Next Article in Journal
Magnetic Field-Controlled Electrical Conductivity in AA Bilayer Graphene
Next Article in Special Issue
IR Spectroscopy of Vacancy Clusters (Amber Centers) in CVD Diamonds Nanostructured by Fast Neutron Irradiation
Previous Article in Journal
Arsenic, Iron, and Manganese Adsorption in Single and Trinary Heavy Metal Solution Systems by Bamboo-Derived Biochars
Previous Article in Special Issue
On the Highly Ordered Graphene Structure of Non-Graphitic Carbons (NGCs)—A Wide-Angle Neutron Scattering (WANS) Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermochemistry of the Smallest Hyperbolic Paraboloid Hydrocarbon: A High-Level Quantum Chemical Perspective

School of Science and Technology, University of New England, Armidale, NSW 2351, Australia
Submission received: 18 January 2023 / Revised: 7 April 2023 / Accepted: 10 April 2023 / Published: 19 April 2023
(This article belongs to the Special Issue Characterization of Disorder in Carbons (2nd Edition))

Abstract

:
[5.5.5.5]hexaene is a [12]annulene ring with a symmetrically bound carbon atom in its center. This is the smallest hydrocarbon with a hyperbolic paraboloid shape. [5.5.5.5]hexaene and related hydrocarbons are important building blocks in organic and materials chemistry. For example, penta-graphene—a puckered 2D allotrope of carbon—is comprised of similar repeating subunits. Here, we investigate the thermochemical and kinetic properties of [5.5.5.5]hexaene at the CCSD(T) level by means of the G4 thermochemical protocol. We find that this system is energetically stable relative to its isomeric forms. For example, isomers containing a phenyl ring with one or more acetylenic side chains are higher in energy by ∆H298 = 17.5–51.4 kJ mol−1. [5.5.5.5]hexaene can undergo skeletal inversion via a completely planar transition structure; however, the activation energy for this process is ∆H298 = 249.2 kJ mol−1 at the G4 level. This demonstrates the high configurational stability of [5.5.5.5]hexaene towards skeletal inversion. [5.5.5.5]hexaene can also undergo a π-bond shift reaction which proceeds via a relatively low-lying transition structure with an activation energy of ∆H298 = 67.6 kJ mol−1. Therefore, this process is expected to proceed rapidly at room temperature.

Graphical Abstract

1. Introduction

Carbon atoms that are surrounded by an annulene ring have attracted considerable attention over the past 50 years [1,2,3,4,5]. A [12]annulene with a symmetrically bound carbon in its center is of particular importance since it adopts a hyperbolic paraboloidal shape due to a central sp3 carbon surrounded by 12 sp2 carbons. Scheme 1 illustrates the hyperbolic paraboloidal structure of the [12]annulene–[5.5.5.5]hexaene (1). In this structure, the central sp3 carbon is located at the saddle point of the surface created by the four adjacent pentagons. It is well known that hyperbolic paraboloid surfaces of the equation z = axy possess rigidity in the x and y directions [6,7]. Indeed, this may explain the high rigidity of penta-graphene—a two-dimensional allotrope of carbon. Penta-graphene, which consists of repeating hyperbolic paraboloid units, each comprising four pentagon rings sharing a vertex, has been suggested to be more rigid than graphene [8,9].
Computational quantum chemistry is a branch of chemistry that uses computational simulations to study the chemical properties of molecules and materials [10,11]. Computational quantum chemistry provides means for the rational design and development of new molecules and materials with tailored chemical properties. Due to significant advances in quantum theory and supercomputers, computational simulations are capable of unprecedented predictive accuracy. The present work focuses on the structural, thermodynamic, and kinetic properties of [5.5.5.5]hexaene. In particular, we use high-level composite ab initio methods [12,13] to show that [5.5.5.5]hexaene is (i) configurationally stable with respect to structural inversion, (ii) energetically stable relative to its isomeric forms, and (iii) has a relatively low lying π-bond shift transition state.

2. Computational Details

In order to obtain accurate thermochemical and kinetic properties for the hydrocarbons considered in the present work, calculations have been carried out using the high-level, ab initio, G4 and G4(MP2) procedures [14,15,16]. These procedures combine coupled-cluster with single, double, and perturbative triple excitation (CCSD(T)) calculations as well as second- and fourth-order Møller–Plesset perturbation theories to approximate the CCSD(T) energy in conjunction with a triple-ζ-quality basis set (CCSD(T)/TZ). Both theories have been found to provide thermochemical and kinetic properties for hydrocarbons with chemical accuracy (arbitrarily defined as 1 kcal mol−1 = 4.2 kJ mol−1) [14,15,16,17,18,19,20,21,22,23,24,25,26,27]. G4 theory is computationally more demanding and in most cases more accurate than G4(MP2) theory; therefore, the main text reports the G4 results whilst the G4(MP2) results are provided as Supporting Information. The geometries of all structures have been obtained at the B3LYP/6-31G(2df,p) level of theory [28,29,30] as prescribed in the G4 and G4(MP2) protocols. Harmonic vibrational frequencies have been calculated at the same level of theory to confirm that all stationary points are equilibrium structures (i.e., with all real frequencies) or transition structures (i.e., with one imaginary frequency). The connectivity of the transition structures were confirmed by performing intrinsic reaction coordinate calculations [31,32]. Zero-point vibrational energies and enthalpic temperature corrections have been obtained from the harmonic frequencies and scaled by appropriate scaling factors as recommended in G4 and G4(MP2) theories [15,16]. All DFT and ab initio calculations were performed using the Gaussian 09 program suite [33].
We have additionally assessed several DFT functionals for their ability to compute the isomerization energies of the highly unsaturated C13H8 isomers considered in the present work relative to the bottom-of-the-well G4 reference values. The considered DFT methods, ordered by their rung on Jacob’s ladder [34], are the generalized gradient approximation (GGA) functionals BLYP [28,35], PBE [36], BPBE [35,36] and BP86 [35,37] the meta-GGA functionals TPSS [38] and MN15-L [39] the hybrid-GGAs B3LYP [28,29,30] B3PW91 [27,40], PBE0 [41] and the range-separated CAM-B3LYP [42] the hybrid-meta GGAs M05-2X [43], M06-2X [44], MN15 [39], BMK [45], and PW6B95 [46]. All calculations have been performed in conjunction with the def2-QZVPP basis set [47]. With the exception of the Minnesota functionals, which account for dispersion in the functional form, the empirical D3 dispersion correction has been used in conjunction with the Becke–Johnson damping potential (denoted by D3BJ) [48,49,50].

3. Results and Discussion

[5.5.5.5]hexaene (1, Scheme 1) has the molecular formula of C13H8, Scheme 2 depicts key structural C13H8 isomers 214 that have been identified using the ChemSpider database [51]. [5.5.5.5]hexaene and the C13H8 isomers in Scheme 2 are, by definition, highly unsaturated. As such, most of the C13H8 isomers are polyynes (isomers 212, Scheme 2), and many isomers include a phenyl ring (isomers 210 and 13, Scheme 2). Isomer 1 does not contain a phenyl ring; however, it contains a nonplanar conjugated π-system. Therefore, [5.5.5.5]hexaene is expected to be relatively stable compared to the isomers in Scheme 2. Yet, to the best of our knowledge, [5.5.5.5]hexaene (1) has not been synthesized; however, several of the C13H8 isomers are synthetically accessible, for example, isomers 2 and 3 [52,53,54]. In the present work, we examine the relative stability of the C13H8 isomers and show that [5.5.5.5]hexaene (1) is the energetically most stable isomer within this space of highly unsaturated structures. This stability may be attributed to the conjugated [12]annulene ring. We also show that, with the exception of isomer 13, which contains a highly strained cyclobutene ring, all of the isomers that contain a phenyl ring are relatively stable with isomerization energies of ∆H298 = 17.5–52.8 kJ mol−1 relative to isomer 1. For example, isomers 2 and 3 are associated with isomerization energies of 17.5 and 24.3 kJ mol−1, respectively, relative to isomer 1.
Table 1 gives the G4 isomerization energies on the electronic (∆Ee), enthalpic at 0 K (∆H0), and enthalpic at 298 K (∆H298) potential energy surfaces (PESs). Remarkably, our G4 calculations show that 1 is the energetically most stable isomer on the ∆Ee, ∆H0, and ∆H298 PESs. From here onwards, we will focus on the ∆H298 values; however, we note that the same trends are observed on the ∆Ee and ∆H0 PESs. We also note that there is a reasonably good agreement between the G4 and G4(MP2) values. For example, the largest deviation of 12.7 kJ mol−1 between the two theories is obtained for isomer 14. All the G4(MP2) isomerization energies are given in Table S1 of the Supporting Information. All the isomers 210 contain an aromatic phenyl ring with one (or more) linear acetylenic carbon chain. Isomer 2, which contains a single (−C≡C−)3 chain terminated with a methyl group, is less stable than [5.5.5.5]hexaene (1) by 17.5 kJ mol−1. Moving the methyl group to the Ph ring in the ortho, meta, and para positions relative to the acetylenic chain (isomers 35) destabilizes the isomers. In particular, we obtain isomerization energies of 24.3 (3), 28.0 (5), and 28.3 (4) kJ mol−1 relative to isomer 1. Isomers 69 all involve three acetylene substituents and one methyl substituent. These isomers, along with isomer 10, are energetically less stable than isomer 1 by 46.8–52.8 kJ mol−1. All the other isomers (1114) involve long acetylenic or cumulenic chains or a highly strained cyclobutene ring. These isomers are highly energetic and lie as much as 277.3–488.2 kJ mol−1 above isomer 1.
Having established that [5.5.5.5]hexaene (1) is a relatively stable C13H8 isomer, it is of interest to examine its structure in more detail. This is a symmetric molecule of D2 symmetry. The bond lengths and angles are shown in Figure 1a. We begin by noting that even though [5.5.5.5]hexaene is highly nonplanar, the double and single bonds of the outer [12]annulene ring exhibit bond length alternation, indicating a partly delocalized π-system. The double C=C bonds are of similar lengths to those of typical C=C bonds, namely, 1.360 and 1.343 Å, and the single C–C bonds are shorter than typical single C–C bonds, namely, 1.441 and 1.468 Å (Figure 1). For comparison, the length of the conjugated double bonds in cyclopentadiene is 1.346 Å, i.e., in between the lengths of the double bonds in 1, and the length of the single bond sandwiched between the two double bonds in cyclopentadiene is 1.468 Å.
[5.5.5.5]hexaene and related systems have been initially proposed in the quest to form a tetracoordinate carbon with a planar configuration [1,2,3,4,55,56,57]. Indeed, the outer conjugated annulene imposes, to some extent, a planar configuration around the central sp3 carbon. This is demonstrated by bond angles around this carbon being smaller than the ideal tetrahedral angle. The two bond angles around the central carbon are 96.4° and 98.1°. The planarity of the central carbon can be assessed by examining the angle between the planes of the two opposite cyclopentadiene rings, which is 56.0° (cfr. to an angle of 90° in a perfect tetrahedral configuration). We note that the angle between the planes of the two opposite cyclopentene rings is very similar and amounts to 56.9°.
The above geometrical parameters show that the central carbon in [5.5.5.5]hexaene is closer to planarity than a tetrahedral carbon; it is therefore of interest to examine the reaction barrier height for the skeletal inversion in this structure. The inversion transition structure is completely planar and has a D2h symmetry (see Figure 1b and Scheme 3a). The four bonds between the central carbon and the carbons of the [12]annulene ring are significantly elongated relative to the equilibrium structure (1, Figure 1a). All of the bonds of the [12]annulene ring are slightly shortened relative to the equilibrium structure; therefore, the bond length alternation in the [12]annulene ring is maintained. As expected, the skeletal inversion has a high barrier height of ∆H298 = 249.2 kJ mol−1 at the G4 level, demonstrating the high configurational stability of [5.5.5.5]hexaene towards skeletal inversion. The high energy of the planar TS for the skeletal inversion may be attributed to two factors: (i) the strain energy associated with a planar structure of four fused five-membered rings sharing a central carbon atom, and (ii) the antiaromaticity of this planar TS, which involves 12 (4n) π-electrons [5,58].
It is of interest to examine whether the conjugated [12]annulene ring can undergo a π-bond shift reaction (illustrated schematically in Scheme 3). Figure 1c shows the optimized transition structure for this reaction. The geometric configuration of the central carbon atom in this transition structure is very similar to that in the equilibrium structure. The bond length alternation in this TS is significantly reduced, with the difference in length between the two types of bonds being merely 0.039 Å. The activation energy for the π-bond shift reaction is ∆H298 = 67.6 kJ mol−1 at the G4 level. Therefore, this process is expected to proceed rapidly at room temperature.
Finally, it is useful to examine the performance of a representative selection of DFT methods across the rungs of Jacob’s Ladder in order to accommodate future computational investigations of larger fragments of penta-graphene and related structures. Table 2 gives the root mean square deviations (RMSDs), mean absolute deviations (MADs), mean signed deviations (MSDs), and largest deviations (LDs) for the considered DFT methods. We begin by noting that highly unsaturated hydrocarbons have been shown to be an extremely challenging target for most DFT functionals [20,59]. Let us begin with examining the performance of the pure GGA methods. BLYP-D3BJ results in an unacceptably large RMSD of 87.7 kJ mol−1, which is mostly attributed to the poor performance of the LYP correlation functional. Replacing the LYP functional with either the P86 or PBE correlation functionals results in significant improvements in performance. Namely, the BP86-D3BJ and BPBE-D3BJ methods result in RMSDs of 27.8 and 23.6 kJ mol−1, respectively. We note that PBE-D3BJ results in a similar performance to BP86-D3BJ, albeit PBE-D3BJ is associated with a much smaller MSD of merely −2.2 kJ mol−1 (Table 2). The considered meta-GGA methods (TPSS-D3BJ and MN15-L) do not offer an improvement over the best-performing GGA methods. Moving on to the hybrid-GGA methods, B3LYP-D3BJ shows poor performance, which is considerably improved by the range-separated CAM-B3LYP method. PBE0-D3BJ results in an RMSD of 23.1 kJ mol−1, which is a noticeable improvement over the pure PBE-D3BJ method. Overall, the hybrid-GGA B3PW91-D3BJ provides the best performance of all the considered DFT methods with an RMSD of 14.9 kJ mol−1. Moving on to the hybrid meta-GGA methods, M06-2X provides poor performance with an RMSD of 47.9 kJ mol−1. Interestingly, its predecessor M05-2X results in an RMSD of just 22.0 kJ mol−1. We note that both M05-2X and M06-2X include similar amounts of exact Hartree–Fock exchange, namely, 56% and 54%, respectively. However, M06-2X is more heavily parametrized than M05-2X, which seems to work less well for the highly challenging isomerization energies at hand. Both MN15 and PW6B95-D3BJ do not offer an improvement over the performance of M05-2X. However, BMK-D3BJ results in a significantly lower RMSD of 16.2 kJ mol−1 (which is similar to that of B3PW91-D3BJ for which we obtain an RMSD of 14.9 kJ mol−1).

4. Conclusions

In this work, we use the high-level G4 composite ab initio method to investigate the thermodynamic and kinetic properties of the smallest prototypical hyperbolic paraboloidal hydrocarbon—[5.5.5.5]hexaene (1). We find that this system is energetically stable relative to its isomeric forms. For example, isomers containing a phenyl ring with one or more acetylenic side chains are higher in energy on the enthalpic potential energy surface at 298 K by ∆H298 = 17.5–51.4 kJ mol−1, whereas long acetylenic carbon chains are higher in energy by as much as ∆H298 = 277.3–488.2 kJ mol−1. [5.5.5.5]hexaene can undergo skeletal inversion via a completely planar transition structure; however, the activation energy for this process is relatively high, being ∆H298 = 249.2 kJ mol−1 at the G4 level. This demonstrates the high structural stability of [5.5.5.5]hexaene. Finally, we find that the transition structure for the π-bond shift reaction of the conjugated [12]annulene ring is relatively low-lying with an activation energy of ∆H298 = 67.6 kJ mol−1. Therefore, this process is expected to proceed rapidly at room temperature. We hope that these accurate theoretical results will inspire further experimental explorations of [5.5.5.5]hexaene.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/c9020041/s1: G4(MP2) reaction energies and barrier heights for all the reactions considered in the present work (Table S1); optimized geometries for all the species considered in this work (Table S2).

Funding

This research received no external funding.

Data Availability Statement

The data that support the findings of this study are available in the supplementary material of this article and from the corresponding author upon reasonable request.

Acknowledgments

We gratefully acknowledge the generous allocation of computing time from the National Computational Infrastructure (NCI) National Facility, and system administration support provided by the Faculty of Science Agriculture Business and Law to the Linux cluster of the Karton group.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Hoffmann, R.; Alder, R.W.; Wilcox, C.F. Planar tetracoordinate carbon. J. Am. Chem. Soc. 1970, 92, 4992. [Google Scholar] [CrossRef]
  2. Chandrasekhar, J.; Wurthwein, E.-U.; Schleyer, P.V.R. On the planarity of tetracoordinate carbon enclosed by annulene perimeters. Tetrahedron 1981, 37, 921. [Google Scholar] [CrossRef]
  3. Minkin, V.I.; Minyaev, R.M.; Hoffmann, R. Non-classical structures of organic compounds: Unusual stereochemistry and hypercoordination. Russ. Chem. Rev. 2002, 71, 869. [Google Scholar] [CrossRef]
  4. Keese, R. Carbon Flatland: Planar Tetracoordinate Carbon and Fenestranes. Chem. Rev. 2006, 106, 4787. [Google Scholar] [CrossRef] [PubMed]
  5. Spitler, E.L.; Johnson, C.A.; Haley, M.M. Renaissance of Annulene Chemistry. Chem. Rev. 2006, 106, 5344. [Google Scholar] [CrossRef] [PubMed]
  6. Pellegrino, S. On the Rigidity of Triangulated Hyperbolic Paraboloids. Proc. R. Soc. Land. A 1988, 418, 425. [Google Scholar]
  7. Ten, L.V. Rigidity of complete surfaces of negative curvature that coincide with a hyperbolic paraboloid outside a compact region. Russ. Math. Surv. 1980, 35, 111. [Google Scholar] [CrossRef]
  8. Nazir, M.A.; Hassan, A.; Shen, Y.; Wang, Q. Research progress on penta-graphene and its related materials: Properties and applications. Nano Today 2022, 44, 101501. [Google Scholar] [CrossRef]
  9. Zhang, S.; Zhou, J.; Wang, Q.; Chen, X.; Kawazoe, Y.; Jena, P. Penta-graphene: A new carbon allotrope. Proc. Natl. Acad. Sci. USA 2015, 112, 2372. [Google Scholar] [CrossRef]
  10. Cramer, C.J. Essentials of Computational Chemistry: Theories and Models, 2nd ed.; Wiley: Hoboken, NJ, USA, 2004. [Google Scholar]
  11. Levine, I.N. Quantum Chemistry, 7th ed.; Prentice Hall: New York, NY, USA, 2013. [Google Scholar]
  12. Karton, A. Quantum mechanical thermochemical predictions 100 years after the Schrödinger equation. Annu. Rep. Comput. Chem. 2022, 18, 123. [Google Scholar]
  13. Karton, A. A computational chemist’s guide to accurate thermochemistry for organic molecules. WIREs Comput. Mol. Sci. 2016, 6, 292. [Google Scholar] [CrossRef]
  14. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-4 theory. J. Chem. Phys. 2007, 126, 084108. [Google Scholar] [CrossRef] [PubMed]
  15. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-4 theory using reduced order perturbation theory. J. Chem. Phys. 2007, 127, 124105. [Google Scholar] [CrossRef] [PubMed]
  16. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gn Theory. WIREs Comput. Mol. Sci. 2011, 1, 810. [Google Scholar] [CrossRef]
  17. Zheng, J.; Zhao, Y.; Truhlar, D.G. The DBH24/08 Database and Its Use to Assess Electronic Structure Model Chemistries for Chemical Reaction Barrier Heights. J. Chem. Theory Comput. 2009, 5, 808. [Google Scholar] [CrossRef] [PubMed]
  18. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Assessment of Gaussian-4 theory for energy barriers. Chem. Phys. Lett. 2010, 499, 168. [Google Scholar] [CrossRef]
  19. Karton, A.; O’Reilly, R.J.; Radom, L. Assessment of theoretical procedures for calculating barrier heights for a diverse set of water-catalyzed proton-transfer reactions. J. Phys. Chem. A 2012, 116, 4211. [Google Scholar] [CrossRef]
  20. Karton, A.; Martin, J.M.L. Explicitly correlated benchmark calculations on C8H8 isomer energy separations: How accurate are DFT, double-hybrid and composite ab initio procedures? Mol. Phys. 2012, 110, 2477. [Google Scholar] [CrossRef]
  21. Yu, L.-J.; Karton, A. Assessment of theoretical procedures for a diverse set of isomerization reactions involving double-bond migration in conjugated dienes. Chem. Phys. 2014, 441, 166. [Google Scholar] [CrossRef]
  22. Karton, A.; Goerigk, L. Accurate reaction barrier heights of pericyclic reactions: Surprisingly large deviations for the CBS-QB3 composite method and their consequences in DFT benchmark studies. J. Comput. Chem. 2015, 36, 622. [Google Scholar] [CrossRef]
  23. Yu, L.J.; Sarrami, F.; O’Reilly, R.J.; Karton, A. Reaction barrier heights for cycloreversion of heterocyclic rings: An Achilles’ heel for DFT and standard ab initio procedures. Chem. Phys. 2015, 458, 1–8. [Google Scholar] [CrossRef]
  24. Karton, A.; Schreiner, P.R.; Martin, J.M.L. Heats of formation of platonic hydrocarbon cages by means of high-level thermochemical procedures. J. Comput. Chem. 2016, 37, 49. [Google Scholar] [CrossRef] [PubMed]
  25. Karton, A. How reliable is DFT in predicting the relative energies of polycyclic aromatic hydrocarbon isomers? Comparison of functionals from different rungs of Jacob’s Ladder. J. Comput. Chem. 2017, 38, 370. [Google Scholar] [CrossRef]
  26. Karton, A.; Sylvetsky, N.; Martin, J.M.L. W4-17: A diverse and high-confidence dataset of atomization energies for benchmarking high-level electronic structure methods. J. Comput. Chem. 2017, 38, 2063. [Google Scholar] [CrossRef]
  27. Karton, A. Fullerenes Pose a Strain on Hybrid Density Functional Theory. J. Phys. Chem. A 2022, 126, 4709. [Google Scholar] [CrossRef]
  28. 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. [Google Scholar] [CrossRef] [PubMed]
  29. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648. [Google Scholar] [CrossRef]
  30. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623. [Google Scholar] [CrossRef]
  31. Gonzalez, C.; Schlegel, H.B. An improved algorithm for reaction path following. J. Chem. Phys. 1989, 90, 2154. [Google Scholar] [CrossRef]
  32. Gonzalez, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Chem. Phys. 1990, 94, 5523. [Google Scholar] [CrossRef]
  33. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 16, Revision E.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  34. Perdew, J.P.; Schmidt, K. Jacob’s ladder of density functional approximations for the exchange-correlation energy. AIP Conf. Proc. 2000, 577, 1–20. [Google Scholar]
  35. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098. [Google Scholar] [CrossRef] [PubMed]
  36. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  37. Perdew, J.P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33, 8822. [Google Scholar] [CrossRef] [PubMed]
  38. Tao, J.M.; 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]
  39. Yu, H.S.; He, X.; Li, S.L.; Truhlar, D.G. MN15: A Kohn–Sham global-hybrid exchange–correlation density functional with broad accuracy for multi-reference and single-reference systems and noncovalent interactions. Chem. Sci. 2016, 7, 5032. [Google Scholar] [CrossRef] [PubMed]
  40. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671. [Google Scholar] [CrossRef]
  41. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158. [Google Scholar] [CrossRef]
  42. Yanai, T.; Tew, D.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51. [Google Scholar] [CrossRef]
  43. Zhao, Y.; Schultz, N.E.; Truhlar, D.G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364. [Google Scholar] [CrossRef]
  44. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215. [Google Scholar]
  45. Boese, A.D.; Martin, J.M.L. Development of density functionals for thermochemical kinetics. J. Chem. Phys. 2004, 121, 3405. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, Y.; Truhlar, D.G. Design of Density Functionals That Are Broadly Accurate for Thermochemistry, Thermochemical Kinetics, and Nonbonded Interactions. J. Phys. Chem. A 2005, 109, 5656. [Google Scholar] [CrossRef] [PubMed]
  47. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [Google Scholar] [CrossRef] [PubMed]
  48. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  49. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456. [Google Scholar] [CrossRef] [PubMed]
  50. Becke, A.D.; Johnson, E.R. A density-functional model of the dispersion interaction. J. Chem. Phys. 2005, 123, 154101. [Google Scholar] [CrossRef]
  51. The C13H8 Isomers Were Identified Using the ChemSpider Database. ChemSpider Is a Comprehensive Freely Available Database with over 100 Million Chemical Structures. Available online: http://www.chemspider.com (accessed on 1 January 2023).
  52. Sorensen, J.S.; Sorensen, N.A. Studies Related to Naturally Occurring Acetylene Compounds. XXII* Correctional Studies on the Constitution of the Polyacetylenes of some Annual Coreopsis Species. Acta Chem. Scand. 1958, 12, 756. [Google Scholar] [CrossRef]
  53. Akiyama, S.; Misumi, S.; Nakagawa, M. Cyclic Acetylenes. IX. Synthesis of a Cyclic Octaacetylene Containing Anthracene Nuclei. Bull. Chem. Soc. Jpn. 1962, 35, 1829. [Google Scholar] [CrossRef]
  54. Jones, E.R.H.; Skattebøl, L.; Whiting, M.C. Researches on acetylenic compounds. Part LX. The synthesis of three natural polyacetylenic hydrocarbons. J. Chem. Soc. 1958, 1054. [Google Scholar] [CrossRef]
  55. Keese, R.; Pfenninger, A.; Roesle, A. Planarization of Tetracoordinate Carbon Atom. Synthesis of 13-oxa-14-oxo-pentacyclo [5.5.2.1.04,15010,15]pentadecane, a bridged ‘tetraquinacane’. Helv. Chim. Acta 1979, 62, 326. [Google Scholar] [CrossRef]
  56. Bohm, M.C.; Gleiter, R.; Schang, P. On the planarity of tetracyclo [5.5.1.04,13o10,13] tridecahexaene. Tetrahedron Lett. 1979, 20, 2575. [Google Scholar] [CrossRef]
  57. Kubiak, G.G.I. Mechanistic and Synthetic Studies on the Scope of the Weiss Reaction. II. Studies Directed toward the Preparation of Staurane-1,3,5,7,9,11-Hexaene on the Route towards Tetracoordinate Planar Carbon. Ph.D. Thesis, The University of Wisconsin, Milwaukee, WI, USA, 1989. Available online: https://www.proquest.com/docview/303806144?pq-origsite=gscholar&fromopenview=true (accessed on 1 January 2023).
  58. Klärner, F.-G. About the Antiaromaticity of Planar Cyclooctatetraene. Angew. Chem. Int. Ed. 2001, 40, 3977. [Google Scholar] [CrossRef]
  59. Karton, A.; Tarnopolsky, A.; Martin, J.M.L. Atomization energies of the carbon clusters Cn (n = 2–10) revisited by means of W4 theory as well as density functional, Gn, and CBS methods. Mol. Phys. 2009, 107, 977. [Google Scholar] [CrossRef]
Scheme 1. (a) Skeletal structure of a hyperbolic paraboloidal shape created by four pentagon rings sharing a common vertex, (b) the optimized structure of [5.5.5.5] hexaene, and (c) a 3D representation of the hyperbolic paraboloidal structure.
Scheme 1. (a) Skeletal structure of a hyperbolic paraboloidal shape created by four pentagon rings sharing a common vertex, (b) the optimized structure of [5.5.5.5] hexaene, and (c) a 3D representation of the hyperbolic paraboloidal structure.
Carbon 09 00041 sch001
Scheme 2. Illustration of key C13H8 isomers taken from the ChemSpider database.
Scheme 2. Illustration of key C13H8 isomers taken from the ChemSpider database.
Carbon 09 00041 sch002
Figure 1. Top and side views of the optimized B3LYP/6-31G(2df,p) structures of (a) 1 ([5.5.5.5]hexaene, D2 symmetry), (b) planar transition structure for structural inversion of 1-TSinv (D2h symmetry), and (c) transition structure for the double-bond shift in 1-TSshift (C2v symmetry); bond lengths are given in Å (black font), and angles are given in degrees (green and red font). See Table 1 for the relative G4 enthalpies at 298 K.
Figure 1. Top and side views of the optimized B3LYP/6-31G(2df,p) structures of (a) 1 ([5.5.5.5]hexaene, D2 symmetry), (b) planar transition structure for structural inversion of 1-TSinv (D2h symmetry), and (c) transition structure for the double-bond shift in 1-TSshift (C2v symmetry); bond lengths are given in Å (black font), and angles are given in degrees (green and red font). See Table 1 for the relative G4 enthalpies at 298 K.
Carbon 09 00041 g001
Scheme 3. Schematic illustration of the (a) skeletal inversion (1-TSinv) and (b) π-bond shift (1-TSshift) reactions in [5.5.5.5]hexaene.
Scheme 3. Schematic illustration of the (a) skeletal inversion (1-TSinv) and (b) π-bond shift (1-TSshift) reactions in [5.5.5.5]hexaene.
Carbon 09 00041 sch003
Table 1. CCSD(T) energies relative to isomer 1 (in kJ mol−1) obtained from G4 theory for the fourteen C13H8 isomers shown in Scheme 1 and Scheme 2 and transition structures in Figure 1.
Table 1. CCSD(T) energies relative to isomer 1 (in kJ mol−1) obtained from G4 theory for the fourteen C13H8 isomers shown in Scheme 1 and Scheme 2 and transition structures in Figure 1.
Struct.EeH0H298Struct.EeH0H298
10.00.00.01052.343.151.4
215.38.117.5961.242.052.8
325.815.224.311281.8263.2277.3
529.518.328.012283.6264.3278.7
429.818.728.313363.6359.9361.4
654.936.346.814498.7472.8488.2
754.836.746.91-TSinv a262.7247.5249.2
857.438.548.91-TSshift b74.668.467.6
a transition structure of 1 for skeletal inversion (see Figure 1b and Scheme 3a). b transition structure of 1 for π-bond shift (see Figure 1c and Scheme 3b).
Table 2. Statistical analysis for the performance of selected DFT procedures for calculating the relative energies of the isomers in Scheme 2 relative to the electronic ∆Ee reference values from G4 theory (in kJ mol−1) a,b.
Table 2. Statistical analysis for the performance of selected DFT procedures for calculating the relative energies of the isomers in Scheme 2 relative to the electronic ∆Ee reference values from G4 theory (in kJ mol−1) a,b.
RMSDMADMSDLD
BLYP-D3BJ87.779.8−76.2−143.4
BP86-D3BJ27.824.3−11.0−52.2
PBE-D3BJ27.424.1−2.2−42.5
BPBE-D3BJ23.619.91.8−34.2
TPSS-D3BJ38.432.8−26.3−69.6
MN15-L30.627.9−16.9−58.7
B3LYP-D3BJ62.057.9−54.1−91.8
CAM-B3LYP-D3BJ38.537.0−35.1−50.2
PBE0-D3BJ23.117.817.438.8
B3PW91-D3BJ14.913.32.722.0
M06-2X47.945.7−45.7−61.3
MN1524.722.6−22.6−36.5
PW6B95-D3BJ28.822.8−22.8−46.8
M05-2X22.020.7−19.8−34.0
BMK-D3BJ16.215.211.5−22.2
a RMSD = root mean square deviation, MAD = mean absolute deviation, MSD = mean signed deviation, and LD = largest deviation. b the cumulenic isomer 14 is highly challenging for DFT methods and is excluded from the evaluation dataset.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Karton, A. Thermochemistry of the Smallest Hyperbolic Paraboloid Hydrocarbon: A High-Level Quantum Chemical Perspective. C 2023, 9, 41. https://doi.org/10.3390/c9020041

AMA Style

Karton A. Thermochemistry of the Smallest Hyperbolic Paraboloid Hydrocarbon: A High-Level Quantum Chemical Perspective. C. 2023; 9(2):41. https://doi.org/10.3390/c9020041

Chicago/Turabian Style

Karton, Amir. 2023. "Thermochemistry of the Smallest Hyperbolic Paraboloid Hydrocarbon: A High-Level Quantum Chemical Perspective" C 9, no. 2: 41. https://doi.org/10.3390/c9020041

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop