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

Latent harmony in dicarbon between VB and MO theories through orthogonal hybridization of 3σg and 2σu

Ronglin Zhong , Min Zhang *, Hongliang Xu and Zhongmin Su *
Institute of Functional Material Chemistry & Local United Engineering Lab for Power Battery, Faculty of Chemistry, Northeast Normal University, Changchun 130024, China. E-mail: mzhang@nenu.edu.cn; zmsu@nenu.edu.cn

Received 11th September 2015 , Accepted 26th October 2015

First published on 27th October 2015


Abstract

Besides the classic double bond scheme, several novel schemes have been proposed to describe the nature of the chemical bond in dicarbon (C2), including a quadruple bond and a singlet diradical state. The results from a symmetry-broken CASSCF(8,8)/aug-cc-pVTZ study present a harmony between MO and VB theories, based on the orthogonal hybridization of the 3σg and 2σu orbitals together with the other six pristine valence orbitals. This scheme achieves the same bonding energy, RC–C, ωe and one electron density as that from the eight pristine valence orbitals. A quadruple bond scheme, identical to Prof. Shaik's result from VB theory, is achieved with the 4th bond energy in the range of 12.8–27.6 kcal mol−1. Meanwhile, the weight of a singlet open-shell configuration is the highest among all the possible configurations.


Introduction

Dicarbon (C2) is a simple molecule with just two atoms. However, it has aroused many fundamental questions, fascinating mysteries and active discussions in chemistry.1–6 It is a colourless gas and is very unstable. A tiny quantity of C2 can be prepared from electric arc strikes, and a good amount of C2 can exist in comets, stellar atmospheres, blue hydrocarbon flames, etc. The C–C bond length (RC–C) is 1.243 Å in its 1Σ+g ground state, and the corresponding vibration frequency (ωe) is 1855 cm−1.2,7 At least another 12 excited states have been observed experimentally and the RC–C values are found to be in a large range of 1.23 Å to 1.53 Å.2,4 Among them, two RC–C values bear shorter bond lengths in the excited states (3Σ+u and 1Σ+u) than in the ground state. Similar shorter bond lengths in the excited states have also been reported in its cation and anion (C2+ and C2).8–11 The RC–C of the dicarbide ion (C22−) in crystalline calcium carbide and lithium carbide is shorter than 1.20 Å,12,13 which is generally accepted as a traditional triple bond (1σ + 2π bonds) analogue of N2.

It is worthy of note that the ground RC–C distance of C2 is 1.243 Å, shorter than the length of any classic C[double bond, length as m-dash]C double bond (1σ + 1π bonds), such as in ethylene.14 Hence, Prof. Shaik15 pointed out that suspended π bonds may be responsible, since they prefer shorter lengths than σ bonds. The essential point of this assumption is that the occupied number of the 2σu antibond is approximately equal to that of the 2σg bond. However, 2σu is a weak antibonding orbital due to a lower occupied number, which can not counteract the stronger bonding of 2σg. Furthermore, the ωe value of C2 is higher than that of ethylene. Hence, soon after, the nature of the C2 bond was said to approach that of a triple bond (C[triple bond, length as m-dash]C).16,17 Based on the characteristics of a triple bond in C2, a scheme of a triple bond plus weak coupling by a pair of opposite spinning electrons was proposed in valence bond (VB) theory.18 The opposite spinning electron coupling energy was found to be ∼12–20.2 kcal mol−1 (ref. 19 and 20) at various levels of the theory. In this context, the corresponding 4th bonding scheme of C2 (Fig. 1c) was proposed with VB theory.19


image file: c5sc03437j-f1.tif
Fig. 1 The configurations of MO’s double bond, MO’s quadruple bond and VB's quadruple bond schemes in C2. The difference in the order of the occupied orbitals is highlighted with red lines.

A quadruple bond in C2 is certainly reasonable based on quantum mechanics,19,21 similar to that in [1.1.1]-propellane.22 However, whether it is the best picture to describe the ground state of C2 or not has aroused many active discussions.21,23–26 According to traditional molecular orbital (MO) theory, the quadruple bond configuration of C2 can be achieved by doubly exciting 2σu electrons to the 3σg orbitals. However, a calculation at the CASSCF(8,8)/cc-pVTZ level indicates that the weight of the double bond state (Fig. 1a) is 71.0% while the weight of the 2σu → 3σg quadruple bond counterpart (Fig. 1b) is only 13.6%.21 On the other hand, the weak 4th bond (the inset in Fig. 1c) was proposed by inequivalent hybridization of the 3σg and 2σu occupied states, covering only MO double and quadruple bond configurations, which does not seem perfect. Even in an egalitarian mode, the corresponding bond order of C2 is between two and three. What is the dominant configuration like if orthogonal hybridization of 3σg and 2σu is utilized and 100% weight of multi-configurational self-consistent field (MCSCF) is covered in the calculation? Is there any other configuration that can possess a higher weight in the ground state (1Σ+g) of C2 among all the possible configurations?

Results and discussion

As we know, the ground state of C2 is generally accepted as double π bonds in MO theory (shown in Fig. 1a). However, the quasi-degeneracy of the 2σu, 3σg and 1πu orbitals is well known in C2 and its ions, which results in many low-lying excited states of C2, C2 and C2+.1,2,4 The antiferromagnetic diradical characteristic of C2 has also been proposed by the finite-difference pseudopotential method, local spin analysis and VB theory.24–26 Silicon resides in the same column of the periodic table as carbon. The singlet diradical characteristic on a silicon (100) surface is well known.27,28 Moreover, dicarbon is a very unstable molecule with a short lifetime, which is easy to dimerize into C4 for instance.

As mentioned above, CASSCF(8,8)/aug-cc-pVTZ was used in our study because C2 has multi-reference configurations in nature. In the beginning, the level was benchmarked for the 1Σ+g ground state (KK2σ2g2xu2yu2u) and four low-lying (3Πu, 3Σ+u, 3Σg and 1Πu) excited states. All the potential energy curves are plotted in Fig. 2. Furthermore, the important data of these optimized states contrasting to the experimental data are listed in Table 1.


image file: c5sc03437j-f2.tif
Fig. 2 Plots of C2 potential energy curves of 1Σ+g state and four (3Πu, 3Σ+u, 3Σg and 1Πu) excited states at CASSCF(8,8)/aug-cc-pVTZ level. The bonding energy is relative to two isolated 3P carbon atoms at CASSCF(4,4)/aug-cc-pVTZ level. The vertical transition energies from 1Σ+g state are also shown.
Table 1 The most important theoretical data of the five states, including RC–Cs (Å) and vibrational frequencies (ωe, cm−1), together with their experimental data (shown in bracket)
  1Σ+g 3Πu 3Σ+u 3Σg 1Πu
R C–C (Å) 1.255 (1.243) 1.330 (1.312) 1.224 (1.230) 1.384 (1.369) 1.338 (1.318)
ΔErelat. (kcal mol−1) −142.5 (−142) −131.8 −114.9 −113.1 −104.7
ω e (cm−1) 1839 (1855) 1599 (1641) 1975 (2084) 1436 (1470) 1553 (1608)
Dominant configuration KK2σ2g KK2σ2g2u KK2σ2g2xu KK2σ2g2u KK2σ2g2u
2xu2yu2u 2xuαyuαg 2yuαuβg 2gαxuαyu 2xuαyuβg
EBO 2.15 1.90 2.74 1.92 1.90
Weight 71.0% 87.5% 85.1% 93.0% 89.9%
One electron density 0.014e (3σu) 0.020e (3σu) 0.016e (3σu) 0.026e (3σu) 0.027e (3σu)
0.115e (1πyg) 0.074e (1πyg) 0.109e (1πyg) 0.048e (1πyg) 0.041e (1πyg)
0.115e (1πxg) 0.098e (1πxg) 0.109e (1πxg) 0.048e (1πxg) 0.097e (1πxg)
0.393e (3σg) 0.965e (1πyu) 0.976e (3σg) 0.990e (1πxu) 0.985e (1πyu)
1.602e (2σu) 1.045e (3σg) 1.031e (2σu) 0.990e (1πyu) 1.023e (3σg)
1.888e (1πyu) 1.905e (1πxu) 1.891e (1πyu) 1.956e (2σu) 1.912e (1πxu)
1.888e (1πxu) 1.910e (2σu) 1.891e (1πxu) 1.966e (3σg) 1.933e (2σu)
1.984e (2σg) 1.982e (2σg) 1.977e (2σg) 1.978e (2σg) 1.982e (2σg)


The calculated RC–C and ωe values of the 1Σ+g state are close to the experimental values. For example, the bonding energy of the 1Σ+g state is −142.5 kcal mol−1, which is almost equivalent to the ideal RC–C bonding breakage of C2 based on the heats of formation.23 The RC–C differences of the other four states compared to the experimental data are only less than 0.02 Å. These results clearly show that the results from the CASSCF(8,8)/aug-cc-pVTZ level are reliable. In the ground state, 3σg is a weak bond with a one electron density of ∼0.4e. It contributes somewhat to the stabilization of C2. Hence, a triple bond scheme of C2 is also reasonable.16,17 However, the one electron density of the 2σu orbital is ∼1.6e, which is ∼1.2e higher than that of the 3σg orbital. If the populations of all the other three valence orbitals (1πxg, 1πyg and 3σu) are included, the relative 2σu antibond electrons are still ∼1.0e higher. In this context, the quadruple bond scheme is hard to be accepted by naive application of MO theory.

Based on the traditional valence MOs, it is hard to interpret the singlet diradical characteristic of the C2 ground state, while it has been shown through the finite-difference pseudopotential method and LSA analysis.25,26 Even though a singlet diradical state (KK2σ2g4uug) was achieved, it is still a 1Σ+u state. The diagram of 2σug occupation is shown in the left part of Fig. 3. As a result, the MO and VB theories fall into an apparent contradiction. In our opinion, this just indicates that opposite spinning electrons do not locate around the C–C bonds, which is in accordance with Prof. Shaik's proposal. However, the difficulty with the VB interpretation of the full configuration interaction (CI) wave function may be due to the nonorthogonal transform and the neglect of ∼15% weight of the configuration state functions (CSFs). Can VB quadruple bond schemes be reliable through the reformed valence MO orbitals? Is it feasible to describe the singlet diradical characteristic of C2 through the reformed valence MO orbitals simultaneously?


image file: c5sc03437j-f3.tif
Fig. 3 Equivalent hybridized orbitals of φL and φR from 2σu and 3σg localized around a single C atom (a) and the corresponding potential occupied styles before and after hybridization (b).

We hereby propose another scheme through hybrid orbitals of the 2σu and 3σg orbitals to φu+3σgL) and φu−3σgR), which was used as the initial active orbital for the CASSCF calculation. It is worthy of note that the φL and φR orbitals ensure that the single-occupied electrons are located outside of the C–C bond (right part of Fig. 3a). Besides, the other six pristine valence orbitals were maintained. Subsequently, the potential energy curve of the ground state of C2 was re-scanned at the same level with our orbitals. Compared with Prof. Shaik's strategy, the eight CAS orbitals were all orthogonal and the SCFs covered 100% of the weight. As expected, the new results show that the bonding energy, RC–C, ωe and one electron density are all the same as the values from the eight pristine valence orbitals (in the first column of Table 1). This is reasonable because the final result is only determined by the one electron densities in the original natural orbitals and the corresponding gradient analysis when all the CSFs are included in the simulations. CASSCF(8,8) and VBSCF(1764) span the same space of 1764 configurations in MO and VB theories, separately. Theoretically, the same electron correlation energy and bonding energy can be achieved if the same orbitals are adopted in the simulation. Unfortunately, orthogonal orbitals are adopted in the former, but the nonorthogonal orbitals are adopted in the latter.18

The difference between our and Prof. Frenking's CASSCF schemes21 is the input active orbitals and their corresponding CSF weights, which are affected by the input orbital styles. The five highest weight configurations of our scheme are shown in Fig. 4. The weight of the highest configuration (KK2σ2g2xu2yuφLφR) is 73.7%, around 2.7% higher than that of Prof. Frenking's (KK2σ2g2xu2yu2u). The results elucidate that the spin-localization of the 2σu electrons is, indeed, outside of the C–C bond, in accordance with the VB view.18,19 If the bonding nature is estimated by the highest weight configuration, the ground state of C2 is inclined to be a singlet diradical due to a little higher weight than the traditional double bond configuration.


image file: c5sc03437j-f4.tif
Fig. 4 Left: CASSCF orbitals, including φL and φR. Right: The corresponding five most important configurations of a full-valence CASSCF(8,8)/aug-cc-PVTZ calculation of C2, showing the coefficients and the weights of the configurations.

However, with our understanding from VB theory, the first three configurations have to be added up, since there is a combination of a major covalent structure (C–C), and two minor ionic ones (C–C+ and C+–C). They all belong to the 4th bond. To our surprise, the total weight of the three highest configurations is equal to the weight sum of the classic MO double and quadruple bond configurations (Fig. 3b), namely the initial value in Prof. Shaik's nonorthogonal scheme. Hence, the corresponding covalent component of the 4th bond is 84.8% (close to the weight of the 3Πu and 3Σu states). The opposite spin coupling energy between φLφR can not be achieved directly in our orthogonal schemes. If the vertical excited energy from the 1Σ+g to 3Πu states (RC–C is 1.253 Å) is the minimum to break the coupling energy of φL and φR, the relevant value is 12.8–15.1 kcal mol−1 (depending on the calculated weight), which is also consistent with the VB results. If the correct decoupled triplet state 3Σ+u is utilized as the electron spin-flip energy, the maximum of the 4th bonding energy is 27.56 kcal mol−1. Hence, the VB 4th bond based on Lewis electron pairing is reasonable. The results of our scheme implement an inherent harmony between VB and MO theories.

Furthermore, we focus on the Effective Bond Order (EBO) based on the one electron density of the ground state in Table 1.29 The EBO of 2σg–2σu is 0.191 and that of 3σg–3σu is 0.190. How to define their bond orders is still a problem. No σ bond or two weak σ bonds? If no σ bond, then it is a classic double bond scheme. If two weak σ bonds, then it is another quadruple bond scheme. Anyway, the EBO results at least demonstrate that the two π bonds contribute most to the bonding energy if the 3P state of the carbon atom is a starting point. That is the reason why the C–C bonding breakage of C2 is smaller than that of ethylene (−172 kcal mol−1). In our simulation, the highest EBO is 2.74 for the 3Σ+u state, because of the occupied number of 0.976e in the 3σg orbitals. Some distribution of 3σg locates between the two carbon atoms. A similar contribution is from the lower occupied number of 1.031e in the 2σu orbitals. Hence, its RC–C value is ∼0.03 Å shorter than the RC–C value of the ground state, and the corresponding ωe value is the highest.

In the end, we would like to say a little more about the nature of the chemical bond in C2 with an ancient Chinese poem about Mountain Lu written by Su Shi: “It's a range viewed in face and peaks from the side. Assuming different shapes viewed from far and wide. Of the Mountain Lu we cannot make out the true face. For we are lost in the heart of the very place”.30 We are shown that the shape and scenery of Mountain Lu are different from different perspectives. Similarly, the understanding of “the most rigorous theory23 for C2 perhaps depends on the various viewpoints of chemists.

Conclusions

In summary, a quadruple bond scheme identical to Prof. Shaik's result from VB theory is achieved, which is related to its RC–C length. Meanwhile, the weak 4th bond or the singlet diradical characteristic of C2 is also easy to be understood, and is related to its instable/reactive nature. Our study conquers the shortcoming of traditional valence MOs. It is worthy of note that C2 must have multi-reference configurations in nature due to having no energy difference among these schemes. The only difference is the dominant contribution in the total CSFs, and how to understand them.

Methodology

Based on the calculations in previous references,21,23 the precision achieved by CASSCF(8,8), which covers the CSFs excited including all the valence electron orbitals, is as good as full CI, since it is commonly recognized that the weights of the inner 1σ2g2u orbitals are always 100% in full CI simulations. The configurations of the C2 electronic states are constructed directly from combinations of natural atomic orbitals in our simulations, because natural orbitals, as a particularly efficient choice, possess the unique advantage of minimizing the mixing effect of the 2s–2p orbitals in carbon and eliminating the diversification of LCAO–MOs in MCSCF simulations.31–34 All the calculations were performed mainly based on the GAUSSIAN 09 program package.35

Acknowledgements

This work was financially supported by the National Nature Science Foundation of China (No. 21203020 and 21131001), National Nature Science Foundation of Jilin Prov. (No. 20150101006JC) and Thirteen Five-Year Sci-tech Research Guideline of the Education Department of Jilin Prov, China.

Notes and references

  1. R. S. Mulliken, Phys. Rev., 1939, 56, 778–781 CrossRef CAS.
  2. R. Hoffmann, Am. Sci., 1995, 83, 309–311 Search PubMed.
  3. W. Weltner and R. J. van Zee, Chem. Rev., 1989, 89, 1713–1747 CrossRef CAS.
  4. M. Martin, J. Photochem. Photobiol., A, 1992, 66, 263–289 CrossRef CAS.
  5. P. S. Skell and J. H. Plonka, J. Am. Chem. Soc., 1970, 92, 5620–5624 CrossRef CAS.
  6. P. S. Skell, K. J. Klabunde and F. A. Fagone, J. Am. Chem. Soc., 1972, 94, 7862–7866 CrossRef CAS.
  7. M. Douay, R. Nietmann and P. F. Bernath, J. Mol. Spectrosc., 1988, 131, 250–260 CrossRef CAS.
  8. R. D. Mead, U. Hefter, P. A. Schulz and W. C. Lineberger, J. Chem. Phys., 1985, 82, 1723–1731 CrossRef.
  9. D. Forney, H. Althaus and J. P. Maier, J. Phys. Chem., 1987, 91, 6458–6461 CrossRef CAS.
  10. M. Rösslein, M. Wyttenbach and J. P. Maier, J. Chem. Phys., 1987, 87, 6770–6772 CrossRef.
  11. J. P. Maier and M. Rösslein, J. Chem. Phys., 1988, 88, 4614–4620 CrossRef CAS.
  12. M. Atoji, J. Chem. Phys., 1961, 35, 1950–1960 CrossRef CAS.
  13. R. Juza, V. Wehle and H.-U. Schuster, Z. Anorg. Allg. Chem., 1967, 352, 252–257 CrossRef CAS.
  14. E. D. Jemmis, B. Pathak, R. B. King and H. F. Schaefer Iii, Chem. Commun., 2006, 2164–2166 RSC.
  15. J. M. Galbraith, E. Blank, S. Shaik and P. C. Hiberty, Chem.–Eur. J., 2000, 6, 2425–2434 CrossRef CAS.
  16. P. Pyykkö, S. Riedel and M. Patzschke, Chem.–Eur. J., 2005, 11, 3511–3520 CrossRef PubMed.
  17. D. R. Huntley, G. Markopoulos, P. M. Donovan, L. T. Scott and R. Hoffmann, Angew. Chem., Int. Ed., 2005, 44, 7549–7553 CrossRef CAS PubMed.
  18. P. Su, J. Wu, J. Gu, W. Wu, S. Shaik and P. C. Hiberty, J. Chem. Theory Comput., 2011, 7, 121–130 CrossRef CAS.
  19. S. Shaik, D. Danovich, W. Wu, P. Su, H. S. Rzepa and P. C. Hiberty, Nat. Chem., 2012, 4, 195–200 CrossRef CAS PubMed.
  20. D. Danovich, P. C. Hiberty, W. Wu, H. S. Rzepa and S. Shaik, Chem.–Eur. J., 2014, 20, 6220–6232 CrossRef CAS PubMed.
  21. G. Frenking and M. Hermann, Angew. Chem., Int. Ed., 2013, 52, 5922–5925 CrossRef CAS PubMed.
  22. W. Wu, J. Gu, J. Song, S. Shaik and P. C. Hiberty, Angew. Chem., Int. Ed., 2009, 48, 1407–1410 CrossRef CAS PubMed.
  23. S. Shaik, H. S. Rzepa and R. Hoffmann, Angew. Chem., Int. Ed., 2013, 52, 3020–3033 CrossRef CAS PubMed.
  24. L. T. Xu and T. H. Dunning, J. Chem. Theory Comput., 2014, 10, 195–201 CrossRef CAS.
  25. J. R. Chelikowsky, N. Troullier, K. Wu and Y. Saad, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 11355–11364 CrossRef CAS.
  26. E. Ramos-Cordoba, P. Salvador and M. Reiher, Chem.–Eur. J., 2013, 19, 15267–15275 CrossRef CAS PubMed.
  27. C. H. Choi, D.-J. Liu, J. W. Evans and M. S. Gordon, J. Am. Chem. Soc., 2002, 124, 8730–8740 CrossRef CAS PubMed.
  28. X. Lu, M. Zhu, X. Wang and Q. Zhang, J. Phys. Chem. B, 2004, 108, 4478–4484 CrossRef CAS.
  29. B. O. Roos, A. C. Borin and L. Gagliardi, Angew. Chem., Int. Ed., 2007, 46, 1469–1472 CrossRef CAS PubMed.
  30. S. Su, Written on the wall at West Forest Temple. This version is translated by Xu Yuanchong.
  31. P. O. Löwdin, Phys. Rev., 1955, 97, 1474–1489 CrossRef.
  32. P. O. Löwdin, Phys. Rev., 1955, 97, 1490–1508 CrossRef.
  33. P. O. Löwdin, Phys. Rev., 1955, 97, 1509–1520 CrossRef.
  34. E. R. Davidson, Rev. Mod. Phys., 1972, 44, 451–464 CrossRef CAS.
  35. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, J. Vreven, T. K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 09w, Revision A.02, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.

Footnote

Ronglin Zhong and Min Zhang contributed equally.

This journal is © The Royal Society of Chemistry 2016