 044602,  DOI:10.1103/PhysRevC.74.044602.
  3. A. Turler and V. Pershina, Advances in the production and chemistry of the heaviest elements, Chem. Rev., 2013, 113, 1237–1312,  DOI:10.1021/cr3002438.
  4. P. Pyykkö, Theoretical chemistry of gold, Angew. Chem., Int. Ed., 2004, 43, 4412–4456,  DOI:10.1002/anie.200300624.
  5. E. Pastorczak and C. Corminboeuf, Perspective: Found in translation: Quantum chemical tools for grasping non-covalent interactions, J. Chem. Phys., 2017, 146, 120901,  DOI:10.1063/1.4978951.
  6. Q. Cui, Perspective: Quantum mechanical methods in biochemistry and biophysics, J. Chem. Phys., 2016, 145, 140901,  DOI:10.1063/1.4964410.
  7. A. D. Buckingham, P. W. Fowler and J. M. Hutson, Theoretical studies of van der Waals molecules and intermolecular forces, Chem. Rev., 1988, 88, 963–988,  DOI:10.1021/cr00088a008.
  8. K. Liu, Y. Kang, Z. Wang and X. Zhang, 25th Anniversary Article: Reversible and Adaptive Functional Supramolecular Materials: “Noncovalent Interaction” Matters, Adv. Mater., 2013, 25, 5530–5548,  DOI:10.1002/adma201302015.
  9. J. Řezáč and P. Hobza, Benchmark calculations of interaction energies in noncovalent complexes and their applications, Chem. Rev., 2016, 116, 5038–5071,  DOI:10.1021/acs.chemrev.5b00526.
  10. S. F. Boys and F. Bernardi, The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors, Mol. Phys., 1970, 19, 553–566,  DOI:10.1080/00268977000101561.
  11. R. Cambi, D. Cappelletti, G. Liuti and F. Pirani, Generalized correlations in terms of polarizability for van der Waals interaction potential parameter calculations, J. Chem. Phys., 1991, 95, 1852–1861,  DOI:10.1063/1.461035.
  12. P. Jerabek, O. R. Smits, J. M.l Mewes, K. A. Peterson and P. Schwerdtfeger, Solid Oganesson via a Many-Body Interaction Expansion Based on Relativistic Coupled-Cluster Theory and from Plane-Wave Relativistic Density Functional Theory, J. Phys. Chem. A, 2019, 123, 4201–4211,  DOI:10.1021/acs.jpca.9b01947.
  13. A. Shee, S. Knecht and T. Saue., A theoretical benchmark study of the spectroscopic constants of the very heavy rare gas dimers, Phys. Chem. Chem. Phys., 2015, 17, 10978–10985,  10.1039/c5cp01094b.
  14. O. R. Smits, J. M. Mewes, P. Jerabek and P. Schwerdtfeger, Oganesson: A Noble Gas Element That Is Neither Noble Nor a Gas, Angew. Chem., Int. Ed., 2020, 59, 23636–23640,  DOI:10.1002/anie.202011976.
  15. R. M. de Oliveira, L. G. M. de Macedo, T. F. da Cunha, F. Pirani and R. Gargano, A Spectroscopic Validation of the Improved Lennard-Jones Model, Molecules, 2021, 26, 3906,  DOI:10.3390/molecules26133906.
  16. W. F. Cunha, R. M. Oliveira, L. F. Roncaratti, J. B. L. Martins, G. M. Silva and R. Gargano, Rovibrational energies and spectroscopic constants for H2O–Ng complexes, J. Mol. Model., 2014, 20, 2498,  DOI:10.1007/s00894-014-2498-8.
  17. A. L. A. Oliveira, M. A. Silva, F. Pirani, L. G. M. de Macedo and R. Gargano, Hydrogen sulphide H2S and noble gases (Ng = He, Ne, Ar, Kr, Xe, Rn) complexes: A theoretical study of their dynamics, spectroscopy, and interactions, Int. J. Quantum. Chem., 2020, 120, e26266,  DOI:10.1002/qua.26266.
  18. F. M. Carvalho, A. S. Kiametis, A. L. A. Oliveira, F. Pirani and R. Gargano, Spectroscopy, lifetime, and charge-displacement of the methanol-noble gas complexes: An integrated experimental-theoretical investigation, Spectrochim. Acta, Part A, 2021, 246, 119049,  DOI:10.1016/j.saa.2020.119049.
  19. A. L. A. Oliveira, L. G. M. de Macedo, Y. A. de Oliveira Só, J. B. L. Martins, F. Pirani and R. Gargano, Nature and role of the weak intermolecular bond in enantiomeric conformations of H2O2-noble gas adducts: a chiral prototypical model, New J. Chem., 2021, 45, 8240–8247,  10.1039/D0NJ06135B.
  20. F. Pirani, S. Brizi, L. Roncaratti, P. Casavecchia, D. Cappelletti and F. Vecchiocattivi, Beyond the Lennard-Jones model: A simple and accurate potential function probed byhigh resolution scattering data useful for molecular dynamics simulations, Phys. Chem. Chem. Phys., 2008, 10, 5489–5503,  10.1039/B808524B.
  21. F. Pirani, M. Alberti, A. Castro, M. Moix Teixidor and D. Cappelletti, Atom-bond pairwise additive representation for intermolecular potential energy surfaces, Chem. Phys. Lett., 2004, 394, 37–44,  DOI:10.1016/j.cplett.2004.06.100.
  22. P. Candori, D. Cappelletti, S. Falcinelli, F. Pirani, L. F. Roncaratti, F. Tarantelli and F. Vecchiocattivi, Benchmarking a model potential for the investigation of intermolecular interactions, Phys. Scr., 2008, 78, 038102,  DOI:10.1088/0031-8949/78/03/038102.
  23. C. Douketis, G. Scoles, S. Marchetti, M. Zen and A. J. Thakkar, Intermolecular forces via hybrid Hartree-Fock-SCF plus damped dispersion (HFD) energy calculations. An improved spherical model, J. Chem. Phys., 1982, 76, 3057–3063,  DOI:10.1063/1.443345.
  24. K. T. Tang and J. P. Toennies, An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients, J. Chem. Phys., 1984, 80, 3726–3741,  DOI:10.1063/1.447150.
  25. K. T. Tang and J. P. Toennies, The van der Waals potentials between all the rare gas atoms from He to Rn, J. Chem. Phys., 2003, 118, 4976–4983,  DOI:10.1063/1.1543944.
  26. K. T. Tang, J. P. Toennies and C. L. You, Accurate Analytical He-He van der Waals Potential Based on Perturbation Theory, Phys. Rev. Lett., 1995, 74, 1546–1549,  DOI:10.1103/PhysRevLett.74.1546.
  27. S. Longo, P. Diomede, A. Laricchiuta, G. Colonna, M. Capitelli, D. Ascenzi, M. Scotoni, P. Tosi and F. Pirani, Lect. Notes Comput. Sci., 2008, 5072, 1131 Search PubMed.
  28. J. N. Murrel, S. Carter, S. C. Farantos, P. Huxley and A. J. C. Varandas, Molecular Potential Energy Functions, John Wiley & Sons, Chichester, Sussex, UK, 1984 Search PubMed.
  29. H. V. R. Vila, L. A. Leal, A. L. A. Fonseca and R. Gargano, Calculation of the H2+ rovibrational energies and spectroscopic constants in the 2π, 3dσ, 4dσ, 4fπ, 4fσ, 5gσ, and 6iσ electronic states, Int. J. Quantum Chem., 2012, 112, 829–833,  DOI:10.1002/qua.23070.
  30. E. N. C. Paura, W. F. da Cunha, P. H. de Oliveira Neto, G. M. e Silva, J. B. L. Martins and R. Gargano, Vibrational and Electronic Structure Analysis of a Carbon Dioxide Interaction with Functionalized Single-Walled Carbon Nanotubes, J. Phys. Chem. A, 2013, 3013(117), 2854–2861,  DOI:10.1021/jp312622s.
  31. W. F. Cunha, R. Gargano, E. Garcia, J. R. S. Politi, A. F. Albernaz and J. B. L. Martins, Rovibrational energy and spectroscopic constant calculations of CH4⋯H2O, CH4⋯CHF3, and H2O⋯CHF3 dimers, J. Mol. Model., 2014, 20, 2298,  DOI:10.1007/s00894-014-2298-1.
  32. J. Soares Neto and L. Costa, Numerical Generation of Optimized Discrete Variable Representations, Braz. J. Phys., 1998, 28, 1–11,  DOI:10.1590/S0103-97331998000100001.
  33. J. L. Dunham, The Energy Levels of a Rotating Vibrator, Phys. Rev., 1932, 41, 721–731,  DOI:10.1103/PhysRev.41.721.
  34. J. C. Slater, The theory of complex spectra, Phys. Rev., 1929, 34, 1293,  DOI:10.1103/PhysRev.34.1293.
  35. R. F. de Menezes, L. G. M. de Macedo, J. B. L. Martins, F. Pirani and R. Gargano, Investigation of strength and nature of the weak intermolecular bond in NH2 radical-noble gas atom adducts and evaluation of their basic spectroscopic features, Chem. Phys. Lett., 2021, 769, 138386,  DOI:10.1016/j.cplett.2021.138386.
  36. DIRAC17, a relativistic ab initio electronic structure program, Release DIRAC17 (2017), written by L. Visscher, H. J. A. Jensen, R. Bast, T. Saue, with contributions from V. Bakken, K. G. Dyall, S. Dubillard, U. Ekström, E. Eliav, T. Enevoldsen, E. Faßhauer, T. Fleig, O. Fossgaard, A. S. P. Gomes, E. D. Hedegård, T. Helgaker, J. Henriksson, M. Iliaš, Ch. R. Jacob, S. Knecht, S. Komorovský, O. Kullie, J. K. Lærdahl, C. V. Larsen, Y. S. Lee, H. S. Nataraj, M. K. Nayak, P. Norman, G. Olejniczak, J. Olsen, J. M. H. Olsen, Y. C. Park, J. K. Pedersen, M. Pernpointner, R.di Remigio, K. Ruud, P. Sałek, B. Schimmelpfennig, A. Shee, J. Sikkema, A. J. Thorvaldsen, J. Thyssen, J.van Stralen, S. Villaume, O. Visser, T. Winther, S. Yamamoto. http://www.diracprogram.org.
  37. K. Dyall and K. Faegri. Introduction to relativistic quantum chemistry, Oxford University Press, 2007 Search PubMed.
  38. T. Saue, Relativistic Hamiltonians for chemistry: A primer, ChemPhysChem, 2011, 12, 3077–3094,  DOI:10.1002/cphc.201100682.
  39. L. Visscher, Approximate molecular relativistic Dirac-Coulomb calculations using a simple Coulombic correction, Theor. Chem. Acc., 1997, 98, 68–70,  DOI:10.1007/s002140050280.
  40. T. Hangele and M. Dolg, Coupled-cluster and DFT studies of the Copernicium dimer including QED effects, Chem. Phys. Lett., 2014, 616, 222–225,  DOI:10.1016/j.cplett.2014.10.048.
  41. D. H. T. Amador, H. C. B. de Oliveira, J. Sambrano, R. Gargano and L. G. M. de Macedo, 4-Component correlated all-electron study on Eka-actinium Fluoride (E121F) including Gaunt interaction: Accurate analytical form, bonding and influence on rovibrational spectra, Chem. Phys. Lett., 2016, 662, 169–175,  DOI:10.1016/j.cplett.2016.09.025.
  42. O. Kullie and T. Saue., Range-separated density functional theory: a 4-component relativistic study of the rare gas dimers He2, Ne2, Ar2, Kr2, Xe2, Rn2 and Uuo2, Chem. Phys., 2012, 395, 54–62,  DOI:10.1016/j.chemphys.2011.06.024.
  43. P. M. W. Gill, R. D. Adamson and J. A. Pople., Coulomb-attenuated exchange energy density functionals, Mol. Phys., 1996, 88, 1005–1009,  DOI:10.1080/00268979609484488.
  44. I. C. Gerber and J. G. Angyan., London dispersion forces by range-separated hybrid density functional with second order perturbational corrections: The case of rare gas complexes, J. Chem. Phys., 2007, 126, 044103,  DOI:10.1063/1.2431644.
  45. T. H. Dunning, Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen, J. Chem. Phys., 1989, 90, 1007–1023,  DOI:10.1063/1.456153.
  46. R. A. Kendall, T. H. Dunning and R. J. Harrison., Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions, J. Chem. Phys., 1992, 96, 6796–6806,  DOI:10.1063/1.462569.
  47. D. E. Woon and T. H. Dunning Jr., Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon, J. Chem. Phys., 1993, 98, 1358–1371,  DOI:10.1063/1.464303.
  48. D. E. Woon and T. H. Dunning Jr., Gaussian Basis Sets for Use in Correlated Molecular Calculations. IV. Calculation of Static Electrical Response Properties, J. Chem. Phys., 1994, 100, 2975–2988,  DOI:10.1063/1.466439.
  49. K. G. Dyall, Relativistic quadruple-zeta and revised triple-zeta and double-zeta basis sets for the 4p, 5p, and 6p elements, Theor. Chem. Acc., 2006, 115, 441–447,  DOI:10.1007/s00214-006-0126-0.
  50. K. Dyall, Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the 7p elements, with atomic and molecular applications, Theor. Chem. Acc., 2012, 131, 1172–1174,  DOI:10.1007/s00214-012-1172-4.
  51. A. D. Powell and R. Dawes, Calculating potential energy curves with fixed-node diffusion Monte Carlo: CO and N2, J. Chem. Phys., 2016, 145, 224308,  DOI:10.1063/1.4971378.
  52. M. J. Jamieson, G. W. F. Drake and A. Dalgarno, Retarded dipole-dipole dispersion interaction potential for helium, Phys. Rev. A: At., Mol., Opt. Phys., 1995, 51, 3358,  DOI:10.1103/PhysRevA.51.3358.
  53. F. Nunzi, G. Pannacci, F. Tarantelli, L. Belpassi, D. Cappelletti, S. Falcinelli and F. Pirani, Leading interaction components in the structure and reactivity of noble gases compounds, Molecules, 2020, 25, 2367,  DOI:10.3390/molecules25102367.
  54. D. Cappelletti, E. Ronca, L. Belpassi, F. Tarantelli and F. Pirani, Revealing Charge-Transfer Effects in Gas-Phase Water Chemistry, Acc. Chem. Res., 2012, 45, 1571–1580,  DOI:10.1021/ar3000635.
  55. G. Liuti and F. Pirani, Regularities in van der Waals forces: correlation between the potential parameters and polarizability, Chem. Phys. Lett., 1985, 122, 245–250,  DOI:10.1016/0009-2614(85)80571-6.
  56. D. V. Fedorov, M. Sadhunkhn, M. Stöhr and A. Tkatchenko, Quantum-Mechanical Relation between Atomic Dipole Polarizability and the van der Waals Radius, Phys. Rev. Lett., 2018, 121, 183401,  DOI:10.1103/PhysRevLett.121.183401.
  57. D. R. Lide, Handbook of Chemistry and Physics, CRC Press, Florida, 67th edn, 2005, pp. 1986–1987 Search PubMed.
  58. R. Wolfgang, Energy and chemical reaction. II. intermediate complexes vs. direct mechanisms, Acc. Chem. Res., 1970, 3,  DOI:10.1021/ar50026a002.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2cp04456k

This journal is © the Owner Societies 2023