Show simple item record

Molecular Structure of XeF6. II. Internal Motion and Mean Geometry Deduced by Electron Diffraction

dc.contributor.authorBartell, Lawrence S.en_US
dc.contributor.authorGavin, R. M.en_US
dc.date.accessioned2010-05-06T22:28:55Z
dc.date.available2010-05-06T22:28:55Z
dc.date.issued1968-03-15en_US
dc.identifier.citationBartell, L. S.; Gavin, R. M. (1968). "Molecular Structure of XeF6. II. Internal Motion and Mean Geometry Deduced by Electron Diffraction." The Journal of Chemical Physics 48(6): 2466-2483. <http://hdl.handle.net/2027.42/70641>en_US
dc.identifier.urihttps://hdl.handle.net/2027.42/70641
dc.description.abstractThe distribution of internuclear distances in gaseous XeF6 exhibits unusually diffuse XeF6 bonded and F–F geminal nonbonded peaks, the latter of which is severely skewed. The distribution proves the molecule cannot be a regular octahedron vibrating in independent normal modes. The instantaneous molecular configurations encountered by the incident electrons are predominantly in the broad vicinity of C3υC3υ structures conveniently described as distorted octahedra in which the xenon lone pair avoids the bonding pairs. In these distorted structures the XeF bond lengths are distributed over a range of approximately 0.08 Å with the longer bonds tending to be those adjacent to the avoided region of the coordination sphere. Fluorines suffer angular displacements from octahedral sites which range up to 5° or 10° in the vicinity of the avoided region.Alternative interpretations of the diffraction data are developed in detail, ranging from models of statically deformed molecules to those of dynamically inverting molecules. In all cases it is necessary to assume that t1ut1u bending amplitudes are enormous and correlated in a certain way with substantial t2gt2g deformations. Notwith‐standing the small fraction of time that XeF. spends near OhOh symmetry, it is possible to construct a molecular potential‐energy function more or less compatiable with the diffraction data in which the minimum energy occurs at OhOh symmerty. The most notable feature of this model is the almost vanishing restoring force for small t1ut1u bending distortions. Indeed, the mean curvature of the potential surface for this model corresponds to a υ4υ4 force constant F44F44 of 10−2 mdyn/Å or less. Various rapidly inverting non‐OhOh structures embodying particular combinations of t2gt2g and t1ut1u deformations from OhOh symmetry give slightly better radial distribution functions, however. In the region of molecular configuration where the gas molecules spend most of their time, the form of the potential‐energy function required to represent the data does not distinguish between a Jahn–Teller first‐order term or a cubic V445V445 term as the agent responsible for introducing the t2gt2g deformation. The Jahn–Teller term is consistent with Goodman's interpretation of the molecule. On the other hand, the cubic term is found to be exactly analogous to that for other molecules with stereochemically active lone pairs (e.g., SF4, ClF3). Therefore, the question as to why the XeF6 molecule is distorted remains open. The reported absence of any observable gas‐phase paramagnetism weighs against the Jahn–Teller interpretation.The qualitative success but quantitative failure of the valence‐shell–electron‐pair‐repulsion theory is discussed and the relevance of the “pseudo‐Jahn–Teller” formalism of Longuet‐Higgins et al. is pointed out. Brief comparisons are made with isoelectronic ions.en_US
dc.format.extent3102 bytes
dc.format.extent1546783 bytes
dc.format.mimetypetext/plain
dc.format.mimetypeapplication/pdf
dc.publisherThe American Institute of Physicsen_US
dc.rights© The American Institute of Physicsen_US
dc.titleMolecular Structure of XeF6. II. Internal Motion and Mean Geometry Deduced by Electron Diffractionen_US
dc.typeArticleen_US
dc.subject.hlbsecondlevelPhysicsen_US
dc.subject.hlbtoplevelScienceen_US
dc.description.peerreviewedPeer Revieweden_US
dc.contributor.affiliationumDepartment of Chemistry, University of Michigan, Ann Arbor, Michiganen_US
dc.description.bitstreamurlhttp://deepblue.lib.umich.edu/bitstream/2027.42/70641/2/JCPSA6-48-6-2466-1.pdf
dc.identifier.doi10.1063/1.1669471en_US
dc.identifier.sourceThe Journal of Chemical Physicsen_US
dc.identifier.citedreferenceSee, for example, the discussions in Noble Gas Compounds, H. H. Hyman, Ed. (University of Chicago Press, Chicago, Ill., 1963), and the references therein.en_US
dc.identifier.citedreferenceJ. G. Malm, H. Selig, J. Jortner, and S. A. Rice, Chem. Rev. 65, 199 (1965).en_US
dc.identifier.citedreferenceH. H. Claassen, The Noble Gases (D. C. Heath and Co., Boston, Mass., 1966).en_US
dc.identifier.citedreferenceB. Weinstock, Chem. Eng. News 42, 86 (1964).en_US
dc.identifier.citedreferenceN. V. Sidgwick and H. M. Powell, Proc. Roy. Soc. (London) A176, 153 (1940).en_US
dc.identifier.citedreferenceR. J. Gillespie and R. S. Nyholm, Quart. Rev. (London) 11, 339 (1957).en_US
dc.identifier.citedreferenceR. J. Gillespie, J. Chem. Educ. 40, 295 (1963).en_US
dc.identifier.citedreferenceR. M. Gavin, Jr., and L. S. Bartell, J. Chem. Phys. 48, 2460 (1968) (preceding article).en_US
dc.identifier.citedreferenceIn Analysis I Hartree‐Fock x‐ray elastic scattering factors and Heisenberg‐Bewilogua inelastic scattering factors were employed. Corrections for the failure of the Born approximation were made only through the use of the Thomas‐Fermi phase shifts of Hoerni and Ibers rescaled in effective atomic number to fit the experimental data. In Analysis II elastic scattering was based on the new partial wave calculations of Cox and Bonham (H. L. Cox, Doctoral dissertation, Indiana University, 1967). Hartree‐Fock inelastic scattering factors for F [C. Tavard, D. Nicholas, and M. Roualt, J. Chim. Phys. 64, 540 (1967)] and for Xe [extrapolated from the iodine factors of R. F. Pohler and H. P. Hansen, J. Chem. Phys. 42, 2347 (1965)] were used.en_US
dc.identifier.citedreferenceG. Goodman, Bull. Am. Phys. Soc. 12, 296 (1967).en_US
dc.identifier.citedreferenceE. Meisingseth and S. T. Cyvin, Acta Chem. Scand. 16, 2452 (1963).en_US
dc.identifier.citedreferenceH. M. Seip and R. Stoelevik, Acta Chem. Scand. 20, 1535 (1966).en_US
dc.identifier.citedreferenceR. J. Gillespie, Ref. 1, p. 333.en_US
dc.identifier.citedreferenceR. J. Gillespie, Alfred Werner Centennial Symposium, American Chemical Society Meeting, New York, September 1966.en_US
dc.identifier.citedreferenceH. B. Thompson and L. S. Bartell, Trans. Am. Cryst. Assoc. 2, 190 (1966); H. B. Thompson, W. Adams, L. Winstrom, and L. S. Bartell (unpublished).en_US
dc.identifier.citedreferenceE. J. Jacob, H. B. Thompson, and L. S. Bartell, J. Chem. Phys. 47, 3736 (1967).en_US
dc.identifier.citedreferenceD. H. Templeton, A. Zalkin, J. D. Forrester, and S. M. Williamson, J. Am. Chem. Soc. 85, 242 (1963); J. A. Ibers and W. C. Hamilton, Science 139, 106 (1963); J. H. Burns, P. A. Agron, and H. A. Levy, 139, 1208 (1963); R. K. Bohn, K. Katada, J. V. Martinez, and S. H. Bauer, Ref. 1, p. 238.en_US
dc.identifier.citedreferenceS. Siegel and E. Gebert, J. Am. Chem. Soc. 85, 240 (1963); H. A. Levy and P. A. Agron, 85, 241 (1963).en_US
dc.identifier.citedreferenceG. Nagarajan (private communication).en_US
dc.identifier.citedreferenceG. Nagarajan, Acta Phys. Austriaca 18, 11 (1964). Note, however, that the force constants were based on an incorrect assignment of the eueu bending mode.en_US
dc.identifier.citedreferenceM. Kimura, V. Schomaker, D. Smith, and B. Weinstock (unpublished).en_US
dc.identifier.citedreferenceH. M. Seip, Acta Chem. Scand. 19, 1955 (1965); H. M. Seip and R. Stoelevik, 20, 1535 (1966); H. M. Seip and R. Seip, 20, 2698 (1966).en_US
dc.identifier.citedreferenceE. Meisingseth and S. J. Cyvin, Acta Chem. Scand. 16, 2452 (1962); M. Kimura and K. Kimura, J. Mol. Spectry. 11, 368 (1963).en_US
dc.identifier.citedreferenceThe anomaly may be an artifact of the “foot” on the leading edge of the XeF peak.en_US
dc.identifier.citedreferenceW. E. Falconer, A. Büchler, J. L. Stauffer, and W. Klemperer, “Molecular Structure of XeF6XeF6 and IF7,IF7,” J. Chem. Phys. (to be published).en_US
dc.identifier.citedreferenceD. F. Smith, Ref. 1, p. 295.en_US
dc.identifier.citedreferenceB. Weinstock, E. E. Weaver, and C. P. Knop, Inorg. Chem. 5, 2189 (1966); B. Weinstock (private communication).en_US
dc.identifier.citedreferenceE. B. Wilson, Jr., J. C. Decius, and P. C. Cross, Molecular Vibrations (McGraw‐Hill Book Co., New York, 1955).en_US
dc.identifier.citedreferenceIn order to remove any ambiguity about the meaning of force constants we shall encounter later, we present, in the usual notation, the explicit form of several representative symmetry coordinates, or S3a=2−1∕2(Δr5−Δr6),S4a=8−1∕2re(Δα15+Δα25+Δα35+Δα45−Δα16−Δα26−Δα36−Δα46),S5a=(12re)(−Δα12+Δα23−Δα34+Δα14). Except for the numbering scheme these coordinates are those of C. W. F. T. Pistorius, J. Chem. Phys. 29, 1328 (1959). For identification of subscripts see Figs. 1 and 2.en_US
dc.identifier.citedreferenceY. Morino, S. J. Cyvin, K. Kuchitsu, and T. Iijima, J. Chem. Phys. 36, 1109 (1962).en_US
dc.identifier.citedreferenceK. Kuchitsu and L. S. Bartell, J. Chem. Phys. 36, 2460, 2470 (1962).en_US
dc.identifier.citedreferenceH. H. Claassen, Bull. Am. Phys. Soc. 12, 295 (1967); (private communication).en_US
dc.identifier.citedreferenceValue represents an extrapolation of F(qFF)F(qFF) values given by Shimanouchi et al. to the F‐F distance in XeF6XeF6 at OhOh symmetry, [T. Shimanouchi, I. Nakagawa, J. Hiraishi, and M. Ishii, J. Mol. Spectry. 19, 78 (1966)], and agree with (∂2V/∂q2)(∂2V∕∂q2) for Ne‐Ne interactions at the same distance [J. O. Hirschfelder, C. F. Curtiss, and R. B. Bird, Molecular Theory of Gases and Liquids (John Wiley & Sons, Inc., New York, 1964)].en_US
dc.identifier.citedreferenceT. M. Dunn, D. S. McClure, and R. G. Pearson, Crystal Field Theory (Harper and Row Publishers, Inc., New, York, 1965).en_US
dc.identifier.citedreferenceE. B. Wilson. Tr. (Drivate communication).en_US
dc.identifier.citedreferenceL. S. Bartell, J. Chem. Phys. 46, 4530 (1967).en_US
dc.identifier.citedreferenceH. Kim, H. H. Claassen, and E. Pearson (private communication).en_US
dc.identifier.citedreferenceG. Herzberg, Infrared and Raman Spectra of Polyatomic Molecules (D. Van Nostrand Co., Inc., Princeton, N.J., 1945), p. 264.en_US
dc.identifier.citedreferenceH. H. Claassen (private communication).en_US
dc.identifier.citedreferenceJ. C. Hindman and A. Svirmicks, Ref. 1, p. 251.en_US
dc.identifier.citedreferenceJ. Martins and E. B. Wilson, Jr., as quoted in Ref. 25.en_US
dc.identifier.citedreferenceD. S. Urch, J. Chem. Soc. 1964, 1442.en_US
dc.identifier.citedreferenceProvided spin‐orbit coupling did not stabilize OhOh symmetry. H. A. Jahn and E. Teller, Proc. Roy. Soc. (London) A161, 220 (1937). H. A. Jahn, A164, 117 (1938); A168, 469, 495 (1938).en_US
dc.identifier.citedreferenceR. F. Code, W. E. Falconer, W. Klemperer, and I. Ozier, J. Chem. Phys. 47, 4955 (1967).en_US
dc.identifier.citedreferenceF. Schreiner, D. W. Osborne, J. G. Malm, and G. MacDonald, Chem. Eng. News 44, 64 (1966).en_US
dc.identifier.citedreferenceP. A. Agron, C. K. Johnson, and H. A. Levy, Inorg. Nucl. Chem. Letters 1, 145 (1965).en_US
dc.identifier.citedreferenceW. M. Tolles and W. D. Gwinn, J. Chem. Phys. 36, 1119 (1962).en_US
dc.identifier.citedreferenceD. F. Smith, J. Chem. Phys. 21, 609 (1953).en_US
dc.identifier.citedreferenceL. S. Bartell, R. M. Gavin, Jr., H. B. Thompson, and C. L. Chernick, J. Chem. Phys. 43, 2547 (1965).en_US
dc.identifier.citedreferenceL. S. Bartell, Inorg. Chem. 5, 1635 (1966).en_US
dc.identifier.citedreferenceL. L. Lohr, Jr., and W. N. Lipscomb, Ref. 1, p. 347.en_US
dc.identifier.citedreferenceR. Hoffmann, J. Chem. Phys. 39, 1397 (1963).en_US
dc.identifier.citedreferenceL. L. Lohr, Bull. Am. Phys. Soc. 12, 295 (1967).en_US
dc.identifier.citedreferenceR. M. Gavin, Jr., and L. S. Bartell (unpublished).en_US
dc.identifier.citedreferenceU. Öpik and M. H. L. Pryce, Proc. Roy. Soc. (London) A238, 425 (1957); D. H. W. DenBoer, P. C. DenBoer, and H. C. Longuet‐Higgins, Mol. Phys. 5, 387 (1962); B. J. Nicholson and H. C. Longuet‐Higgins, 9, 461 (1965).en_US
dc.identifier.citedreferenceL. S. Bartell, Symposium on Models for Discussion of Molecular Geometry, American Chemical Society Meeting, Chicago, Ill., September 1967; J. Chem. Educ. (to be published).en_US
dc.identifier.citedreferenceR. F. W. Bader, Mol. Phys. 3, 137 (1960).en_US
dc.identifier.citedreferenceR. D. Willett, Theoret. Chim. Acta 6, 186 (1966).en_US
dc.identifier.citedreferenceG. Nagarajan, Bull. Soc. Chem. Belg. 71, 674 (1962).en_US
dc.identifier.citedreferenceG. Engle, Z. Krist. 90, 341 (1935); E. E. Aynsley and A. C. Hazell Chem. Ind. (London) 1963 611; I. D. Brown, Can. J. Chem. 42, 2758 (1964); S. Lawton and R. Jacobson (private communication, 1965).en_US
dc.identifier.citedreferenceR. E. Rundle, Record Chem. Progr. (kresge‐Hooker Sci. Lib.) 23, 195 (1962) A slight modification in the scheme is introduced by R. E. Rundle, Survey Progr. Chem. 1, 81 (1963).en_US
dc.identifier.citedreferenceA treatment in which the weighting of configurations is consistent with the potential function coupling the modes is in progress.en_US
dc.identifier.citedreferenceH. Kim, Bull. Am. Phys. Soc. 13, 425 (1968).en_US
dc.owningcollnamePhysics, Department of


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.