Next Article in Journal
Identification of 3-Methoxycarpachromene and Masticadienonic Acid as New Target Inhibitors against Trypanothione Reductase from Leishmania Infantum Using Molecular Docking and ADMET Prediction
Next Article in Special Issue
σ-Aromatic MAl6S6 (M = Ni, Pd, Pt) Stars Containing Planar Hexacoordinate Transition Metals
Previous Article in Journal
Novel Bis- and Mono-Pyrrolo[2,3-d]pyrimidine and Purine Derivatives: Synthesis, Computational Analysis and Antiproliferative Evaluation
Previous Article in Special Issue
Preamble from the Guest Editor of Special Issue “DFT Quantum-Chemical Calculation of Metal Clusters”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chirality and Relativistic Effects in Os3(CO)12

Nikolaev Institute of Inorganic Chemistry SB RAS, 3, Acad. Lavrentiev Ave., 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(11), 3333; https://doi.org/10.3390/molecules26113333
Submission received: 14 May 2021 / Revised: 28 May 2021 / Accepted: 30 May 2021 / Published: 1 June 2021
(This article belongs to the Special Issue DFT Quantum Chemical Calculation of Metal Clusters)

Abstract

:
The energy and structural parameters were obtained for all forms of the carbonyl complex of osmium Os3(CO)12 with D3h and D3 symmetries using density functional theory (DFT) methods. The calculations took into account various levels of relativistic effects, including those associated with nonconservation of spatial parity. It was shown that the ground state of Os3(CO)12 corresponds to the D3 symmetry and thus may be characterized either as left-twisted (D3S) or right-twisted (D3R). The D3S↔D3R transitions occur through the D3h transition state with an activation barrier of ~10–14 kJ/mol. Parity violation energy difference (PVED) between D3S and D3R states equals to ~5 × 10−10 kJ/mol. An unusual three-center exchange interaction was found inside the {Os3} fragment. It was found that the cooperative effects of the mutual influence of osmium atoms suppress the chirality of the electron system in the cluster.

1. Introduction

It is well known that enantiomerism is directly related to the origin of life on Earth. Many essential biological and chemical processes are stereoselective, involving only one enantiomer, which exists independently from its chiral counterpart. Separation of one enantiomer from another is not only important for practical use, but also sparks much interest as a fundamental problem, which constantly attracts research attention to the spatial and electronic structure of enantiomers [1,2,3]. In particular, especially interesting is how the parity violating weak nuclear forces may manifest in chiral molecules and may be somehow responsible for the choice of which enantiomer would prevail in living organisms [4,5,6,7].
Molecules with D3 symmetry are particularly interesting because they are characterized either with left-handed (D3S) or with right-handed (D3R) torsion, which allows these structures to be treated as chiral enantiomers. Previously, the detailed analysis of the 1,4-diazabicyclo[2.2.2]octane (DABCO) molecule with D3S (left-twisted), D3R (right-twisted), and D3h (untwisted) symmetries in gas phase and in Metal-organic Frameworks (MOFs) was performed [8,9]. There were expectations that MOFs with a DABCO linker may undergo a number of phase transitions related to enantiomers ordering [2,10], and that cooperative effects in such systems may help the experimental search of molecular parity violation effects [7,9,11]. However, the energy barrier between the D3S and D3R states of DABCO appears to be very low (<100 J/mol in gas phase [8] and ~5 kJ/mol in MOF [9]), which makes MOFs with DABCO difficult for experimental studies of such effects.
Similarly to the DABCO molecule, trinuclear transition metal cluster complexes may also have D3 symmetry. A well-known carbonyl osmium complex Os3(CO)12 (Figure 1) is widely used in synthesis in organometallic and inorganic chemistry [12,13,14]. It acts as a catalyst for a wide range of reactions [15]. It was considered as a model system for studies of mechanisms of photoinduced reactions [16,17,18] and vibrational and NMR spectroscopy [19,20,21]. In crystal, which is reported to have a monoclinic space group P21/n, the Os3(CO)12 molecules show approximate (pseudo-) D3h symmetry, where the three Os atoms form an almost equilateral triangle [22]. Previous detailed quantum chemical studies of structure and interatomic interactions in the Os3(CO)12 cluster have shown that in the gas phase, the ground state has the D3 symmetry, while the D3h symmetry refers to a transition state. It was also found that there is another transition state with C2v symmetry and higher energy, which, however, may occur at high temperatures [23,24,25].
In the present study, for the first time, the Os3(CO)12 cluster was considered as a chiral compound. We investigated what exactly is responsible for the stabilization of the twisted D3 structure compared to the untwisted D3h one. We also evaluated the impact of relativistic effects on the structure and energetics of the Os3(CO)12 clusters, including the parity violation energy shift due to the weak nuclear forces. We expected that due to the cooperative effects, such trinuclear transition metal clusters may become a good family of systems for the experimental molecular parity violation search.

2. Results and Discussion

2.1. Structure and Energetics of Os3(CO)12

The analysis of the interatomic distances in Os3(CO)12 (Table 1 and Table S1) shows that the best agreement between calculated and experimental X-ray diffraction (XRD) structures [22,26] was observed for the calculations with relativistic effects. Moreover, the difference between scalar (SR) and spin–orbit (SO) relativistic methods is rather small. However, they both are very different from the nonrelativistic (NR) level of theory. In this case, the D3h is the local minimum without imaginary frequencies in vibrational spectrum, while the D3 symmetry refers to just some point on potential energy surface (PES) and is characterized with two imaginary frequencies. The energy difference between the structures with D3 and D3h symmetries calculated at the relativistic level (SR and SO) indicates that the D3 state is the local minimum on the potential energy surface. At both SR and SO levels of theory, there are no imaginary frequencies for the structure with D3 symmetry, while D3h is a transition state with a single imaginary frequency. Gibbs free energy differences between D3 and D3h structures are 13.8 kJ/mol and 10.5 kJ/mol for SR and SO levels of theory, respectively (Figure 2). Previous calculations with relativistic effective core potential (ECP) also indicated that the D3 state is a local minimum [25]. Thus, it can be concluded that the D3 is the local minimum in the gas phase and the taking into account the relativistic effects at least at ECP level is necessary for correct description of the system.
The main difference between experimental XRD data and relativistic calculations (for both SR and SO methods) is found in the dihedral angles ∠C-Os-Os-C и ∠C3axes-Os-C, which characterize the degree of the twisting of the Os3(CO)12 structure. Such a difference could be related to the packing effects in the solid state. In order to compare the packing energy with the barrier between the cluster structures with D3S and D3R symmetries, the periodic calculations were performed. The packing energy is ~ 44 kJ/mol per cluster, while the barrier is ~10 kJ/mol. Thus, the packing effect is big enough to suppress the twisting of the Os3(CO)12 clusters in the solid state.
The energy difference between D3 and D3h structures characterizes the barrier between D3S and D3R enantiomers of Os3(CO)12 cluster. Without consideration of parity violating interactions, D3S and D3R enantiomers have exactly the same energy and electronic structure. The geometries of D3S and D3R enantiomers differ only by the sign of the Z coordinates of the atoms (Table S1). Thus, any of the enantiomers can be used to analyze the interactions in the cluster.
The barrier between D3S and D3R is low (~10–14 kJ/mol). Thus, similarly to the DABCO molecule [1], at room temperature there should be a dynamical equilibrium between D3S and D3R structures resulting in a quasi-D3h structure.

2.2. Characterization of the Interactions in the Os3(CO)12 Cluster

The energy decomposition analysis (EDA) of the interactions between Os(CO)4 fragments in the Os3(CO)12 cluster show that the orbital interactions’ contributions to stabilization of the D3 structure are larger than for the D3h form (Table 2). On the contrary, the steric interactions destabilize D3 compared to the D3h. The distortions of the Os(CO)4 fragments in the clusters, as well as dispersion interactions, have only minor impacts on the energy difference between structures with D3 and D3h symmetries and, in general, compensate each other. Thus, the stronger orbital interactions between the Os(CO)4 fragments are the main factors that are responsible for the stability of the D3 form.
ELF analysis of the interactions in the osmium triangle showed the presence of the three valence V(Os, Os) basins (Figure 3), which indicate the covalency of the interactions between Os atoms in both D3 and D3h structures of Os3(CO)12 cluster. Also, both structures are characterized with the single trisynaptic basin in the center of Os3 triangle. However, this basin does not have common borders with any of the core basins (Figures S1 and S2). It shares borders only with valence V(Os, Os) basins and, thus, this basin indicates the electron exchange between V(Os, Os) basins. To our best knowledge, it is the first example of such ELF topology. Formally, this basin should be classified as V(V(Os, Os)3). In the list of the molecular orbitals, the HOMO-2 (Figure 4) could be the one responsible for this V(V(Os, Os)3) basin. Note that the populations of the V(V(Os, Os)3) basins are very small (0.03e and 0.01e for D3 and D3h structures, respectively), however it should provide the exchange between V(Os, Os). This assumption is consistent with results of previous QTAIM analysis about three center interactions in Os3(CO)12 cluster [23].

2.3. Chirality and Parity Violation in Os3(CO)12

Our calculations indicate that relativistic effects define the D3 symmetry of the ground state of Os3(CO)12 molecule, which has the two variants: left-twisted D3S and right-twisted D3R. The next level of relativistic effects would be the calculation of the contributions of the nuclear weak forces to the energy of the system. These forces violate spatial parity symmetry; thus, they make the energy of enantiomers slightly different. Experimental evidence of such effect have been observed on nuclear and atomic levels [27]. However, parity violation effects on molecular level have not been detected experimentally yet [7]. Thus, it is important to look for good candidates for experimental search of molecular parity violation.
The major contribution to PVED comes from the nuclear-spin-independent part of the parity-violating Hamiltonian [28]:
H P V S I = G F 2 2 n N n u c ( Q W ( n ) γ 5 ρ n ( r ) ) ,
where GF is the Fermi-coupling constant with a value of 2.22255 × 10−14 a.u.; γ5 is the fifth Dirac gamma matrix, which refers to the electron chirality operator; ρn(r) is the normalized nucleon density; Nnuc is the number of nuclei in the molecule; and QW(n) is the weak nuclear charge of nucleus n, which depends on the number of protons Z(n) and neutrons N(n) in the nucleus. It is given by the expression:
Q W ( n ) = Z ( n ) ( 1 4 sin 2 θ W ) N ( n )
with θW being the Weinberg angle.
Previous Density Functional Theory (DFT), Hartree-Fock method (HF), and Many Body Perturbation Theory (MBPT) calculations for compounds of various six-row elements showed rather low values of PVED: ~3 × 10−12 kJ/mol for Hg [9], ~6 × 10−11 kJ/mol for Re and ~2 × 10−10 for Os [29], up to ~2 × 10−10 for Bi [30], and the largest would be ~8 × 10−9 kJ/mol for the hypothetical analogue of hydrogen peroxide—Po2H2 [31]. However, some of those estimations were made using approximate two-component relativistic methods, while our previous work showed that results of two-component PVED calculations in some systems may differ from the four-component results by up to 30% [11]. Therefore, in this work we used only fully relativistic four-component methods for calculations of PVED.
We may try to estimate the characteristic PVED values for Os compounds. As weak interactions are local and largely depend on the atomic number, the main contribution would be from the heaviest nucleus in the molecule (in our case Os) and may be estimated as follows [32]:
PVED     G F α 3 Z 4 N K r η
where α is the fine-structure constant, Z is the charge of the heaviest nucleus in the molecule, N is the number of neutrons in the heaviest nucleus in the molecule, Kr is some enhancement factor for relativistic effects (for the first row atoms Kr ~ 1, while for the heavier period six elements Kr ~10), and η is the dimensionless molecular asymmetry factor. The estimations of typical η in organic molecules fall in the range from ~10−9 [33] to ~10−2 [34]. Therefore, according to Equation (3), PVED in Os compounds may be preliminarily estimated as ~8 × 10−16–~8 × 10−9 kJ/mol.
Especially interesting are the possible cooperative effects in trinuclear Os3(CO)12 clusters. The contributions from the three Os atoms may just sum up, but also it may either additionally increase or suppress the chirality of the system. To extract such effects, we may compare the PVED values for Os3(CO)12 with some mononuclear reference system. We chose the hypothetical OsOSSeTe complex, which is analogous to the osmium tetraoxide. The results are shown in Table 3. The PVED values for Os3(CO)12 are closer to the upper estimate for Os compounds. We can see that Generalized Gradient Approximation (GGA) methods consistently give lesser values than HF, while hybrid and range-separated methods predictably give half-way values. This is supported by the electron chirality density pictures (Figure 5). The most surprising is the strong dependence of PVED for OsOSSeTe from the method used. This is due to the fact that contribution from the second heavy atom Te is of the opposite sign and DFT GGA methods, for some reason, redistribute the electron chirality in such a way that it becomes much higher on the Te atom. However, even the largest PVED value for OsOSSeTe (HF) is an order of magnitude less than the corresponding value for Os3(CO)12.
Another model for testing the cooperative effects is to calculate PVED for stretched and contracted versions of the Os3(CO)12 cluster, where only the Os-Os distances change. We may expect the decrease of geometric chirality parameter η with the increase of the Os-Os distances. However, the calculated PVED values actually grow with increase of the Os-Os distances (Figure 6) for both HF and DFT PBE methods. This indicates that stronger interaction between Os atoms rather suppresses the electron chirality in the system.

3. Materials and Methods

Computational Details

Geometry optimization of the Os3(CO)12 cluster with D3 and D3h symmetries was performed in ADF2020 program [40,41] with all-electron Slater’s type TZ2P basis set [42], TPSSh [43] density functional, and Grimme D4(EEQ) [44] dispersion corrections. The calculations were performed in the gas phase. In order to analyze the influence of the relativistic effects on the geometry of the cluster, the calculations were performed without relativistic approximation (NR), with scalar relativistic (SR), and with spin-orbit (SO) zero-order regular approximation (ZORA) [45,46]. To minimize the basis set superposition error (BSSE), the single point calculations at TPSSh + D4(EEQ) + SR/QZ4P level of theory were performed for the energy decomposition analysis (EDA) [47] and electron localization function calculations (ELF) [48,49]. ELF analysis was performed in the dgrid-4.6 program [50] on the discreet mesh with the step of 0.05 a.u.
The periodic calculation of the Os3(CO)12 cluster in the solid state were performed in the BAND2020 program [51,52] with SCAN density functional [53], all-electron TZP basis set, and SR ZORA approximation. The starting geometry for the periodic calculations was taken from the experimental XRD structural data [22].
Calculations of parity-violating energy difference (PVED) between left- and right-twisted structures of Os3(CO)12 in the gas phase were performed in Dirac-19 [32,54] with a fully relativistic four-component Dirac–Coulomb Hamiltonian. An all-electron double-zeta dyall.ae2z relativistic basis set [55], in combination with Hartree–Fock and various DFT methods, was used. To highlight the cooperative effect of three Os atoms, we compared the PVED for Os3(CO)12 with PVED for a hypothetical compound OsOSSeTe, where there was only one Os atom. The structure of OsOSSeTe in the gas phase was also optimized in ADF2020 at the TPSSh + D4(EEQ) + SR/QZ4P level of theory.

4. Conclusions

In our study, for the first time, the chirality of the Os3(CO)12 was investigated. With an account of relativistic effects, the twisted structure with the D3 symmetry refers to the local minimum on potential energy surface, while the nonchiral D3h structure refers to the transition state with a single imaginary frequency. The chiral D3 structure is, apparently, stabilized by the relativistic effects. Moreover, the EDA analysis of the interactions in Os3(CO)12 showed that D3 structure has the stronger orbital interactions between Os(CO)4 fragments.
The ELF topological analysis showed the unusual pattern of the basins. The trisynaptic basin in the center of Os triangle has no borders with any of the core basins, apparently, indicating the exchange between disynaptic V(Os, Os) basins.
We also estimated the PVED values for the Os3(CO)12 cluster. At present, the obtained values are the largest predicted, with the exception of the hypothetical H2Po2 system [31]. Moreover, the study of cooperative effects in PVED of Os3(CO)12 clusters showed that the stronger interaction between Os atoms rather reduces the electron chirality in the system. Knowing that, we may hope to construct in future some molecular system with even larger PVED.

Supplementary Materials

The following are available online, Table S1: Coordinates of the Os3(CO)12 cluster optimized with Non Relativistic (NR), Scalar Relativistic (SR) and Spin-Orbit (SO) approximations at TPSSh + D4(EEQ)/TZ2P level of theory, Figure S1: ELF basins for Os3CO12 clusters with D3 (top) and D3h (bottom) symmetries, Figure S2: Critical points between V(Os, Os) and V(V(Os, Os)3) basins.

Author Contributions

M.R.R.—quantum chemical studies of energy and structural parameters, conceptualization, and writing; I.V.M.—calculation of parity violation effects and writing; S.G.K. and Y.V.M.—discussion, writing, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in supplementary material.

Conflicts of Interest

There are no conflicts to declare.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Kozlova, S.G.; Mirzaeva, I.V.; Ryzhikov, M.R. DABCO molecule in the M2(C8H4O4)2·C6H12N2 (M = Co, Ni, Cu, Zn) metal-organic frameworks. Coord. Chem. Rev. 2018, 376, 62–74. [Google Scholar] [CrossRef]
  2. Kozlova, S.; Ryzhikov, M.; Pishchur, D.; Mirzaeva, I. Overview of Low-Temperature Heat Capacity Data for Zn2(C8H4O4)2 C6H12N2 and the Salam Hypothesis. Symmetry 2019, 11, 657. [Google Scholar] [CrossRef] [Green Version]
  3. Puneet, P.; Singh, S.; Fujiki, M.; Nandan, B. Handed Mirror Symmetry Breaking at the Photo-Excited State of π-Conjugated Rotamers in Solutions. Symmetry 2021, 13, 272. [Google Scholar] [CrossRef]
  4. Salam, A. The role of chirality in the origin of life. J. Mol. Evol. 1991, 33, 105–113. [Google Scholar] [CrossRef]
  5. Berger, R. Parity-Violation Effects in Molecules. In Relativistic Electronic Structure Theory, Part 2: Applications; Schwerdtfeger, P., Ed.; Elsevier B.V.: Amsterdam, The Netherlands, 2004; pp. 188–288. [Google Scholar]
  6. MacDermott, A.J.; Fu, T.; Nakatsuka, R.; Coleman, A.P.; Hyde, G.O. Parity-Violating Energy Shifts of Murchison L-Amino Acids are Consistent with an Electroweak Origin of Meteorite L-Enantiomeric Excesses Orig. Life Evol. Biosph. 2009, 39, 459–478. [Google Scholar] [CrossRef]
  7. Schwerdtfeger, P. The Search for Parity Violation in Chiral Molecules. In Computational Spectroscopy: Methods, Experiments and Applications; Grunenberg, J., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010; pp. 201–221. [Google Scholar]
  8. Nizovtsev, A.S.; Ryzhikov, M.R.; Kozlova, S.G. Structural flexibility of DABCO. Ab initio and DFT benchmark study. Chem. Phys. Lett. 2017, 667, 87–90. [Google Scholar] [CrossRef]
  9. Mirzaeva, I.V.; Kozlova, S.G. Parity violating energy difference for mirror conformers of DABCO linker between two M2+ cations (M = Zn, Cd, and Hg). J. Chem. Phys. 2018, 149, 214302. [Google Scholar] [CrossRef] [PubMed]
  10. Gabuda, S.P.; Kozlova, S.G. Abnormal difference between the mobilities of left- and right-twisted conformations of C6H12N2 roto-symmetrical molecules at very low temperatures. J. Chem. Phys. 2015, 142, 234302. [Google Scholar] [CrossRef]
  11. Mirzaeva, I.V.; Kozlova, S.G. Computational estimation of parity violation effects in a metal-organic framework containing DABCO. Chem. Phys. Lett. 2017, 687, 110–115. [Google Scholar] [CrossRef]
  12. Vergeer, F.W.; Kleverlaan, C.J.; Matousek, P.; Towrie, M.; Stufkens, D.J.; Hartl, F. Redox Control of Light-Induced Charge Separation in a Transition Metal Cluster: Photochemistry of a Methyl Viologen-Substituted [Os3(CO)10(α-diimine)] Cluster. Inorg. Chem. 2005, 44, 1319–1331. [Google Scholar] [CrossRef] [PubMed]
  13. Nijhoff, J.; Bakker, M.J.; Hartl, F.; Stufkens, D.J.; Fu; van Eldik, R. Photochemistry of the Triangular Clusters Os3(CO)10(α-diimine): Homolysis of an Os−Os Bond and Solvent-Dependent Formation of Biradicals and Zwitterions. Inorg. Chem. 1998, 37, 661–668. [Google Scholar] [CrossRef]
  14. Adams, R.D.; Captain, B.; Zhu, L. A New Tris(diphenylstannylene)triosmium Carbonyl Cluster Complex and Its Reactions with Pt(PBut3)2 and Pt(PPh3)4. Organometallics 2006, 25, 2049–2054. [Google Scholar] [CrossRef]
  15. Kuznetsov, B.N.; Kovalchuk, V.I.; Chesnokov, N.V.; Mikova, N.M.; Naimushina, L.V. Cluster carbonyls of Os, Fe and Fe-Rh on oxide supports: Synthesis and properties. Macromol. Symp. 1998, 136, 41–46. [Google Scholar] [CrossRef]
  16. Bentsen, J.G.; Wrighton, M.S. Wavelength-, medium-, and temperature-dependent competition between photosubstitution and photofragmentation in triruthenium dodecacarbonyl and triiron dodecacarbonyl: Detection and characterization of coordinatively unsaturated trimetal undecacarbonyl complexes. J. Am. Chem. Soc. 1987, 109, 4530–4544. [Google Scholar]
  17. Leadbeater, N.E. Control of the photochemistry of Ru3(CO)12 and Os3(CO)12 by variation of the solvent. J. Organomet. Chem. 1999, 573, 211–216. [Google Scholar] [CrossRef]
  18. Schalk, O.; Josefsson, I.; Richter, R.; Prince, K.C.; Odelius, M.; Mucke, M. Ionization and photofragmentation of Ru3(CO)12 and Os3(CO)12. J. Chem. Phys. 2015, 143, 154305. [Google Scholar] [CrossRef] [Green Version]
  19. Huggins, D.K.; Flitcroft, N.; Kaesz, H.D. Infrared Spectrum of Osmium Tetracarbonyl Trimer, Os3(CO)12; Assignment of CO and MC Stretching Absorptions. Inorg. Chem. 1965, 4, 166–169. [Google Scholar] [CrossRef]
  20. Yan, S.; Seidel, M.T.; Zhang, Z.; Leong, W.K.; Tan, H.-S. Ultrafast vibrational relaxation dynamics of carbonyl stretching modes in Os3(CO)12. J. Chem. Phys. 2011, 135, 024501. [Google Scholar] [CrossRef]
  21. Koridze, A.A.; Kizas, O.A.; Astakhova, N.M.; Petrovskii, P.V.; Grishin, Y.K. Internuclear exchange of carbonyl groups in Os3(CO)12: Coupling constants J(187Os–13C) in trinuclear osmium carbonyls. J. Chem. Soc. Chem. Commun. 1981, 853–855. [Google Scholar] [CrossRef]
  22. Churchill, M.R.; DeBoer, B.G. Structural studies on polynuclear osmium carbonyl hydrides. 1. Crystal structures of the isomorphous species H2Os3(CO)11 and Os3(CO)12. Role of an equatorial μ2-bridging hydride ligand in perturbing the arrangement of carbonyl ligands in a triangular cluster. Inorg. Chem. 1977, 16, 878–884. [Google Scholar]
  23. Van der Maelen, J.F.; García-Granda, S.; Cabeza, J.A. Theoretical topological analysis of the electron density in a series of triosmium carbonyl clusters: [Os3(CO)12], [Os3(μ-H)2(CO)10], [Os3(μ-H)(μ-OH)(CO)10], and [Os3(μ-H)(μ-Cl)(CO)10]. Comput. Theor. Chem. 2011, 968, 55–63. [Google Scholar] [CrossRef]
  24. Hunstock, E.; Mealli, C.; Calhorda, M.J.; Reinhold, J. Molecular Structures of M2(CO)9 and M3(CO)12 (M = Fe, Ru, Os): New Theoretical Insights. Inorg. Chem. 1999, 38, 5053–5060. [Google Scholar] [CrossRef]
  25. Li, Q.-S.; Xu, B.; Xie, Y.; King, R.B.; Schaefer, H.F. Unsaturated trinuclear osmium carbonyls: Comparison with their iron analogues. Dalt. Trans. 2007, 4312–4322. [Google Scholar] [CrossRef]
  26. Corey, E.R.; Dahl, L.F. The Molecular and Crystal Structure of Os3(CO)12. Inorg. Chem. 1962, 1, 521–526. [Google Scholar] [CrossRef]
  27. Roberts, B.M.; Dzuba, V.A.; Flambaum, V.V. Parity and Time-Reversal Violation in Atomic Systems. Annu. Rev. Nucl. Part. Sci. 2015, 65, 63–86. [Google Scholar] [CrossRef] [Green Version]
  28. Bast, R.; Koers, A.; Gomes, A.S.P.; Iliaš, M.; Visscher, L.; Schwerdtfeger, P.; Saue, T. Analysis of parity violation in chiral molecules. Phys. Chem. Chem. Phys. 2011, 13, 864–876. [Google Scholar] [CrossRef]
  29. Schwerdtfeger, P.; Bast, R. Large Parity Violation Effects in the Vibrational Spectrum of Organometallic Compounds. J. Am. Chem. Soc. 2004, 126, 1652–1653. [Google Scholar] [CrossRef]
  30. Faglioni, F.; Lazzeretti, P. Parity-violation effect on vibrational spectra. Phys. Rev. A 2003, 67, 032101. [Google Scholar] [CrossRef]
  31. Berger, R.; van Wüllen, C. Density functional calculations of molecular parity-violating effects within the zeroth-order regular approximation. J. Chem. Phys. 2005, 122, 134316. [Google Scholar] [CrossRef]
  32. Laerdahl, J.; Schwerdtfeger, P. Fully relativistic ab initio calculations of the energies of chiral molecules including parity-violating weak interactions. Phys. Rev. A. 1999, 60, 4439–4453. [Google Scholar] [CrossRef]
  33. Rein, D.W.; Hegstrom, R.A.; Sandars, P.G.H. Parity non-conserving energy difference between mirror image molecules. Phys. Lett. A 1979, 71, 499–502. [Google Scholar] [CrossRef]
  34. Harris, R.A.; Stodolsky, L. Quantum beats in optical activity and weak interactions. Phys. Lett. B 1978, 78, 313–317. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
  37. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  39. Yanai, T.; Tew, D.; Handy, N. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  40. ADF, Version 2020.102, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. Available online: https://www.scm.com/product/adf/ (accessed on 14 May 2021).
  41. te Velde, G.; Bickelhaupt, F.M.; Baerends, E.J.; Fonseca Guerra, C.A.; van Gisbergen, S.J.; Snijders, J.G.; Ziegler, T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931–967. [Google Scholar] [CrossRef]
  42. Van Lenthe, E.; Baerends, E.J. Optimized Slater-type basis sets for the elements 1–118. J. Comput. Chem. 2003, 24, 1142–1156. [Google Scholar] [CrossRef] [PubMed]
  43. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Caldeweyher, E.; Ehlert, S.; Hansen, A.; Neugebauer, H.; Spicher, S.; Bannwarth, C.; Grimme, S. A generally applicable atomic-charge dependent London dispersion correction. J. Chem. Phys. 2019, 150, 154122. [Google Scholar] [CrossRef]
  45. van Lenthe, E.; Ehlers, A.; Baerends, E.-J. Geometry optimizations in the zero order regular approximation for relativistic effects. J. Chem. Phys. 1999, 110, 8943–8953. [Google Scholar] [CrossRef] [Green Version]
  46. van Lenthe, E.; Snijders, J.G.; Baerends, E.J. The zero–order regular approximation for relativistic effects: The effect of spin–orbit coupling in closed shell molecules. J. Chem. Phys. 1996, 105, 6505–6516. [Google Scholar] [CrossRef] [Green Version]
  47. Ziegler, T.; Rauk, A. On the calculation of bonding energies by the Hartree Fock Slater method. Theor. Chim. Acta. 1977, 46, 1–10. [Google Scholar] [CrossRef]
  48. Savin, A.; Jepsen, O.; Flad, J.; Andersen, O.K.; Preuss, H.; von Schnering, H.G. Electron Localization in Solid-State Structures of the Elements: The Diamond Structure. Angew. Chem. Int. Ed. English. 1992, 31, 187–188. [Google Scholar] [CrossRef]
  49. Silvi, B.; Savin, A. Classification of chemical bonds based on topological analysis of electron localization functions. Nature 1994, 371, 683–686. [Google Scholar] [CrossRef]
  50. Kohout, M. DGrid, Version 4.6, Radebeul, 2011. Available online: http://www.cpfs.mpg.de/~kohout/dgrid.html (accessed on 14 May 2021).
  51. BAND, Version 2020.102, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands. Available online: https://www.scm.com/product/band_periodicdft/ (accessed on 14 May 2021).
  52. te Velde, G.; Baerends, E.J. Precise density-functional method for periodic structures. Phys. Rev. B 1991, 44, 7888–7903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sun, J.; Ruzsinszky, A.; Perdew, J.P. Strongly Constrained and Appropriately Normed Semilocal Density Functional. Phys. Rev. Lett. 2015, 115, 036402. [Google Scholar] [CrossRef] [Green Version]
  54. Saue, T.; Bast, R.; Gomes, A.S.P.; Jensen, H.J.A.; Visscher, L.; Aucar, I.A.; Di Remigio, R.; Dyall, K.G.; Eliav, E.; Fasshauer, E.; et al. The DIRAC code for relativistic molecular calculations. J. Chem. Phys. 2020, 152, 204104. [Google Scholar] [CrossRef]
  55. Dyall, K.G. Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the 6d elements Rf–Cn. Theor. Chem. Acc. 2011, 129, 603–613. [Google Scholar] [CrossRef]
Figure 1. The structure of Os3(CO)12 cluster corresponding to the local minimum on PES.
Figure 1. The structure of Os3(CO)12 cluster corresponding to the local minimum on PES.
Molecules 26 03333 g001
Figure 2. The barrier height for the D3S↔D3R enatiomerization through the D3h transition state calculated at nonrelativistic (NR), scalar (SR) and spin-orbit (SO) relativistic levels of theory.
Figure 2. The barrier height for the D3S↔D3R enatiomerization through the D3h transition state calculated at nonrelativistic (NR), scalar (SR) and spin-orbit (SO) relativistic levels of theory.
Molecules 26 03333 g002
Figure 3. ELF slice planes for Os3(CO)12 clusters with D3 (a) and D3h (b) symmetries.
Figure 3. ELF slice planes for Os3(CO)12 clusters with D3 (a) and D3h (b) symmetries.
Molecules 26 03333 g003
Figure 4. HOMO-2 orbitals for Os3(CO)12 clusters with D3 (a) and D3h (b) symmetries.
Figure 4. HOMO-2 orbitals for Os3(CO)12 clusters with D3 (a) and D3h (b) symmetries.
Molecules 26 03333 g004
Figure 5. Electron chirality density (γ5) calculated at HF (a), DFT PBE0 (b), and PBE (c) levels of theory.
Figure 5. Electron chirality density (γ5) calculated at HF (a), DFT PBE0 (b), and PBE (c) levels of theory.
Molecules 26 03333 g005
Figure 6. PVED as a function of Os-Os distance change, zero refers to the equilibrium Os-Os distance.
Figure 6. PVED as a function of Os-Os distance change, zero refers to the equilibrium Os-Os distance.
Molecules 26 03333 g006
Table 1. Interatomic distances (Å), dihedral angles (º), energy differences (kJ/mol), and number of imaginary frequencies (ni) for Os3(CO)12 clusters with D3 and D3h symmetries. Error estimates shown in parentheses for average (<x>) distances or angles are the exterior estimates of the precision of the average value given by [Ʃ(<x> − x)2/(m2 − m)]1/2 (m is the number of x values) [22]. The TPSSh + D4(EEQ)/TZ2P and MPW1PW91/SDD are the theoretical levels used for geometry optimization (for details see Section 3.1). NR, SR, SO and ECP are the relativistic levels used in quantum chemical calculations. The average experimental distances and angles were taken from X-ray diffraction (XRD) data [22,26].
Table 1. Interatomic distances (Å), dihedral angles (º), energy differences (kJ/mol), and number of imaginary frequencies (ni) for Os3(CO)12 clusters with D3 and D3h symmetries. Error estimates shown in parentheses for average (<x>) distances or angles are the exterior estimates of the precision of the average value given by [Ʃ(<x> − x)2/(m2 − m)]1/2 (m is the number of x values) [22]. The TPSSh + D4(EEQ)/TZ2P and MPW1PW91/SDD are the theoretical levels used for geometry optimization (for details see Section 3.1). NR, SR, SO and ECP are the relativistic levels used in quantum chemical calculations. The average experimental distances and angles were taken from X-ray diffraction (XRD) data [22,26].
TPSSh + D4(EEQ)/
TZ2P
TPSSh + D4(EEQ)/
TZ2P
XRD [22]XRD [26]MPW1PW91
/SDD [25]
SymmetryD3D3hPseudo-D3hPseudo-D3hD3D3h
Relativity levelNRSRSONRSRSO ECPECP
d(Os-Os)2.8082.8692.8692.8012.8832.884<2.877>(3)<2.881>(4)2.8952.907
d(Os-Ceq)1.9681.9161.9141.9661.9151.914<1.912>(7)<1.919>(36)1.9171.917
d(Os-Cax)2.0012.111.9551.9921.9541.953<1.946>(6)<1.973>(12)1.9531.95
∠C-Os-Os-C37.229.329.2000<2.1>(5)<1.3>(1.3)31.40
∠C3axes-Os-C21.216.816.8000<1.2>(5)<0.7>(1.0)180
ΔE000−3.07.57.7--−2.30
ΔG000−2.513.810.5----
ni200011--01
Table 2. Results of energy decomposition analysis (EDA) for interactions between Os(CO)4 fragments in D3 and D3h forms of Os3(CO)12 calculated at TPSSh + D4(EEQ) + SR/QZ4P//TPSSh + D4(EEQ) + SR/TZ2P level of theory. The formation energies of the fragments (Efrag(Os(CO)4) as well as energies of steric (ESteric), orbital (EOrbital), dispersion (EDisp) and total (EInt) interactions between fragments are given in kJ/mol.
Table 2. Results of energy decomposition analysis (EDA) for interactions between Os(CO)4 fragments in D3 and D3h forms of Os3(CO)12 calculated at TPSSh + D4(EEQ) + SR/QZ4P//TPSSh + D4(EEQ) + SR/TZ2P level of theory. The formation energies of the fragments (Efrag(Os(CO)4) as well as energies of steric (ESteric), orbital (EOrbital), dispersion (EDisp) and total (EInt) interactions between fragments are given in kJ/mol.
Efrag(Os(CO)4)EStericEOrbitalEDispEInt
D3−7838.3520.1−1191.6−118.4−789.9
D3h−7839.6489.5−1153.1−114.2−778.2
∆(D3−D3h)1.330.5−38.5−4.2−11.7
Table 3. PVED (kJ/mol) calculated with different methods: Hartree-Fock (HF) and DFT with functionals PBE [35], PBE0 [36], BLYP [37], B3LYP [38], CAMB3LYP [39].
Table 3. PVED (kJ/mol) calculated with different methods: Hartree-Fock (HF) and DFT with functionals PBE [35], PBE0 [36], BLYP [37], B3LYP [38], CAMB3LYP [39].
HFPBEPBE0BLYPB3LYPCAMB3LYP
Os3(CO)126.66 × 10−103.87 × 10−104.61 × 10−104.14 × 10−104.77 × 10−105.51 × 10−10
OsOSSeTe4.08 × 10−115.11 × 10−142.23 × 10−133.40 × 10−141.17 × 10−131.35 × 10−12
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ryzhikov, M.R.; Mirzaeva, I.V.; Kozlova, S.G.; Mironov, Y.V. Chirality and Relativistic Effects in Os3(CO)12. Molecules 2021, 26, 3333. https://doi.org/10.3390/molecules26113333

AMA Style

Ryzhikov MR, Mirzaeva IV, Kozlova SG, Mironov YV. Chirality and Relativistic Effects in Os3(CO)12. Molecules. 2021; 26(11):3333. https://doi.org/10.3390/molecules26113333

Chicago/Turabian Style

Ryzhikov, Maxim R., Irina V. Mirzaeva, Svetlana G. Kozlova, and Yuri V. Mironov. 2021. "Chirality and Relativistic Effects in Os3(CO)12" Molecules 26, no. 11: 3333. https://doi.org/10.3390/molecules26113333

Article Metrics

Back to TopTop