Archive for the ‘Hypervalency’ Category

Chasing ever higher bond orders; the strange case of beryllium.

Monday, February 7th, 2022

Ever since the concept of a shared two-electron bond was conjured by Gilbert N. Lewis in 1916,[1] chemists have been fascinated by the related concept of a bond order (the number of such bonds that two atoms can participate in, however a bond is defined) and pushing it ever higher for pairs of like-atoms. Lewis first showed in 1916[1] how two carbon atoms could share two, four or six electrons to achieve a bond order of up to three. It took quite a few decades for this to be extended to four for carbon (and nitrogen) and that only with some measure of controversy and dispute (for one recent brief summary, see[2]).

For the transition elements over the last forty years or so, bond orders of four, five and even six between like atom pairs have been mooted and many characterised.[3] Moving to the left of the transition elements in the periodic table, this hunt has looked at elements such as beryllium. Eleven years back, I explored here how a Be=Be double bond could be formed, strangely enough as an electronically excited state of the dispersion-bound weak Be2 dimer.[4] This species had a calculated Be-Be distance of 1.78Å, resulting from double excitation from the 2s σ*-antibonding orbital into the degenerate π-bonding orbital above it, giving four electrons in bonding valence orbitals. In 2019, three articles appeared which showed how this bond order might be extended to the lofty heights of three as in Be≡Be[5],[6],[7] for (hypothetical) molecules in their ground electronic state. Here I discuss one example from these articles and compare it to the excited state observations made previously.

A useful starting point is the standard molecular orbital diagram for Be2, illustrating why the ground state singlet actually has a bond order of zero.

The three 2019 suggestions[5],[6],[7] modified this to surround the Be2 core with e.g. six Li atoms, resulting in a stable singlet species with a Be-Be distance (calculated at e.g. the CCSD/Def2-TZVP level) of 1.99Å and exhibiting C2h symmetry. The role of the Li is to polarise and repopulate Be orbitals by delocalization of e.g. a 2c-2e bond in Be2 dimer into a 6c-2e bond in Be2Li6. The reported calculations (as successfully replicated here, FAIR DOI: 10.14469/hpc/10106) show the resulting molecular orbitals for Be2Li6 comprise an (accidentally) degenerate π-pair and a higher energy weak σ-orbital, together forming the proposed triple bond. This of course inverts the normal ordering of such bonds, for which the σ-orbital is lower in energy (more stable) than π-bonds. The form of the σ-orbital also reminds to some extent of the fourth σ-bond in C⩸C.

MOs for Be2Li6

HOMO, σ orbital

-0.158au

HOMO-1, π-pair,

-0.175au

HOMO-2, π-pair

-0.176au

Because the static 2D projections shown in the articles cited above do not always make for easy interpretation, if you click on the orbital thumbnails, you will get dynamic 3D isosurfaces to rotate and inspect. These were generated using the tool at https://www.ch.ic.ac.uk/rzepa/cub2jvxl/

The two lower energy 2s σ-orbitals, which taken together do not apparently contribute to the overall bond order in Be2Li6, are shown below.

Lower energy MOs for Be2Li6
σ -0.235au σ-0.496au

ELF (electron localisation function) integrations for Be2Li6 show each beryllium has two basins in the Be-Be region of about 2.5e each (red arrows) typical of triple bonds and two terminal Li-Be basins of 2.3e.

One aspect arising from my earlier post on the excited state Be=Be double bond relates to the reported calculated Be-Be bond length of 1.99Å and ν 718 cm-1 for ground state Be2Li6. To quote one article[5], “the Be≡Be triple bond in Li6Be2 may also be considered as another example of an ultraweak but ultrashort triple bond.” I had noted earlier that the electronically excited state of the Be2 dimer has a computed bond length of 1.78Å and ν 917 cm-1 for a double bond order, this being significantly shorter than the suggested ultrashort triple bond. We learn from this that the relationship between a bond order and a bond length may not always be linear. In other words, a longer bond may in fact have a higher bond order than a shorter bond between the same two atoms. The same was true as it happens with C⩸C; the mooted quadruple bond had a longer bond length than the triple bond in HC≡CH. That observation was controversial at the time; I suspect a similar phenomenon for Be has become less controversial.

To go back to the Be=Be dimer which started things off and that excited state with one electron in each of the degenerate π-orbitals (actually a triplet state). What would happen if two electrons were to be added, making an excited state of Be22-? Yes indeed, this species (CCSD/Def2-TZVPPD) has a calculated bond length of 1.885Å and ν 766 cm-1. If this di-anion is stabilised with a continuum water field (a milder version of surrounding the dimer with Li atoms), the Be-Be length contracts to 1.74Å, the Be-Be stretch increases to 949 cm-1 and the σ-orbital becomes more stable than the π-orbitals. At the higher CCSD(T)/Def2-TZVPPD/SCRF=water level, the bond length still has the ultrashort value of 1.761Å, which might be assumed as the natural value for Be≡Be, a classical triple bond. From that perspective, the “ultraweak but ultrashort triple bond” predicted for Be2Li6 actually emerges as a relatively long triple bond!

Our final exploration is to add two lithium atoms to Be2 to form the neutral LiBe≡BeLi. This was done in stages (see FAIR DOI 10.14469/hpc/10106), starting with a linear arrangement of atoms which revealed two negative force constants, a C2h shape with one negative force constant and ending with a C2 (chiral!) geometry with no negative force constants. This has a Be≡Be length of 1.705Å (ωB97XD/Def2-TZVPPD/SCRF=water), ν 1129 cm-1, a Wiberg bond index of 2.98 and a Li-Be bond index of 0.0065, indicating an entirely ionic lithium and again a central Be22- unit. As an excited state, it is 49.8 kcal/mol higher than the ground state of Be2Li2.

NBOs for LiBe≡BeLi

HOMO, π-pair,

-0.175au

HOMO, π-pair

-0.176au

HOMO-2, σ orbital

-0.158au

So to conclude, we have seen two different motifs for constructing a model of a Be≡Be triple bond, one recently reported in the literature for a ground state species with six lithium atoms surrounding the Be2 dimer and a simpler one with just two lithiums exhibiting a much shorter Be≡Be bond but which requires electronic excitation to achieve. So these two motifs are not equivalent. But hopefully this exercise shows how playing around with atoms and electrons can achieve very unusual bonding states and elevated bond orders from which one can learn a lot, although with the caveat that one does not always produce molecules capable of facile synthesis!


On a slightly different theme, Cs can be shown to sustain three bonds, albeit all to different atoms. See DOI: 10.6084/m9.figshare.861030 Li≡Li4- can also be calculated as the tetra-anion showing almost identical properties to Be≡Be2- with a Li≡Li triple bond distance of 2.11Å. See DOI: 10.14469/hpc/10122. Replication was necessary because the appropriate wavefunction files for analysis were not included in the supporting information. Only the coordinates were available for interoperation, and due to a quirk in the way Adobe Acrobat works, even those could not be easily transferred by a simple copy/paste operation to create a job input file. See e.g. here or DOI: 10.14469/hpc/10043 for more discussion. All the wavefunction files for this replication are available at the FAIR DOI noted above. The Be-Be distance in catena(dimethylberyllium), a polymer comprising two bridging Me units connecting Be atoms, is only slightly longer at 2.09Å[8] This fascinating transannular Be-Be interaction is one to be explored elsewhere.


The post has DOI: 10.14469/hpc/10125


References

  1. G.N. Lewis, "THE ATOM AND THE MOLECULE.", Journal of the American Chemical Society, vol. 38, pp. 762-785, 1916. https://doi.org/10.1021/ja02261a002
  2. H.S. Rzepa, "Routes involving no free C<sub>2</sub> in a DFT-computed mechanistic model for the reported room-temperature chemical synthesis of C<sub>2</sub>", Physical Chemistry Chemical Physics, vol. 23, pp. 12630-12636, 2021. https://doi.org/10.1039/d1cp02056k
  3. D. Lu, P.P. Chen, T. Kuo, and Y. Tsai, "The MoMo Quintuple Bond as a Ligand to Stabilize Transition‐Metal Complexes", Angewandte Chemie International Edition, vol. 54, pp. 9106-9110, 2015. https://doi.org/10.1002/anie.201504414
  4. P.J. Bruna, and J.S. Wright, "Strongly bound doubly excited states of Be<sub>2</sub>", Canadian Journal of Chemistry, vol. 74, pp. 998-1004, 1996. https://doi.org/10.1139/v96-111
  5. S.S. Rohman, C. Kashyap, S.S. Ullah, A.K. Guha, L.J. Mazumder, and P.K. Sharma, "Ultra‐Weak Metal−Metal Bonding: Is There a Beryllium‐Beryllium Triple Bond?", ChemPhysChem, vol. 20, pp. 516-518, 2019. https://doi.org/10.1002/cphc.201900051
  6. X. Liu, R. Zhong, M. Zhang, S. Wu, Y. Geng, and Z. Su, "BeBe triple bond in Be<sub>2</sub>X<sub>4</sub>Y<sub>2</sub> clusters (X = Li, Na and Y = Li, Na, K) and a perfect classical BeBe triple bond presented in Be<sub>2</sub>Na<sub>4</sub>K<sub>2</sub>", Dalton Transactions, vol. 48, pp. 14590-14594, 2019. https://doi.org/10.1039/c9dt03321a
  7. S.S. Rohman, C. Kashyap, S.S. Ullah, L.J. Mazumder, P.P. Sahu, A. Kalita, S. Reza, P.P. Hazarika, B. Borah, and A.K. Guha, "Revisiting ultra-weak metal-metal bonding", Chemical Physics Letters, vol. 730, pp. 411-415, 2019. https://doi.org/10.1016/j.cplett.2019.06.023
  8. A.I. Snow, and R.E. Rundle, "The structure of dimethylberyllium", Acta Crystallographica, vol. 4, pp. 348-352, 1951. https://doi.org/10.1107/s0365110x51001100

Never mind main group “hypervalency”, what about transition metal “hypervalency”?

Sunday, March 18th, 2018

I have posted often on the chemical phenomenon known as hypervalency, being careful to state that as defined it applies just to “octet excess” in main group elements. But what about the next valence shell, occurring in transition metals and known as the “18-electron rule”? You rarely hear the term hypervalency being applied to such molecules, defined presumably by the 18-electron valence shell count being exceeded. So the following molecule (drawn in three possible valence bond representations) first made in 1992 intrigues.[1]

The molecule comprises six phosphinidene ligands (RP:, R=tert-butyl), the P analogues of nitrenes and arranged around nickel to form an unusual hexagonal planar coordinate species and with three of the t-butyl groups facing up and three down. This arrangement totally obscures the two nickel diaxial positions, preventing any ligand from occupying them. One may even speculate that the dispersion attractions between the two pairs of three t-butyl groups might be unusually stabilising, maybe even on a par with those reported by Schreiner and co-workers for t-butyl substituted triphenylmethanes and noted on this blog.

Nitrenes can be represented coordinating to a metal as RN=M. If the analogy extends to P, the valence bond structure 1 above would result and the six P atoms would contribute 12 electrons to the Ni valence shell. Since the Ni shell is  [Ar].3d8.4s2 adding another 12 electrons would make 22 electrons, thus exceeding the 18-electron rule. In fact it was never suggested as such; in the 1992 analysis of the bonding[1], the authors clearly state “The electronic structure … cannot be described in terms of the 18e rule”. To support this, they draw the structure in form 3 above, which implies a Ni valence shell of 16e, albeit also implying a Ni bond index of ~6.

Time I thought for calculations (wB97XD/Def2-TZVPP). The calculated NBO bond index at Ni is in fact 2.38 and the individual Ni-P bond orders are 0.37. The P bond indices are each 3.24 and the P-P bond orders are 0.88. The final electronic configuration is [Ar]3d9.53.4s0.22.4p0.64.4d0.01.5p0.01 and the natural charge on Ni is -0.39. I show one NBO orbital which illustrates the two-electron-three-centre interaction spanning two P atoms and the Ni and giving rise to the modest Ni-P bond order.

The Ni electronic structure of [Ar].3d8.4s2 normally corresponds to its divalency, so in this sense it is mildly hypervalent. Overall, representation 2 above is perhaps more accurate than 3. The ELF (Electron localisation function) of the electron density is shown below (t-butyl groups represented by a single carbon) with the ELF basins highlighted. Basin 2 is a P lone pair, with an integration close to 2e. Basin 1 is a P-P basin with an integration of 1.89 and finally 3 is indeed a P-Ni basin, but with an integration of only 0.06e. 

So this molecule really is a ring of six P atoms, encapsulating at its centre a lone Ni(II) atom. Rather than being hypervalent, perhaps it is most interesting as a complex in which a metal atom is contained in and perturbed by a dispersion-stabilized sphere of ligands.[2] We need more examples!


The original authors[1] merely stated that electron correlation effects are decisive for the stability. Of course, dispersion attractions are indeed a form of electron correlation! FAIR data doi: 10.14469/hpc/3925

References

  1. R. Ahlrichs, D. Fenske, H. Oesen, and U. Schneider, "Synthesis and Structure of [Ni(P<i>t</i>Bu<sub>6</sub>)] and [Ni<sub>5</sub>(P<i>t</i>Bu)<sub>6</sub>(CO)<sub>5</sub>] and Calculations on the Electronic Structure of [Ni(P<i>t</i>Bu)<sub>6</sub>] and (PR)<sub>6</sub>, R = <i>t</i>Bu,Me", Angewandte Chemie International Edition in English, vol. 31, pp. 323-326, 1992. https://doi.org/10.1002/anie.199203231
  2. https://doi.org/

Never mind main group "hypervalency", what about transition metal "hypervalency"?

Sunday, March 18th, 2018

I have posted often on the chemical phenomenon known as hypervalency, being careful to state that as defined it applies just to “octet excess” in main group elements. But what about the next valence shell, occurring in transition metals and known as the “18-electron rule”? You rarely hear the term hypervalency being applied to such molecules, defined presumably by the 18-electron valence shell count being exceeded. So the following molecule (drawn in three possible valence bond representations) first made in 1992 intrigues.[1]

The molecule comprises six phosphinidene ligands (RP:, R=tert-butyl), the P analogues of nitrenes and arranged around nickel to form an unusual hexagonal planar coordinate species and with three of the t-butyl groups facing up and three down. This arrangement totally obscures the two nickel diaxial positions, preventing any ligand from occupying them. One may even speculate that the dispersion attractions between the two pairs of three t-butyl groups might be unusually stabilising, maybe even on a par with those reported by Schreiner and co-workers for t-butyl substituted triphenylmethanes and noted on this blog.

Nitrenes can be represented coordinating to a metal as RN=M. If the analogy extends to P, the valence bond structure 1 above would result and the six P atoms would contribute 12 electrons to the Ni valence shell. Since the Ni shell is  [Ar].3d8.4s2 adding another 12 electrons would make 22 electrons, thus exceeding the 18-electron rule. In fact it was never suggested as such; in the 1992 analysis of the bonding[1], the authors clearly state “The electronic structure … cannot be described in terms of the 18e rule”. To support this, they draw the structure in form 3 above, which implies a Ni valence shell of 16e, albeit also implying a Ni bond index of ~6.

Time I thought for calculations (wB97XD/Def2-TZVPP). The calculated NBO bond index at Ni is in fact 2.38 and the individual Ni-P bond orders are 0.37. The P bond indices are each 3.24 and the P-P bond orders are 0.88. The final electronic configuration is [Ar]3d9.53.4s0.22.4p0.64.4d0.01.5p0.01 and the natural charge on Ni is -0.39. I show one NBO orbital which illustrates the two-electron-three-centre interaction spanning two P atoms and the Ni and giving rise to the modest Ni-P bond order.

The Ni electronic structure of [Ar].3d8.4s2 normally corresponds to its divalency, so in this sense it is mildly hypervalent. Overall, representation 2 above is perhaps more accurate than 3. The ELF (Electron localisation function) of the electron density is shown below (t-butyl groups represented by a single carbon) with the ELF basins highlighted. Basin 2 is a P lone pair, with an integration close to 2e. Basin 1 is a P-P basin with an integration of 1.89 and finally 3 is indeed a P-Ni basin, but with an integration of only 0.06e. 

So this molecule really is a ring of six P atoms, encapsulating at its centre a lone Ni(II) atom. Rather than being hypervalent, perhaps it is most interesting as a complex in which a metal atom is contained in and perturbed by a dispersion-stabilized sphere of ligands.[2] We need more examples!


The original authors[1] merely stated that electron correlation effects are decisive for the stability. Of course, dispersion attractions are indeed a form of electron correlation! FAIR data doi: 10.14469/hpc/3925

References

  1. R. Ahlrichs, D. Fenske, H. Oesen, and U. Schneider, "Synthesis and Structure of [Ni(P<i>t</i>Bu<sub>6</sub>)] and [Ni<sub>5</sub>(P<i>t</i>Bu)<sub>6</sub>(CO)<sub>5</sub>] and Calculations on the Electronic Structure of [Ni(P<i>t</i>Bu)<sub>6</sub>] and (PR)<sub>6</sub>, R = <i>t</i>Bu,Me", Angewandte Chemie International Edition in English, vol. 31, pp. 323-326, 1992. https://doi.org/10.1002/anie.199203231
  2. https://doi.org/

Hypervalent or not? A fluxional triselenide.

Saturday, February 24th, 2018

Another post inspired by a comment on an earlier one; I had been discussing compounds of the type I.In (n=4,6) as possible candidates for hypervalency. The comment suggests the below as a similar analogue, deriving from observations made in 1989.[1]

This compound was investigated using 77Se NMR, with the following conclusions:

  1. The compound is fluxional, with the lines at room temperature broadened compared to those at -50°C.
  2. At -50°C the peaks are sharp enough to discern 1JSe-Se couplings, with multiplicities and integrations that suggest a central Se is surrounded by four equivalent further Se atoms, with shifts of 655.1 and 251.2 ppm.
  3. The magnitude of this 1JSe-Se coupling (391 Hz) leads to the suggestion of a considerable contribution of a resonance form with Se=Se bonds (structure 2 above).
  4. This was supported by 2J13C-77Se couplings which also imply a symmetrically coordinated central  Se.
  5. Thus the two resonance forms 1 or 2 above were suggested as the predominant form at -50°C, with an increasing incursion of the open chain isomer 3 at higher temperatures giving rise to the observed fluxional dynamic behaviour.
  6. One may surmise from these results that the central Se is certainly hypercoordinated and by the classical interpretations hypervalent.

Here are some calculations (R=H), at the ωB97XD/Def2-TZVPP/SCRF=chloroform level.‡ In red are the calculated Wiberg Se-Se bond orders, which give little indication of any Se=Se double bond character. 

The calculated 77Se shifts are shown in magenta, with the observed values being 655 and 255 ppm. The match is not good, the errors were 120 and 20.5 ppm.  However calculated shifts for elements adjacent to e.g. Se or Br etc suffer from relativistic effects such as spin orbit coupling.[2] Thus the shift for the central Se, surrounded by four other Se atoms is likely to have a significant error, but the error for the four other Se atoms should be less. The reverse is true.

However, all the calculations of this species (up to Def2-TZVPPD basis set) showed this symmetric form of D2h symmetry to actually be a transition state, as per below.

There is a minimum with the structure below in which one pair of Se-Se lengths are longer than the other pair and for which the free energy is 6.5 kcal/mol lower. The Wiberg bond orders for the two sets of Se-Se bonds are now 0.16 and 0.86, which very much corresponds to structure 3 above.

Assuming that this compound is fluxional even at -50°C, the average of the pairs of Se atoms gives calculated shifts of 667 ppm (655 obs) whilst the central Se is 204.6 ppm (251 obs). The latter, influenced by two especially short Se-Se distances, is likely to have a very large spin-orbit coupling error, whilst for the former the error will be smaller (13C shifts adjacent to one Br typically have induced calculated errors of about 14 ppm[2]).

At this point I searched the Cambridge structure database for Se coordinated by four other Se atoms. A close analogue[3] has the structure shown below, in which pairs of Se-Se interactions have unequal bond lengths, the shorter being ~2.45Å. This matches the calculation above reasonably well.

Reconciling these various observations, we might assume that even at -50°C the fluxional behaviour has not been frozen out. Given that the fluxional barrier is only 6.5 kcal/mol, it is unlikely that the spectrum could be measured at a sufficiently low temperature to reveal not two sets of Se signals in the ratio 4:1 but three in the ratio 2:2:1. The spin-spin couplings reported presumably are a result of averaging a genuine 1JSe-Se coupling with a through space coupling.

So it appears that the analysis of the 77Se NMR reported in this article [1] may not be quite what it seems. A better interpretation is that structure 3 is the most realistic. This means no hypercoordination for the Se, never mind hypervalence!


FAIR data at DOI: 10.14469/hpc/3724. The original reference, Me2Se was incorrectly calculated without solvation by chloroform. The values shown here are now corrected from those shown in the original post.

References

  1. Y. Mazaki, and K. Kobayashi, "Structure and intramolecular dynamics of bis(diisobutylselenocarbamoyl) triselenide as identified in solution by the 77Se-NMR spectroscopy", Tetrahedron Letters, vol. 30, pp. 2813-2816, 1989. https://doi.org/10.1016/s0040-4039(00)99132-9
  2. D.C. Braddock, and H.S. Rzepa, "Structural Reassignment of Obtusallenes V, VI, and VII by GIAO-Based Density Functional Prediction", Journal of Natural Products, vol. 71, pp. 728-730, 2008. https://doi.org/10.1021/np0705918
  3. R.O. Gould, C.L. Jones, W.J. Savage, and T.A. Stephenson, "Crystal and molecular structure of bis(NN-diethyldiselenocarbamato)-selenium(II)", Journal of the Chemical Society, Dalton Transactions, pp. 908, 1976. https://doi.org/10.1039/dt9760000908

Hypervalent Helium – not!

Friday, February 16th, 2018

Last year, this article[1] attracted a lot of attention as the first example of molecular helium in the form of Na2He. In fact, the helium in this species has a calculated bond index of only 0.15 and it is better classified as a sodium electride with the ionisation induced by pressure and the presence of helium atoms. The helium is neither valent, nor indeed hypervalent (the meanings are in fact equivalent for this element). In a separate blog posted in 2013, I noted a cobalt carbonyl complex containing a hexacoordinate hydrogen in the form of hydride, H. A comment appended to this blog insightfully asked about the isoelectronic complex containing He instead of H. Here, rather belatedly, I respond to this comment!

The complex [HCo6(CO)15] has a calculated bond index at the hydrogen of 0.988 and a calculated NMR chemical shift of 21.6 ppm (ωB97XD/Def2-TZVPPD calculation) compared to a measured value of 23.2 ppm. Despite being six-coordinate, the hydride has a bond index that does not exceed one (it is not hypervalent).

So here is the neutral helium analogue. The He bond index emerges as 0.71 at the geometry of the hydride complex. Compare this with the bond index of 0.15 calculated for Na2He and it would be fair to say that at this geometry, the helium in [HeCo6(CO)15] would have a greater claim to be a molecular compound. Back in 2010, extrapolating from a series of posts here, I had speculated[2] about other molecular species of He, including the di-cation below. This has a He bond index of 0.54, rather less than that in [HeCo6(CO)15] but much more than in Na2He. It is also vibrationally stable.

But now, [HeCo6(CO)15] goes “pear-shaped” (why do pears have such a bad press?). I started a process of optimizing the geometry of this complex (ωB97Xd/Def2-TZVPPD). Slowly, the He started to creep out of the centre of the complex and emerge from the cavity. After about 100 steps it reached the geometry shown below, at which point the Wiberg bond index has dropped to 0.62 and still going down. I think it might take a few more steps to be completely expelled, but I have stopped the geometry optimisation at this stage.

So helium appears not to be valent in [HeCo6(CO)15]. However, I have yet to try Ne, which is both larger and softer. I will post results here.


All data at 10.14469/hpc/3587.

References

  1. X. Dong, A.R. Oganov, A.F. Goncharov, E. Stavrou, S. Lobanov, G. Saleh, G. Qian, Q. Zhu, C. Gatti, V.L. Deringer, R. Dronskowski, X. Zhou, V.B. Prakapenka, Z. Konôpková, I.A. Popov, A.I. Boldyrev, and H. Wang, "A stable compound of helium and sodium at high pressure", Nature Chemistry, vol. 9, pp. 440-445, 2017. https://doi.org/10.1038/nchem.2716
  2. H.S. Rzepa, "The rational design of helium bonds", Nature Chemistry, vol. 2, pp. 390-393, 2010. https://doi.org/10.1038/nchem.596

Hypervalent hydrogen?

Saturday, January 13th, 2018

I discussed the molecule the molecule CH3F2- a while back. It was a very rare computed example of a system where the added two electrons populate the higher valence shells known as Rydberg orbitals as an alternative to populating the C-F antibonding σ-orbital to produce CH3 and F. The net result was the creation of a weak C-F “hyperbond”, in which the C-F region has an inner conventional bond, with an outer “sheath” encircling the first bond. But this system very easily dissociates to CH3 and F and is hardly a viable candidate for experimental detection.  In an effort to “tune” this effect to see if a better candidate for such detection might be found, I tried CMe3F2-. Here is its story.

The calculation is at the ωB97XD/Def2-TZVPPD/SCRF=water level (water is here used as an approximate model for a condensed environment, helping to bind the two added electrons).

  1. An NBO (Natural Bond orbital) analysis reveals a total Rydberg orbital population of 1.186e and the following bond indices; F 0.853, C 3.977, C(methyl) 4.051, H(*3) 1.332. The latter corresponds to the three methyl hydrogens aligned antiperiplanar to the C-F bond.
  2. To put this value into context, the hydrogen in the FHF anion has an NBO H bond index of 0.724, and the bridging hydrogens in diborane only have a value of 0.988. Even the hexa-coordinate hydride system [Co6H(CO)15] discussed in an earlier blog  has an H bond index of just 0.86. Actually, coordination of six or even higher for hydrogen is no longer rare; some 28 crystal structures of the type HM6 (M=metal) are known (it would be useful to find out if any of the other 27 such structures might have a hydrogen bond index >1).
  3. Next, the ELF analysis (Electron localisation function), analysed firstly using the excellent MultiWFN program.[1]

    This reveals an attractor basin integrating to 1.663e and located along the axis of the F-C bond and extended into the region of the three antiperiplanar methyl hydrogens. The C-F bond itself only supports a basin of 0.729e, typical of the fairly ionic C-F bond. The covalent C-Me bonds are also pretty normal, as are the other hydrogens.
  4. I also show ELF analysis using the alternative TopMod program[2]; the numerical values on this diagram are the calculated bond lengths in Å. The basin integrations are very similar to those obtained using MultiWFN.

    The Wiberg bond orders of the three H…H regions shown connected by dashed lines above are 0.154, which contributes to the bond index of >1 at these three hydrogens.
  5. The predicted 1H chemical shift of these three “hypervalent” hydrogens is +3.0 ppm, whilst the other six methyl hydrogens are at -0.87ppm.

So changing CH3F2- to CMe3F2- has dramatically changed the bonding picture that emerges, rather than a fine-tuning. The C-F is no longer a “hyperbond”, although the Rydberg occupancy of 1.186e remains unusually large. Most of the additional electrons have fled the torus surrounding the C-F bond and relocated to the exo-region of that bond where they now influence the three antiperiplanar methyl hydrogens. A two-electron-three-centre interaction if you like, but with the electron basin occupying a tetrahedral vertex rather than the triatom centroid.

I end with a challenge. Is it possible to find “real” molecules containing hydrogen where the formal bond index for at least one hydrogen exceeds 1.0 significantly, thus making it hypervalent? 


The calculations are all collected at FAIR doi; 10.14469/hpc/3372.

References

  1. T. Lu, and F. Chen, "Multiwfn: A multifunctional wavefunction analyzer", Journal of Computational Chemistry, vol. 33, pp. 580-592, 2011. https://doi.org/10.1002/jcc.22885
  2. S. Noury, X. Krokidis, F. Fuster, and B. Silvi, "Computational tools for the electron localization function topological analysis", Computers & Chemistry, vol. 23, pp. 597-604, 1999. https://doi.org/10.1016/s0097-8485(99)00039-x

Are diazomethanes hypervalent molecules? An attempt into more insight by more “tuning” with substituents.

Tuesday, December 26th, 2017

Recollect the suggestion that diazomethane has hypervalent character[1]. When I looked into this, I came to the conclusion that it probably was mildly hypervalent, but on carbon and not nitrogen. Here I try some variations with substituents to see what light if any this casts.

I have expanded the resonance forms of diazomethane by one structure from those shown in the previous two posts (a form by the way not considered in the original article[1]) to include a nitrene. This takes us back to an earlier suggestion on this blog that HC≡S≡CH is not a stable species but a higher order saddle point which distorts down to a bis-carbene, together with the suggestion that hypervalent triple bonds have the option of converting four of the six electrons into two carbene lone pairs, replacing the triple bond with a single bond. This in turn harks back to G. N. Lewis’ 101 year old idea for acetylene itself!

To explore this mode, I start by replacing the terminal ≡N in diazomethane with a ≡C-Me group, which cannot absorb electrons into lone-pairs in the manner that nitrogen can. A ωB97XD/Def2-TZVPP calculation reveals that the linear form is a transition state for interconversion into a carbene. The IRC for the process (below) shows this carbene is ~10 kcal/mol lower than the linear “hypervalent” form. 

NBO analysis of this transition state reveals a similar orbital pattern to diazomethane itself, including a non-bonding orbital on the H2C carbon. The Wiberg carbon bond indices are 3.6764 and N 3.6454 and the bond orders C=N 1.1390 and N=CMe 1.6192.

ELF analysis of this transition state reveals the presence of two non-bonding pairs on the carbon atoms either side of the nitrogen but unshared with it, with populations of 1.19e and 1.37e (DFT). That nitrogen really does not like excess electrons! The four atoms C,N,C,C have ELF valence basins totalling 8.00, 6.94, 7.69 and 7.92e (DFT) or 8.07, 7.07 and 7.61e (CASSCF), suggesting that unlike diazomethane itself, the octet-excess induced hypervalence on carbon is slightly decreased.

Pumping even more electrons in by replacing the ≡C-Me group with ≡C-NH2 does not increase any hypervalence, but does induce more electrons to reside in “lone pairs”. Of the four atoms along the chain, three have “lone pairs” associated with them, a total of 4.83e that do not contribute to bonds (valence).

An electron withdrawing ≡C-CN group replacing the ≡C-NH2 reverses the effect of the latter, but this linear species is still a transition state for carbon isomerisation:

Finally, combining all we have learnt by adding in nitro groups on the first carbon. This is no longer a transition state but now a stable species; the sum of the ELF basin integrations around the carbon on the left reaches 8.95e, slightly higher than the dinitro-diazomethane discussed in the previous post. The numerical Wiberg atom bond indices are C 3.8713, N 3.6898, C 3.8503, C 3.9958 and N 3.0288 for the atoms along the chain, with the first nitrogen the “least-valent”.

So we see that “hypervalence”, or at least “octet-excess”, which is not exactly the same as hypervalence since it includes contributions from non-bonding electrons, is balanced on a knife-edge. Trying to increase the octet-excess by pumping electrons in turns the system into a transition state for carbene formation. Octet-excess is seen as a metastable property, to be relieved by geometric distortions where possible or localization of electrons into non-bonding lone pairs. And I remind yet again that no evidence has manifested in calculations of the molecules above that the central nitrogen of these diazomethane-like systems has any propensity for octet or valence-excess as implied by the formula C=N≡X.[1]


FAIR data for all calculations is available at DOI: 10.14469/hpc/3476

References

  1. M.C. Durrant, "A quantitative definition of hypervalency", Chemical Science, vol. 6, pp. 6614-6623, 2015. https://doi.org/10.1039/c5sc02076j

Are diazomethanes hypervalent molecules? An attempt into more insight by more "tuning" with substituents.

Tuesday, December 26th, 2017

Recollect the suggestion that diazomethane has hypervalent character[1]. When I looked into this, I came to the conclusion that it probably was mildly hypervalent, but on carbon and not nitrogen. Here I try some variations with substituents to see what light if any this casts.

I have expanded the resonance forms of diazomethane by one structure from those shown in the previous two posts (a form by the way not considered in the original article[1]) to include a nitrene. This takes us back to an earlier suggestion on this blog that HC≡S≡CH is not a stable species but a higher order saddle point which distorts down to a bis-carbene, together with the suggestion that hypervalent triple bonds have the option of converting four of the six electrons into two carbene lone pairs, replacing the triple bond with a single bond. This in turn harks back to G. N. Lewis’ 101 year old idea for acetylene itself!

To explore this mode, I start by replacing the terminal ≡N in diazomethane with a ≡C-Me group, which cannot absorb electrons into lone-pairs in the manner that nitrogen can. A ωB97XD/Def2-TZVPP calculation reveals that the linear form is a transition state for interconversion into a carbene. The IRC for the process (below) shows this carbene is ~10 kcal/mol lower than the linear “hypervalent” form. 

NBO analysis of this transition state reveals a similar orbital pattern to diazomethane itself, including a non-bonding orbital on the H2C carbon. The Wiberg carbon bond indices are 3.6764 and N 3.6454 and the bond orders C=N 1.1390 and N=CMe 1.6192.

ELF analysis of this transition state reveals the presence of two non-bonding pairs on the carbon atoms either side of the nitrogen but unshared with it, with populations of 1.19e and 1.37e (DFT). That nitrogen really does not like excess electrons! The four atoms C,N,C,C have ELF valence basins totalling 8.00, 6.94, 7.69 and 7.92e (DFT) or 8.07, 7.07 and 7.61e (CASSCF), suggesting that unlike diazomethane itself, the octet-excess induced hypervalence on carbon is slightly decreased.

Pumping even more electrons in by replacing the ≡C-Me group with ≡C-NH2 does not increase any hypervalence, but does induce more electrons to reside in “lone pairs”. Of the four atoms along the chain, three have “lone pairs” associated with them, a total of 4.83e that do not contribute to bonds (valence).

An electron withdrawing ≡C-CN group replacing the ≡C-NH2 reverses the effect of the latter, but this linear species is still a transition state for carbon isomerisation:

Finally, combining all we have learnt by adding in nitro groups on the first carbon. This is no longer a transition state but now a stable species; the sum of the ELF basin integrations around the carbon on the left reaches 8.95e, slightly higher than the dinitro-diazomethane discussed in the previous post. The numerical Wiberg atom bond indices are C 3.8713, N 3.6898, C 3.8503, C 3.9958 and N 3.0288 for the atoms along the chain, with the first nitrogen the “least-valent”.

So we see that “hypervalence”, or at least “octet-excess”, which is not exactly the same as hypervalence since it includes contributions from non-bonding electrons, is balanced on a knife-edge. Trying to increase the octet-excess by pumping electrons in turns the system into a transition state for carbene formation. Octet-excess is seen as a metastable property, to be relieved by geometric distortions where possible or localization of electrons into non-bonding lone pairs. And I remind yet again that no evidence has manifested in calculations of the molecules above that the central nitrogen of these diazomethane-like systems has any propensity for octet or valence-excess as implied by the formula C=N≡X.[1]


FAIR data for all calculations is available at DOI: 10.14469/hpc/3476

References

  1. M.C. Durrant, "A quantitative definition of hypervalency", Chemical Science, vol. 6, pp. 6614-6623, 2015. https://doi.org/10.1039/c5sc02076j

Can any hypervalence in diazomethanes be amplified?

Saturday, December 23rd, 2017

In the previous post, I referred to a recently published review on hypervalency[1] which introduced a very simple way (the valence electron equivalent γ) of quantifying the effect. Diazomethane was cited as one example of a small molecule exhibiting hypervalency (on nitrogen) by this measure. Here I explore the effect of substituting diazomethane with cyano and nitro groups.

Firstly, dicyanodiazomethane. NBO analysis reveals the following atom bond indices; C, 3.810; N 3.834; N 2.971. Compare these values to diazomethane itself, C, 3.716; N 3.802; N 2.907 and you can see that the carbon bond index has increased slightly. The ELF basin integrations (below) which also take into account the “lone pair” on carbon are: C, 8.55, N, 6.65, N, 7.52 (DFT), again compared with diazomethane as C, 8.16; N, 6.59; N, 7.52. The CASSCF(14,14) result is very similar.

So the “γ(C)” has increased from 8.2 to 8.55. Next, dinitrodiazomethane;

The NBO bond indices are C, 3.8203; N 3.8255; N 2.9802 and ELF integrations C, 8.82, N, 6.68, N, 7.49 (DFT).

So “γ(C)”  increases along the series 8.16 → 8.55 → 8.82, whereas “γ(N)” changes as 6.59 → 6.65 → 6.68, a smaller effect. Whilst 8.82 is still some way off the value of γ(N)=10 quoted[1] for diazomethane, dinitrodiazomethane is still a pretty good candidate for hypervalent carbon. The question now is whether even larger values of “γ(C)” can be identified in other molecules. 


FAIR data for all calculations is available at DOI: 10.14469/hpc/3476. The quotes in “γ(C)” indicate it is calculated here using ELF integrations rather than charge maps.

References

  1. M.C. Durrant, "A quantitative definition of hypervalency", Chemical Science, vol. 6, pp. 6614-6623, 2015. https://doi.org/10.1039/c5sc02076j

Are diazomethanes hypervalent molecules? Probably, but in an unexpected way!

Saturday, December 23rd, 2017

A recently published review on hypervalency[1] introduced a very simple way of quantifying the effect. One of the molecules which was suggested to be hypervalent using this method was diazomethane. Here I take a closer look.

The new method is called the valence electron equivalent γ. It is defined as “the formal shared electron count at a given atom, obtained by any combination of valid ionic and covalent resonance forms that reproduces the observed charge distribution”. These atom charge maps can be obtained from various kinds of quantum mechanical calculation; the one adopted in the review was Bader’s QTAIM analysis.

Three resonance forms used to estimate γ for the atoms in diazomethane are shown below. According to Durrant[1], “if γ(X) > 8, neither form of the octet rule is obeyed and the atom is hypervalent”. His procedure gives γ(N) = 10 (table 3 and also structure 27a[1]), suggesting that the third resonance form shown below is the most appropriate.

As a result, the nitrogen is to be considered hypervalent, with five formal covalent bonds and hence a ten shared-electron valence shell. If such hypervalence is to be considered as more than just a convenient representation and to have a deeper underlying foundation, similar conclusions should ideally emerge from other ways of analysing the wavefunctions for such species. Here I try the NBO (Natural Bond Orbital) and ELF (Electron localization Function) analyses of two types of wavefunction using both DFT (density functional theory) ωB97XD/Def2-TZVPP and multi-reference CASSCF(12,12)/Def2-TZVPP Hamiltonians.

Firstly, the NBO method, which localizes electron pairs. There are four bonding NBOs associated with the central nitrogen and a modest contribution from a fifth, the terminal nitrogen lone pair (bottom).

The carbon has four NBOs comprising one “non-bonding lone pair”, two C-Hs, a C-N and one partial C-N (bottom). This partitioning can be quantified using Wiberg atom bond indices: C, 3.716; N 3.802; N 2.907 (the C-N and N=N bond orders are 1.425 and 2.368). Note that the “non-bonding carbon lone pair” would not contribute to the C bond index, which is already 3.72. Is numerical evidence perhaps emerging that the carbon may exceed the octet in its valence shell? In contrast, the central nitrogen, with a bond index of 3.802, but not having any associated “non-bonding lone pair”, does not look to exceeding the octet. It seems well short of achieving γ(N) = 10, equivalent to a bond index of 5 as shown in the resonance form above.

Now for ELF, which is derived from the electron density and the basin locations from its kinetic energy density. The values are the integrations for the individual ELF basins, of which 0.498*2 = 0.996 is the monosynaptic basin for the carbon lone pair noted above. The total for this carbon comes to 8.16, the adjacent nitrogen 6.59 and the terminal nitrogen 7.52e. 

The CASSCF(12,12) values are respectively are 8.18, 6.50 and 7.50e, which are very similar. So these two types of electron partitioning show no evidence that the central nitrogen has γ(N) = 10; it’s actually ~6.5, which is a long way short of 10! However, γ(C) = ~8.2, which does exceed the octet, if only very modestly. Perhaps this can best be summarised with the following representation based on the bond orders, which indicates three shared covalent bonds to carbon, a partial (dashed) C-N interaction and a lone pair implied by the minus sign, the total of which exceeds γ(C) = ~8. It all boils down really to whether that “lone pair” of electrons on the carbon is truly a non-bonding electron pair or whether it should be considered a fully covalent pair associated also with the nitrogen. Certainly, there appears to be no evidence from NBO or ELF that this is the case.

The next logical question to ask is whether the effect can be “optimised” such that γ(C) > 8 or even >> 8. I will address this in the next post.


FAIR data for all calculations is available at DOI: 10.14469/hpc/3476

References

  1. M.C. Durrant, "A quantitative definition of hypervalency", Chemical Science, vol. 6, pp. 6614-6623, 2015. https://doi.org/10.1039/c5sc02076j